Flux
In physics and mathematics, flux is a measure of the amount of a physical quantity that flows through a given surface per unit time, often quantified as the surface integral of a vector field. The concept is fundamental in describing phenomena such as the flow of fluids, heat, mass, electric fields, and magnetic fields.[1] The term derives from the Latin fluxus, meaning "flow," and is widely applied in electromagnetism (e.g., electric and magnetic flux), transport phenomena, and vector calculus, where it relates to theorems like the divergence theorem.[2]Terminology and Fundamentals
Terminology
The term flux derives from the Latin fluxus, meaning "a flow," which stems from the verb fluere, "to flow." It entered English in the late 14th century, initially denoting an abnormally copious discharge, such as in medical descriptions of fluid motion or bodily flows, before broadening to scientific contexts by the 16th century.[3][4] In scientific usage, flux refers to the rate at which a quantity—such as mass, energy, particles, or a field—passes through a given surface, typically measured per unit time and often normalized per unit area to indicate intensity. This concept captures the transfer or flow across a boundary, emphasizing the net amount crossing rather than motion parallel to the surface. Flux is always defined with respect to the direction perpendicular to the surface, as only the normal component contributes to the passage through it; oblique flows reduce the effective flux via the cosine of the angle between the flow direction and the surface normal.[5][6][7] A key distinction exists between total flux, which integrates the flow over the entire surface area to yield the overall quantity transferred, and flux density, which quantifies the flow per unit area at a local point, providing a measure of concentration or strength. In physics, flux commonly describes phenomena like particle streams in transport processes or energy radiation, while in mathematics, it denotes the surface integral of a vector field over a region. Beyond these domains, the term appears briefly in non-physical fields, such as economics, where "cash flux" or flux analysis refers to the movement or variation in financial quantities like revenue streams.[8][9][7][10]Basic Properties and Units
Flux, in its general physical sense, represents the rate of flow of a quantity through a surface, and its density form—often simply called flux—carries dimensions of that quantity per unit time per unit area.[11] This dimensional structure arises from the fundamental definition of flux as a surface integral of a flow vector field, ensuring homogeneity in physical equations.[12] For instance, mass flux density has dimensions of mass per time per area, expressed as kg/(s·m²) in SI units. In the International System of Units (SI), common flux densities are derived from base units without special names in most cases, promoting coherence across physical domains. Energy flux density, or irradiance, is measured in watts per square meter (W/m²), equivalent to joules per second per square meter (J/(s·m²)).[13] Molar flux density uses moles per second per square meter (mol/(s·m²)), reflecting the flow of substance amount.[14] These units stem directly from the SI base units of mass (kg), time (s), length (m), and amount of substance (mol).[15] While SI emphasizes coherence, historical non-SI units persist in specialized contexts, such as luminous flux measured in lumens (lm), which is the SI unit but derived from candela (cd) and steradian (sr) as cd·sr.[16] Luminous flux density, or illuminance, is then in lux (lx), or lm/m², bridging photometry with SI standards.[17] These units maintain compatibility with SI base quantities like luminous intensity, underscoring the system's adaptability without sacrificing uniformity.[13] A key invariant property of flux is its adherence to conservation laws, encapsulated in the continuity equation, which states that the divergence of the flux vector equals the negative rate of change of the quantity density plus any sources or sinks.[18] This relation holds across diverse physical systems, from mass and charge to energy, ensuring local conservation without net creation or destruction in the absence of sources.[19] Flux is inherently signed, with its value positive when the flow vector aligns with the chosen surface normal and negative when opposing it, providing directional information essential for vector calculus applications.[20] This convention, often taking the outward normal for closed surfaces, standardizes calculations and reflects the oriented nature of flow through surfaces.[21]Flux as Flow Rate per Unit Area
