The volt (symbol: V) is the derived unit in the International System of Units (SI) for measuring electric potential difference, also known as voltage, as well as electromotive force.[1] It is named after the Italian physicist Alessandro Volta (1745–1827), who invented the voltaic pile in 1800, the first device capable of producing a steady electric current through chemical reactions.[2][3]One volt represents the electric potential difference between two points in a conducting wire carrying a constant current of one ampere when the power dissipated between those points is one watt.[2] Equivalently, it is the amount of energy, one joule, required to move one coulomb of electric charge between the points.[4] In terms of SI base units, the volt is expressed as kilogram meter squared per second cubed per ampere, or kg⋅m²⋅s⁻³⋅A⁻¹.[4] This unit was formally adopted as part of the SI in 1960 through CGPM Resolution 12, building on earlier definitions established in 1948.[1]The volt plays a central role in electrical engineering, physics, and everyday technology, quantifying the "push" that drives electric current through circuits and devices ranging from household appliances to advanced electronics.[2] Since the 2019 SI revision, which fixed the values of fundamental constants like the elementary charge and speed of light, the volt's realization has become more precise through quantum standards such as the Josephson effect, enabling highly accurate voltage measurements independent of material artifacts.[5] Common multiples include the kilovolt (kV) for high-voltage transmission lines and the millivolt (mV) for low-level signals in sensors.[4]
Definition and Fundamentals
Formal SI Definition
The volt (symbol: V) is the SI unit of electric potential difference (voltage). It is defined as the electric potential difference between two points of a conducting wire carrying a constant current of 1 ampere when the power dissipated between these points is equal to 1 watt.[6]This definition derives from the fundamental relation for electric power, P = V I, where P is power in watts, V is potential difference in volts, and I is current in amperes, yielding $1 \, \mathrm{V} = 1 \, \mathrm{W} / 1 \, \mathrm{A}.[6] It connects to resistance through Ohm's law, V = I R, where R is resistance in ohms, such that power dissipation P = I^2 R aligns with the unit definition when I = 1 \, \mathrm{A} and P = 1 \, \mathrm{W}.[7]The definition was proposed by the International Committee for Weights and Measures (CIPM) in 1946 and formally adopted by the 9th General Conference on Weights and Measures (CGPM) in 1948 as part of the electrical units in the metre-kilogram-second (MKS) system.[6] It was incorporated into the SI upon its establishment in 1960 and remains unchanged in the 2019 revision of the SI, which redefined base units via fixed constants but preserved the relational definitions of derived units like the volt.[7] The volt is a derived SI unit, expressed in terms of the base units kilogram, metre, second, and ampere as \mathrm{V} = \mathrm{kg \cdot m^2 \cdot s^{-3} \cdot A^{-1}}.[6]
Relation to Base Units
The dimensional formula for the volt in the International System of Units (SI) is [V] = kg·m²·s⁻³·A⁻¹, derived from the relationship between electric potential difference and work per unit charge, where one volt is defined as the potential difference across a conductor when one joule of energy is dissipated per coulomb of charge transferred.[4][8] This formula arises because the joule (J) has dimensions kg·m²·s⁻² and the coulomb (C) has dimensions A·s, so dividing the former by the latter yields the volt's dimensions.[9]The volt depends on four of the seven SI base units: mass (kg), length (m), time (s), and electric current (A), with no involvement of temperature (K), amount of substance (mol), or luminous intensity (cd).[8] This composition reflects the volt's role as a derived unit in electromagnetism, tying mechanical and electrical quantities together through fundamental interactions.[4]The 2019 redefinition of the SI units fixed the value of the elementary charge e to exactly 1.602176634 × 10⁻¹⁹ coulombs, which anchors the ampere—and by extension the volt—to a reproducible quantum constant rather than physical artifacts like the international prototype kilogram.[10][11] This change ensures that the volt's realization is exact and independent of material standards, enhancing precision in metrology.[12]For instance, one volt corresponds to one joule per coulomb, or in base units, the energy of 1 kg·m²·s⁻² imparted to a charge of 1 A·s, illustrating the unit's expression as a potential energy gradient per unit charge in fundamental terms.[4][8]
Physical Interpretation
Connection to Energy and Charge