General Mathematical Definition
In transport phenomena, the flux density \vec{J} represents the rate at which a physical quantity Q (such as mass, energy, or momentum) flows through a unit area perpendicular to the flow direction per unit time. It is formally defined as the limit of the incremental flow rate divided by the area, as the area approaches zero: \vec{J} = \lim_{\Delta A \to 0} \frac{\Delta Q}{\Delta t \, \Delta A} \hat{n}, where \Delta Q is the amount of quantity transported across the small area \Delta A in time \Delta t, and \hat{n} is the unit normal vector to the surface defining the direction of the flux. The vector nature of \vec{J} ensures it points in the direction normal to the surface, capturing both magnitude and orientation of the flow. For mass transport specifically, the mass flux density \vec{j} arises from the product of the local mass density \rho and the velocity field \vec{v}: \vec{j} = \rho \vec{v}. This expression describes the convective transport of mass, where the flux density is proportional to the density and the component of velocity aligned with the surface normal. To obtain the total flux \Phi through an arbitrary surface S, the flux density is integrated over the surface: \Phi = \int_S \vec{J} \cdot d\vec{A}, where d\vec{A} = \hat{n} \, dA is the vector area element; this continuous form derives from summing the discrete contributions in the limit of infinitesimal areas. Flux densities exhibit key mathematical properties, including linearity with respect to the transported quantity and additivity across different transport mechanisms. For instance, the total flux density can be decomposed as the sum of a convective component \rho \vec{v} and a diffusive component, allowing superposition for complex flows without loss of generality. A representative example is the flux of particles through a flat plane of area A. If the particles have number density n and average velocity \vec{v}, the total particle flux is \Phi = n (\vec{v} \cdot \hat{n}) A, where the dot product \vec{v} \cdot \hat{n} projects the velocity onto the surface normal, yielding zero flux for perpendicular flows and maximum for aligned ones; this illustrates how the flux density \vec{j}_p = n \vec{v} quantifies the effective crossing rate per unit area.Transport Phenomena Applications
In transport phenomena, the concept of flux as a flow rate per unit area finds extensive application in describing the movement of mass, momentum, and energy through continuous media. Momentum transport, governed by viscous effects, is exemplified by Newton's law of viscosity, which states that the shear stress \tau_{yx} (flux of x-momentum across a y-oriented plane) is \tau_{yx} = -\mu \frac{\partial u_x}{\partial y}, where \mu is the dynamic viscosity and u_x is the x-component of velocity; this describes frictional forces in fluid flows, such as in pipe friction or boundary layers.[22] Diffusion represents a fundamental diffusive transport process where particles move from regions of higher concentration to lower concentration due to random molecular motion, driven by thermal agitation and collisions. This molecular basis underpins Fick's first law, formulated by Adolf Fick in 1855, which quantifies the diffusive flux \vec{J} of a species as proportional to the negative gradient of its concentration c: \vec{J} = -D \nabla c, where D is the diffusion coefficient, a material-specific property reflecting the ease of particle movement.[23][24] Convective transport, in contrast, arises from the bulk motion of the fluid carrying the species along with it, often dominating in scenarios with significant fluid velocities. The convective flux \vec{J}_c is given by \vec{J}_c = c \vec{v}, where \vec{v} is the fluid velocity vector. In many practical situations, such as in chemical reactors or atmospheric flows, the total mass flux \vec{J} combines both mechanisms: \vec{J} = c \vec{v} - D \nabla c, allowing for the modeling of advection-diffusion processes where convection enhances diffusive spreading.[25][26] Heat transfer by conduction similarly employs flux to describe energy flow, as captured by Fourier's law, proposed by Joseph Fourier in 1822. This law states that the heat flux vector \vec{q} is proportional to the negative temperature gradient \nabla T: \vec{q} = -k \nabla T, with k denoting the thermal conductivity, which quantifies a material's ability to conduct heat through lattice vibrations or free electron movement in solids and fluids.[27] Applications of these flux expressions extend to specialized scenarios, such as gas effusion through porous media under Knudsen conditions, where the mean free path exceeds the pore diameter, leading to molecule-wall collisions dominating transport and yielding a flux \vec{J}_K = -\frac{1}{3} \bar{v} d_p \nabla c (with \bar{v} as the average molecular speed and d_p the pore diameter); this is critical in vacuum systems and membrane separations.[28] Another example is sedimentation flux in fluids, where suspended particles settle under gravity, producing a downward flux J_s = c v_s (with v_s as the settling velocity, often derived from Stokes' law for low Reynolds numbers). This process governs sediment deposition in rivers and industrial clarifiers, influencing erosion patterns and water quality. In chemical engineering, flux concepts apply to transport-limited reactions, particularly molar flux through membranes, where selective permeation drives separation processes like gas purification or desalination; here, the flux N_i of species i follows forms analogous to Fick's law, modulated by membrane permeability and pressure gradients.[29][30] While traditional treatments emphasize bulk continua, flux in biological systems—such as ion flux across cell membranes via channels—highlights selective transport down electrochemical gradients, enabling processes like nerve signaling with fluxes on the order of $10^7 ions per second per channel. Modern extensions include microfluidics, where engineered channels manipulate tiny fluid volumes to study or mimic these fluxes, as in lab-on-a-chip devices for drug delivery, achieving precise control over concentration gradients in volumes below microliters.[31][32]Flux as Surface Integral
Mathematical Formulation
In multivariable calculus, the flux of a vector field \vec{F} through a surface S is defined as the surface integral \Phi = \iint_S \vec{F} \cdot d\vec{A}, where d\vec{A} represents the vector area element, given by d\vec{A} = \hat{n} \, dA with \hat{n} as the unit normal vector to the surface and dA as the scalar area element.[33] This integral quantifies the net "flow" of the field across the surface, weighted by the field's component perpendicular to the surface. The orientation of the surface plays a crucial role in determining the direction of \hat{n}. For an oriented surface, the normal vector is chosen consistently, often following the right-hand rule: if fingers curl along the boundary curve in the positive direction, the thumb points in the direction of \hat{n}.[34] This distinguishes oriented flux, which can be positive or negative depending on the alignment of \vec{F} with \hat{n}, from absolute flux, which considers only the magnitude |\vec{F} \cdot \hat{n}| \, dA.[33] For a parametrized surface \vec{r}(u,v) = \langle x(u,v), y(u,v), z(u,v) \rangle over a domain D in the uv-plane, the flux integral becomes \Phi = \iint_D \vec{F}(\vec{r}(u,v)) \cdot (\vec{r}_u \times \vec{r}_v) \, du \, dv, where \vec{r}_u \times \vec{r}_v provides the normal vector aligned with the parametrization's orientation.[33] This form facilitates computation for non-flat surfaces by transforming the integral into a double integral over D. A key property of flux through a closed surface S enclosing a volume V is its relation to the divergence of the field via the divergence theorem: \iint_S \vec{F} \cdot d\vec{A} = \iiint_V \nabla \cdot \vec{F} \, dV, linking surface flux to the field's local expansion or contraction within the volume.[35] As an illustrative example, consider the flux of a radial velocity field \vec{v} = v \hat{r} (with constant speed v) through a sphere of radius r centered at the origin, oriented outward. Parametrizing the sphere in spherical coordinates and computing the integral yields \iint \vec{v} \cdot d\vec{A} = 4\pi r^2 v, representing the total outward flow across the surface.[36]Relation to Divergence Theorem