The volt represents the electric potential difference between two points in an electric field, defined as the amount of work required to move a unit positive charge from one point to the other. Specifically, one volt is equivalent to one joule of energy expended per coulomb of charge transported, expressed as V = \frac{J}{C}. This relationship underscores the volt as a measure of energy per unit charge, where the potential difference quantifies the electrostatic work done against the field.[7][13]In practical terms, this energy-charge linkage is evident in the acceleration of charged particles. For instance, a potential difference of one volt imparts one electronvolt (eV) of kinetic energy to a single electron, whose charge is approximately $1.6 \times 10^{-19} C, as the electron gains energy equivalent to qV where q is its charge. This unit, the eV, is widely used in atomic and particle physics to describe energy scales on the order of electron interactions.[14]For a uniform electric field, the potential difference V relates directly to the field strength E and the distance d between the points, given by V = Ed. This equation illustrates how voltage arises from the spatial variation of the electric field, providing the driving force for charge separation in capacitors or conductors.[15]In electrical circuits, voltage serves as the motive force that drives current I through a conductor, with the instantaneous power delivered or dissipated expressed as P = VI. This formulation connects voltage to the rate of energy transfer, where higher voltages enable greater power output for a given current, fundamental to the operation of devices from batteries to amplifiers.[16]
Distinction from Related Concepts
The volt (V) is the SI derived unit for electric potential difference, a physical quantity often interchangeably called voltage in technical and colloquial usage. While voltage denotes the magnitude of the potential difference between two points in an electric field, the volt specifically quantifies this difference as the work done per unit charge to move a charge between those points, equivalent to one joule per coulomb. This distinction ensures precision in measurement, as voltage encompasses the broader concept, but only the volt serves as the standardized unit in the International System of Units (SI).[17]A key differentiation exists between the volt and electromotive force (EMF), both expressed in volts but representing distinct aspects of electrical systems. EMF refers to the maximum potential difference a source, such as a battery or generator, can provide when no current flows through an external circuit, essentially the open-circuit voltage generated by the source's internal mechanisms. In contrast, the volt measures the actual potential difference across any two points in a circuit, which may be lower than the EMF due to internal resistance or load effects when current is drawn. This clarifies that while EMF drives the circuit, the volt applies universally to any observed potential drop.[18][13]The volt quantifies potential difference rather than absolute electric potential, which is a scalar field value at a single point relative to an arbitrary reference (often infinity or ground). Electric potential itself lacks a fixed zero point and is meaningful only through differences, so the volt always describes the change in potential (ΔV) between locations, not an absolute value. This relativity underscores that potentials are path-independent in electrostatic fields, but the unit remains the volt for the difference.[19]Common misconceptions arise with alternating current (AC) and direct current (DC) voltages, both denominated in volts yet interpreted differently for effective power delivery. DC voltages are straightforward steady-state values, whereas AC voltages are typically specified as root-mean-square (RMS) values, which equal the DC equivalent for the same heating effect in a resistor—approximately 0.707 times the peak value for sinusoidal waveforms. Peak or instantaneous AC values, while also in volts, do not represent the effective voltage for most practical calculations, leading to errors if not distinguished from RMS.[20]
Analogies for Understanding
Hydraulic Analogy
The hydraulic analogy likens electrical circuits to systems of water flowing through pipes, offering an intuitive framework for grasping the concept of voltage. In this model, voltage represents the pressure difference—measured in pascals for water but conceptualized in volts for electricity—that drives the flow of charges, much like hydrostatic pressure propels water through a conduit. A higher voltage provides a greater "push" on electrons, analogous to increased waterpressure overcoming gravitational or frictional forces to maintain flow.[21] This analogy emphasizes voltage as the potential difference across a circuit, motivating charge movement without directly representing the charges themselves.[22]Central to the analogy are the mappings between electrical quantities and hydraulic counterparts. Electric current, measured in amperes, corresponds to the volume flow rate of water, quantifying how much charge passes a point per unit time, similar to gallons per second in a pipe. Resistance, in ohms, parallels the frictional resistance or constriction in the pipe, such as narrower diameter or rougher surfaces that impede flow for a given pressure. Power, in watts, equates to the hydraulic power delivered, akin to the work rate of waterflow (pressure times flow rate), illustrating how voltage and current together determine energy transfer.[23] These correspondences align with fundamental circuit laws, like Ohm's law (current equals voltage divided by resistance), mirroring Poiseuille's law for fluid dynamics.[21]A practical example is a 12 V battery in a simple circuit with a resistor, where the battery acts as a pressure source equivalent to a 12-unit head of water, forcing current through the resistor like flow through a narrow, friction-filled pipe. The resulting current depends on the resistor's "narrowness," just as pipe restrictions limit water throughput under fixed pressure.While effective for direct current (DC) circuits with ohmic components, the hydraulic analogy has limitations, particularly in capturing alternating current (AC) dynamics or non-linear behaviors, where fluid flow does not naturally replicate oscillatory pressures or reactive elements like capacitors and inductors.[24] It remains particularly valuable, however, for building conceptual understanding of voltage's role in steady-state DC scenarios.[25]
Mechanical and Thermal Analogies
The mechanical analogy for voltage draws a parallel between electric potential difference and gravitational potential difference in a uniform field. Just as the gravitational potential energy of a mass increases with height in a gravitational field, the electric potential energy of a charge increases as it moves against an electric field; thus, voltage, defined as the potential energy per unit charge, is analogous to the height difference per unit mass, or gh where g is gravitational acceleration and h is height.[26] Specifically, a voltage of 1 V corresponds to an electric potential energy change of 1 J for a charge of 1 C, equivalent in energy scale to the gravitational potential energy gained by lifting a mass that yields 1 J when raised by a certain height under gravity.[27] In this analogy, positive charges "fall" from regions of high electric potential to low potential, converting potential energy into kinetic energy, much like objects rolling downhill in a gravitational field and accelerating due to the height difference.[28]The thermal analogy compares voltage to a temperature gradient that drives heat flow, particularly through the Seebeck effect in thermoelectrics. In the Seebeck effect, a temperature difference \Delta T across a junction of dissimilar materials generates an electromotive force (voltage) proportional to \Delta T, where the Seebeck coefficient determines the voltage per unit temperature difference.[29] This voltage arises as charge carriers diffuse from the hot to the cold region, creating a potential difference that opposes further flow, analogous to how a temperature gradient induces heat transfer until equilibrium.[30] For instance, in a thermocouple, the generated voltage can power a circuit when the temperature difference is maintained, highlighting the direct link between thermal disequilibrium and electrical potential.[31]These analogies underscore the conservative nature of electrostatic fields, where the work done by the electric field on a charge between two points is path-independent, depending only on the potential difference (voltage) at those points.[32] This path independence mirrors the behavior in gravitational and thermal conservative systems, emphasizing that voltage represents a state function tied to energy differences rather than the specific trajectory.[33] As noted in the connection between voltage and energy per charge, such analogies aid in conceptualizing how electric potential drives charge motion without dissipation along closed paths.[26]
Measurement and Standards
Historical Measurement Methods
In the late 18th century, early measurements of electric potential relied on electrometers, devices that detected voltage through the repulsion of charged objects. The gold-leaf electrometer, invented by Abraham Bennet in 1787, consisted of two thin gold leaves suspended from a conducting rod within a glass enclosure; when a voltage was applied, the leaves diverged proportionally to the potential difference due to electrostatic repulsion, allowing qualitative and semi-quantitative assessments.[34][35] This instrument marked a significant advance over prior pith-ball electroscopes, providing greater sensitivity for measuring potentials in electrostatic experiments, though it required careful calibration and was limited to low-current, high-impedance scenarios.[34]By the mid-19th century, as electrochemical cells emerged, the Daniell cell served as a key reference for absolute voltage measurements. Invented by John Frederic Daniell in 1836, this cell produced a stable electromotive force of approximately 1.1 volts from the reaction between zinc and copper electrodes in their respective sulfate solutions, separated by a porous barrier to minimize polarization.