The divergence theorem establishes a fundamental connection between the flux of a vector field through a closed surface and the sources or sinks within the enclosed volume. For a vector field \vec{F} that is continuously differentiable on an open set containing the solid region V with piecewise smooth boundary surface S (oriented outward), the theorem states: \iint_S \vec{F} \cdot d\vec{A} = \iiint_V \nabla \cdot \vec{F} \, dV This equates the net outward flux through S to the integral of the divergence over V./16%3A_Vector_Calculus/16.08%3A_The_Divergence_Theorem) A proof can be sketched by considering the theorem separately for each Cartesian component of \vec{F}, say \vec{F} = (P, Q, R), and proving the result for \vec{F} = (P, 0, 0) using coordinate parametrization. By Fubini's theorem, the volume integral \iiint_V \frac{\partial P}{\partial x} \, dV reduces to integrating over slices perpendicular to the x-axis, where the fundamental theorem of calculus applied to each slice yields the surface integral of P over the x-faces, with internal contributions canceling. The cases for Q and R follow analogously, and this approach generalizes the one-dimensional fundamental theorem, akin to Stokes' theorem in higher dimensions./04%3A_Line_and_Surface_Integrals/4.04%3A_Surface_Integrals_and_the_Divergence_Theorem) The theorem implies that if \nabla \cdot \vec{F} = 0 everywhere in V, the net flux through S is zero, characterizing divergence-free (solenoidal) fields, such as those in incompressible fluid flow with no sources or sinks. Conversely, nonzero divergence indicates local expansion (sources, \nabla \cdot \vec{F} > 0) or contraction (sinks, \nabla \cdot \vec{F} < 0)./04%3A_Integral_Theorems/4.02%3A_The_Divergence_Theorem) In applications, the divergence theorem underpins conservation laws, such as the continuity equation for mass in fluid dynamics: \frac{\partial \rho}{\partial t} + \nabla \cdot \vec{j} = 0, where \rho is density and \vec{j} = \rho \vec{v} is mass flux density. Integrating over V and applying the theorem yields \frac{d}{dt} \iiint_V \rho \, dV = - \iint_S \vec{j} \cdot d\vec{A}, expressing the rate of change of mass inside V as the negative of the outward flux through S.[37] As an example, the theorem previews Gauss's law in potential theory, where for a field like the electric field, the flux through a closed Gaussian surface equals the integral of the divergence (related to enclosed sources) over the volume, quantifying how field lines originate from or terminate at interior points without specifying the physical context./16%3A_Vector_Calculus/16.08%3A_The_Divergence_Theorem)Applications in Electromagnetism
Electric Flux
Electric flux, denoted \Phi_E, quantifies the flow of the electric field \vec{E} through a given surface S and is mathematically defined as the surface integral \Phi_E = \iint_S \vec{E} \cdot d\vec{A}, where d\vec{A} is the infinitesimal vector area element pointing outward from the surface.[38] This dot product accounts for the component of the electric field perpendicular to the surface, making the flux positive when field lines exit the surface and negative when they enter.[39] The SI unit of electric flux is the volt-meter (V·m), which is dimensionally equivalent to the newton-meter squared per coulomb (N·m²/C), reflecting the field's intensity integrated over area.[40] Conceptually, electric flux represents the density of electric field lines piercing the surface; denser field lines correspond to higher flux, and the net flux through a closed surface is zero if no charge is enclosed, as field lines entering must equal those exiting.[41] A fundamental relation is provided by Gauss's law, which states that the total electric flux \Phi_E through any closed surface is equal to the net charge Q_{encl} enclosed by that surface divided by the vacuum permittivity \epsilon_0: \Phi_E = \oint_S \vec{E} \cdot d\vec{A} = \frac{Q_{encl}}{\epsilon_0}. This law enables the calculation of electric fields in symmetric charge distributions by choosing appropriate Gaussian surfaces.