[36][37] In the 1860s, during early efforts to define electrical standards, the Daniell cell's output was adopted by bodies like the British Association for the Advancement of Science as a practical benchmark for the volt, enabling more reproducible comparisons in telegraphy and laboratory work.[38][37]Into the early 20th century, galvanometers paired with potentiometers offered precise voltage comparisons via null methods. The slide-wire potentiometer, originating from Johann Christian Poggendorff's 1841 design and refined in commercial forms by the 1900s, used a uniform resistance wire along which a sliding contact balanced an unknown voltage against a known electromotive force from a standard cell, with a sensitive galvanometer detecting zero current at the balance point.[39] This technique, employed in instruments like those from Leeds & Northrup, achieved accuracies to within millivolts without drawing significant current from the source, making it ideal for emf measurements in research and industry.[39]These historical methods faced substantial challenges from environmental and material inconsistencies, hindering precision. Temperature variations notably affected cell outputs, with the Daniell cell's voltage decreasing by about 0.0004 volts per degree Celsius rise due to shifts in electrolyte properties and reactionkinetics. Material impurities and inconsistencies in wire resistances for potentiometers further introduced errors, while electrometers suffered from humidity-induced charge leakage. No unified international standard existed until the 1881 International Electrical Congress in Paris, which formalized the volt and related units to address these discrepancies.[40]
Modern Josephson Junction Standard
The modern realization of the volt relies on the Josephson effect, predicted by Brian Josephson in 1962, which describes a direct current-voltage relationship in a superconducting tunnel junction under microwave irradiation.[41] In this effect, a constant DC voltage V appears across the junction at specific steps given by the equationV = \frac{n h f}{2 e},where n is an integer representing the step number, h is Planck's constant, e is the elementary charge, and f is the frequency of the applied microwave radiation.[41] This quantum phenomenon enables precise voltage generation tied to measurable frequency, independent of material properties beyond the fundamental constants.[42]Since 1972, national metrology institutes have implemented Josephson voltage standards based on this effect to maintain the volt, starting with systems at the National Bureau of Standards (now NIST) that disseminated the U.S. legal volt using 10 mV arrays.[43] Prior to the 2019 SI revision, the conventional Josephson constant K_{J-90} = 2e/[h](/page/H+) was fixed at 483597.9 GHz/V, such that 1 V corresponded to a frequency of approximately 483.598 GHz times an integer n.[44] Following the redefinition of the SI units, which fixed the values of h and e, the Josephson constant became exactly K_J = 483597.8484\ldots GHz/V, allowing direct realization of the volt from frequency measurements alone, with frequency serving as the sole input parameter traceable to the second.[42]These standards offer exceptional reproducibility, achieving relative uncertainties as low as $10^{-10} or better in inter-laboratory comparisons, far surpassing classical methods due to the intrinsic quantum locking of voltage steps.[45] National metrology laboratories, including NIST and PTB, employ them as primary references for calibrating voltage artifacts and ensuring global consistency in electrical measurements.[41]To generate practical voltages beyond millivolts, series arrays of thousands to hundreds of thousands of Josephson junctions are integrated into superconducting chips, operated at cryogenic temperatures around 4 K under microwave excitation.[46] For instance, modern programmable Josephson voltage standards (PJVS) use over 265,000 junctions to produce stable DC outputs up to 10 V, enabling rapid switching and application in both DC and AC metrology.[46] This array configuration ensures Shapiro steps remain sharp and uniform, supporting uncertainties below 1 part in $10^{10} for routine use.[47]As of August 2025, NIST has developed a single integrated device that realizes the SI standards for voltage (via Josephson effect), resistance (via quantum Hall effect), and current (via single-electron transport) within one cryostat, improving efficiency and precision in quantum electrical metrology.[48]
Common Applications and Values
Everyday and Household Voltages
In households around the world, alternating current (AC) mains power provides the primary electrical supply for appliances and lighting. In the United States and Canada, the nominal root mean square (RMS) voltage is 120 V at 60 Hz, with standard tolerances allowing variations from 114 V to 126 V under normal operating conditions, though brief fluctuations up to ±10% can occur due to grid dynamics.[49] In much of Europe and other regions following International Electrotechnical Commission (IEC) standards, the nominal voltage is 230 V at 50 Hz, with tolerances typically limited to -6% to +10%, ensuring compatibility with connected devices.[50] These voltages originate from large-scale generators in power plants and are stepped down through transformers for safe distribution to homes.Direct current (DC) voltages are prevalent in portable and low-power household applications, often derived from batteries or rectified AC sources. A typical AA alkaline battery delivers a nominal 1.5 V, suitable for powering small devices like remote controls and clocks.[51] Automotive lead-acid batteries, used for starting engines and auxiliary systems, operate at a standard 12 V when fully charged.[52] Universal Serial Bus (USB) ports, common on computers, chargers, and wall adapters, supply a regulated 5 V, enabling the operation of peripherals such as external drives and charging cables.[53]Many consumer electronics incorporate even lower voltages for efficient operation. Smartphones and similar devices charge via USB protocols that adjust from 5 V for standard rates up to 20 V for fast charging, optimizing power delivery while minimizing heat.[54] Light-emitting diodes (LEDs), widely used in indicators, displays, and lighting strips, require forward voltages of 1.8 V to 3.3 V depending on the diode's color and material, such as lower values for red LEDs and higher for white or blue ones.[55]From a safety perspective, everyday voltages are designed to minimize risks, but thresholds exist for potential electric shock. Voltages above 50 V RMS for AC or 120 V for DC are classified as hazardous under IEC guidelines, as they can drive sufficient current through the body to cause injury, though the actual danger also hinges on factors like current limiting, skin resistance, and exposure pathway.[56] Protective measures, such as grounding and fuses in household systems, further mitigate these risks.
Voltages in Scientific and Industrial Contexts
In high-voltage power transmission, voltages typically range from 110 kV to 765 kV to enable efficient long-distance electricity delivery while minimizing resistive losses described by the I²R term in power dissipation.[57] These elevated potentials reduce the required current for a given power level, as power P equals voltage V times current I, thereby lowering energy losses in conductors over hundreds of kilometers.[58] Such systems form the backbone of national grids, with examples including 500 kV lines in the United States for bulk power transfer.[59]Scientific applications often demand extreme voltages to accelerate charged particles for probing fundamental physics. In the Large Hadron Collider (LHC) at CERN, superconducting radiofrequency cavities provide accelerating voltages up to 2 MV per cavity, totaling about 16 MV per beam to achieve proton energies of 6.5 TeV.[60] Van de Graaff generators, electrostatic accelerators pivotal in nuclear physics research, can generate terminal voltages reaching 25 MV, enabling ion acceleration for experiments on atomic nuclei and material properties.[61] Linear accelerators in facilities like SLAC extend this to gigavolt scales, with total accelerating potentials around 50 GV over kilometer-long structures to produce high-energy electron beams for X-ray studies.[62] These voltages, often applied in pulsed modes, facilitate collisions that reveal subatomic structures but require precise insulation to prevent breakdowns.At the opposite end of the scale, precision scientific contexts utilize ultra-low voltages for sensitive measurements and control. In quantum computing, superconducting qubits are manipulated using gate voltages on the order of a few microvolts (μV) to tune Josephson junctions without introducing thermalnoise that could decohere quantum states.[63] In biomedical research, electroencephalography (EEG) records brain electrical activity at scalp voltages of 10–100 μV, capturing neural oscillations for diagnosing neurological disorders.[64] These minute signals demand high-impedance amplifiers to preserve fidelity, highlighting voltage's role in non-invasive physiological monitoring.[65]Industrial processes leverage specific voltage ranges tailored to equipment demands, often prioritizing current alongside potential. Electric vehicles employ direct current (DC) battery packs at 400–800 V, predominantly 800 V as of 2025, to drive high-power motors efficiently, enabling rapid acceleration and regenerative braking in models like those from Porsche and Hyundai.[66][67]Arc welding operations maintain arcs at 20–50 V, where low voltage ensures operator safety while high currents (up to 1000 A) generate the necessary heat for metal fusion in manufacturing.[68] These applications underscore voltage optimization for thermal and mechanical performance in heavy industry.