[39] Notably, the net flux through a closed surface is a Lorentz invariant quantity, as it directly corresponds to the enclosed charge, which remains unchanged under Lorentz transformations in special relativity.[42] For example, consider a point charge q at the center of a spherical Gaussian surface of radius r. By symmetry, \vec{E} is radial and constant in magnitude on the sphere, so \Phi_E = E \cdot 4\pi r^2 = q / \epsilon_0, yielding the Coulomb field E = q / (4\pi \epsilon_0 r^2).[38] Similarly, for an infinite parallel-plate capacitor with uniform surface charge density \sigma on one plate, a Gaussian pillbox straddling the plate gives \Phi_E = E \cdot A = (\sigma A) / \epsilon_0, so the field between plates is E = \sigma / \epsilon_0, independent of plate separation.[43]Magnetic Flux
Magnetic flux, denoted \Phi_B, quantifies the total magnetic field passing through a given surface and is defined as the surface integral of the magnetic field vector \vec{B} dotted with the infinitesimal area vector d\vec{A} over the surface S: \Phi_B = \iint_S \vec{B} \cdot d\vec{A}. The SI unit of magnetic flux is the weber (Wb), equivalent to one tesla-square meter (T·m²)./22:_Induction_AC_Circuits_and_Electrical_Technologies/22.1:_Magnetic_Flux_Induction_and_Faradays_Law) This measure arises from the interaction of magnetic fields with conducting circuits, where flux linkage determines electromagnetic induction effects. A key property of magnetic flux stems from Maxwell's equation \nabla \cdot \vec{B} = 0, which implies that the net magnetic flux through any closed surface is always zero.[44] This reflects the nonexistence of magnetic monopoles, as magnetic field lines form continuous loops rather than originating or terminating at isolated points, in contrast to electric flux, which can be nonzero through closed surfaces due to enclosed charges.[45] In electromagnetism, changing magnetic flux through a circuit induces an electromotive force (EMF), as described by Faraday's law of induction: the magnitude of the induced EMF \mathcal{E} equals the negative time derivative of the flux, \mathcal{E} = -\frac{d\Phi_B}{dt}.[46] For a loop or coil enclosing the surface, this law links flux variation—due to moving magnets, changing currents, or varying fields—to generated voltages, forming the basis for transformers, generators, and inductors. Lenz's law complements this by specifying that the induced current flows in a direction opposing the flux change, conserving energy by resisting the alteration.[45] Consider a solenoid, a helical coil of wire carrying current I, which produces a uniform magnetic field B = \mu_0 n I inside, where n is the turn density and \mu_0 the permeability of free space.[47] The flux through each turn of area A is \Phi_B = B A, and for N turns, the total linkage is N \Phi_B; if current changes, the induced EMF opposes this via Lenz's law, generating a field that counters the flux variation. Similarly, in a toroidal coil—a solenoid bent into a doughnut shape—the confined field enhances flux uniformity, minimizing leakage and illustrating induction in closed magnetic paths./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/13:_Electromagnetic_Induction/13.03:_Lenz's_Law) In modern applications, particularly type-II superconductors, magnetic flux penetrating the material forms quantized vortices known as fluxoids, where the flux is discrete in units of the magnetic flux quantum \Phi_0 = h / (2e) \approx 2.07 \times 10^{-15} Wb, with [h](/page/H+) Planck's constant and [e](/page/Elementary_charge) the elementary charge.[48] This quantization, predicted by Fritz London and experimentally confirmed in multiply-connected superconductors, underpins phenomena like the Josephson effect and enables precise magnetic sensing in devices such as SQUIDs (superconducting quantum interference devices).[49]Specialized Flux Concepts
Poynting Flux
The Poynting vector \vec{S} quantifies the flux of electromagnetic energy, representing the power flow per unit area in an electromagnetic field. Introduced by John Henry Poynting in his 1884 paper on energy transfer in electromagnetic fields, it is defined in SI units as \vec{S} = \frac{1}{\mu_0} \vec{E} \times \vec{B}, where \vec{E} is the electric field strength, \vec{B} is the magnetic flux density, and \mu_0 is the vacuum permeability, with units of watts per square meter (W/m²).[50][51] The direction of \vec{S} indicates the instantaneous direction of energy propagation, while its magnitude S = \frac{EB}{\mu_0} \sin \theta gives the energy flux density, with \theta the angle between \vec{E} and \vec{B}; in typical cases like plane waves where fields are perpendicular, \sin \theta = 1.[51] This interpretation arises from the cross-product structure, which aligns energy flow orthogonal to both fields, as derived from Maxwell's equations.[50] Poynting's theorem provides a conservation law linking this flux to energy changes: \nabla \cdot \vec{S} + \frac{\partial u}{\partial t} = -\vec{j} \cdot \vec{E}, where u = \frac{1}{2} \left( \epsilon_0 E^2 + \frac{B^2}{\mu_0} \right) is the electromagnetic energy density and \vec{j} is the current density, showing that the decrease in field energy corresponds to work done on charges or energy outflow.[50] In practical applications, the Poynting vector calculates energy transport in waveguides, where \vec{S} aligns with the propagation direction to determine power handling capacity. For antennas, integrating the time-averaged Poynting vector over a surrounding surface yields the total radiated power, essential for efficiency analysis.[52] In modern relativistic contexts, such as high-intensity laser interactions with plasmas, the Poynting vector describes energy flux driving particle acceleration and wave propagation in dense media.[53] A representative example occurs in a monochromatic plane wave propagating in free space, where the time-averaged Poynting vector magnitude is \langle S \rangle = \frac{E_0^2}{2 Z_0}, with E_0 the electric field amplitude and Z_0 = \sqrt{\mu_0 / \epsilon_0} \approx 377 \, \Omega the impedance of free space; this intensity scales quadratically with field strength, highlighting energy transport in electromagnetic radiation.[54]Radiometric Flux and SI Units
In radiometry, the term "flux" primarily refers to radiant flux, which quantifies the total power of electromagnetic radiation emitted, reflected, transmitted, or received by an object or passing through a surface. Radiant flux, denoted as \Phi_e, is defined as the time derivative of the radiant energy Q_e, expressed mathematically as \Phi_e = \frac{dQ_e}{dt}. This quantity represents the rate at which radiant energy flows, independent of direction or spatial distribution, making it the foundational measure in detector-based radiometry for characterizing light sources./Remote_Sensing_(Knudby)/01%3A_Chapters/1.02%3A_Radiometric_measurements)[55] The SI unit for radiant flux is the watt (W), equivalent to one joule per second (J/s), directly tying it to the base SI unit for power. This unit applies to broadband measurements across the electromagnetic spectrum relevant to radiometry, typically from ultraviolet through infrared wavelengths. For instance, the output power of a lamp or the total energy flux from a star is expressed in watts, providing a scalar measure without regard to beam geometry or area. The adoption of the watt ensures consistency with electrical power measurements, facilitating integration of radiometric data in engineering and scientific applications.[56][57] Spectral radiant flux extends the concept to wavelength- or frequency-resolved measurements, denoted as \Phi_{e,\lambda} (per unit wavelength) or \Phi_{e,\nu} (per unit frequency). These are defined as \Phi_{e,\lambda} = \frac{d\Phi_e}{d\lambda} and \Phi_{e,\nu} = \frac{d\Phi_e}{d\nu}, respectively, allowing analysis of radiation distribution across the spectrum. The SI units are watts per meter (W/m) for \Phi_{e,\lambda} and watts per hertz (W/Hz) for \Phi_{e,\nu}, enabling precise characterization of sources like LEDs or blackbody radiators where spectral content varies significantly. Integrating spectral flux over the appropriate range yields the total radiant flux, underscoring its role in comprehensive radiometric assessments.[58]| Quantity | Symbol | Definition | SI Unit |
|---|---|---|---|
| Radiant flux | \Phi_e | Total power of radiation | W |
| Spectral radiant flux (wavelength) | \Phi_{e,\lambda} | Power per unit wavelength | W/m |
| Spectral radiant flux (frequency) | \Phi_{e,\nu} | Power per unit frequency | W/Hz |