History and Development
Origins and Naming
The unit of electric potential known as the volt is named after Alessandro Volta (1745–1827), an Italianphysicist and chemist renowned for his pioneering work in electricity.[69] In 1800, Volta invented the voltaic pile, the first electrochemical battery capable of producing a steady electric current through chemical reactions between alternating discs of zinc and copper separated by brine-soaked cardboard. This breakthrough demonstrated that electricity could be generated continuously from chemical sources, laying foundational concepts for understanding electromotive force and potential difference.[70]Early efforts to quantify electrical potential predated Volta's contributions. In the 1780s, ItalianphysicistGiovanni Battista de Cavallo developed an improved electrometer—a device using gold leaves to detect and measure electrical charges—that scaled readings in "degrees of electricity" to assess the intensity of static electrical tension.[71] While de Cavallo's instrument advanced the measurement of transient electrical effects, it was Volta's voltaic pile that provided a reliable source of sustained current, enabling more precise standardization of potential concepts in subsequent experiments.The formal adoption of "volt" as the name for the unit of electromotive force occurred in 1881 at the International Electrical Congress in Paris, where international delegates honored Volta's legacy by selecting the term to denote the difference of potential that drives one ampere of current against one ohm of resistance.[72] This decision marked a key step in unifying electrical nomenclature amid growing global standardization efforts. Initially, the volt's practical definition relied on the electromotive force of the Clark cell—a mercury-zinc electrochemical reference cell yielding approximately 1.4328 volts at 15 °C—as a reproducible standard for calibration.[73]
Evolution as an SI Unit
In the mid-20th century, the volt underwent formal standardization through international agreements. At the 9th General Conference on Weights and Measures (CGPM) in 1948, the volt was implicitly defined via the newly ratified international ampere and international ohm, where the international ampere was set as the current producing a specific force between conductors, and the international ohm was maintained by an artifact prototype resistance of 1 ohm; this system relied on physical prototypes for reproducibility across nations.[74] By the 1960s, a shift occurred toward absolute units independent of artifacts. The 11th CGPM in 1960 established the International System of Units (SI), designating the ampere as a base unit defined by the force between two parallel conductors carrying constant current, thereby deriving the volt absolutely from the meter, kilogram, and second without prototype dependence.[75][76]From the 1970s to the 1990s, quantum effects revolutionized the precise realization of the volt, reducing uncertainties from parts per million to parts per billion. The AC Josephson effect, discovered in 1962, was internationally adopted in 1972 as a practical voltage standard, linking voltage to frequency via the relation V = \frac{n f h}{2e}, where n is an integer, f is microwave frequency, h is Planck's constant, and e is the elementary charge; this enabled stable, reproducible voltages using superconducting junctions.[77][78] Similarly, the quantum Hall effect, observed in 1980, provided a resistance standard that complemented Josephson-based voltage measurements for full electrical metrology. In 1988, the International Committee for Weights and Measures (CIPM) adopted conventional values for the Josephson constant K_{J-90} = 483597.9 GHz/V and the von Klitzing constant R_{K-90} = 25812.807 Ω, effective from 1 January 1990, fixing the volt's representation relative to legacy prototypes while aligning quantum realizations globally; this adjustment shifted national standards by small amounts, such as 9.3 ppm for the U.S. volt.[79][80]The 2019 revision of the SI marked a pivotal advancement, defining the volt exactly through fundamental constants without reliance on artifacts or conventional values. At the 26th CGPM in 2018, effective 20 May 2019, the ampere was redefined by fixing the elementary charge e = 1.602176634 \times 10^{-19} C, with the second unchanged; the volt, as the unit of electric potential difference, derives from the watt (kg m² s⁻³) divided by the ampere, ultimately anchored to fixed values of the Planck constant h, elementary charge e, and speed of light c. This change maintained the numerical value of the volt but enhanced universality and stability, as realizations now stem directly from invariant physical constants rather than measurable phenomena or prototypes.[81][6]The Bureau International des Poids et Mesures (BIPM) plays a central role in harmonizing volt realizations worldwide, conducting key comparisons of Josephson voltage standards among national metrology institutes to ensure consistency at the parts-per-billion level. Through ongoing calibrations and resolutions, such as those under the Consultative Committee for Electricity and Magnetism, the BIPM maintains the SI volt's integrity, supporting global trade, scientific collaboration, and technological interoperability.[82][83]