Fact-checked by Grok 2 weeks ago

Continuity equation

The continuity equation is a fundamental partial differential equation in physics that expresses the local conservation of mass (or other conserved quantities, such as electric charge) within a continuous medium, stating that the temporal rate of change of a density field in a given volume equals the negative of the flux divergence through the bounding surface. In its general form, for a conserved quantity with density \rho and flux \mathbf{J}, it is written as \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{J} = 0, where the first term represents local accumulation or depletion, and the second term accounts for net transport across boundaries. This equation arises from applying the integral form of conservation laws to an infinitesimal control volume and is a cornerstone of continuum mechanics, ensuring no sources or sinks exist except through explicit terms. In , the continuity equation specifically governs mass conservation, taking the form \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0, where \rho is fluid density and \mathbf{v} is ; for incompressible flows, it simplifies to \nabla \cdot \mathbf{v} = 0, implying divergence-free velocity fields. It is one of the core Navier-Stokes equations, alongside and , and plays a critical role in modeling phenomena like , where horizontal convergence drives vertical motion, or flows, where it constrains mass flow rates through varying cross-sections. Derivations typically involve balancing mass influx and outflux across a fixed , leading to both Eulerian (fixed-frame) and (material) formulations. In , the continuity equation manifests as , \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{j} = 0, where \rho is and \mathbf{j} is ; this relation is mathematically derived from by taking the divergence of Ampère's law with Maxwell's correction and using . It ensures that charge is neither created nor destroyed locally, underpinning circuit analysis (via Kirchhoff's current law) and relativistic electrodynamics, where it extends to four-vector forms in . Beyond these domains, analogous continuity equations appear in (for probability density), , and , highlighting its versatility as a local expression of global conservation principles.

Mathematical Foundations

Flux Concept

In physics and mathematics, flux represents the rate at which a quantity, such as , , or charge, passes through a surface, quantifying the transport across a per unit time. It arises from the idea of a describing the flow , where the flux measures the net amount crossing the surface in the direction normal to it. For instance, describes the movement of matter in , given by the product of and , while charge flux corresponds to electric current in . This concept captures the directional flow of conserved quantities, essential for understanding conservation laws. Mathematically, the flux of a vector field \mathbf{F}, known as the flux density, through an oriented surface S is defined as the surface integral \Phi = \iint_S \mathbf{F} \cdot d\mathbf{A}, where d\mathbf{A} is the vector area element, with magnitude equal to the infinitesimal area and direction normal to the surface. This integral computes the component of \mathbf{F} perpendicular to S, weighted by the surface element, yielding a scalar value that can be positive (outward flow) or negative (inward flow) depending on the orientation. The formulation originates from the need to generalize line integrals to surfaces in vector analysis. While is fundamentally a scalar representing , the associated flux \mathbf{F} is a , highlighting a key distinction in physical contexts: scalar flux often denotes integrated quantities like across a boundary, whereas vector flux emphasizes directional , such as momentum flux in mechanics. The term "flux" and its rigorous mathematical treatment emerged in 19th-century developments in vector calculus, particularly through Carl Friedrich Gauss's work on the divergence theorem in the 1830s, building on earlier ideas by figures like Joseph-Louis Lagrange and Mikhail Ostrogradsky.

Integral Formulation

The integral formulation of the continuity equation expresses the of a , such as or charge, within an arbitrary fixed V bounded by a closed surface S. Consider a quantity with \rho(\mathbf{r}, t) distributed throughout the volume, where the flux density of this quantity is given by the \mathbf{j}(\mathbf{r}, t). The total amount of the quantity within V at time t is \int_V \rho \, dV. The rate of change of this total amount is \frac{d}{dt} \int_V \rho \, dV, accounting for any temporal variation in \rho. Simultaneously, the net rate at which the quantity leaves the volume through S is the surface of the , \int_S \mathbf{j} \cdot d\mathbf{A}, where d\mathbf{A} is the outward-pointing area element. For a with no sources or sinks inside V, the principle of requires that the time rate of change within the volume equals the negative of the net outward , leading to the integral continuity equation: \frac{d}{dt} \int_V \rho \, dV + \int_S \mathbf{j} \cdot d\mathbf{A} = 0. This equation serves as a global statement of , applicable to any fixed, arbitrary V, regardless of its shape or size, as long as the S is closed and the fields \rho and \mathbf{j} are sufficiently smooth. It encapsulates the balance between the accumulation (or depletion) of the quantity inside V and the across its boundary, embodying the idea that what enters or leaves the volume directly affects its internal content. The formulation is particularly useful in and physics applications where global balances are computed, such as in analysis for systems with complex geometries, without requiring detailed local field information. To connect the integral form to its differential counterpart, apply the , which relates the surface integral of a to a of its : \int_S \mathbf{j} \cdot d\mathbf{A} = \int_V \nabla \cdot \mathbf{j} \, dV. Substituting this into the continuity equation yields \frac{d}{dt} \int_V \rho \, dV + \int_V \nabla \cdot \mathbf{j} \, dV = 0, or equivalently, \int_V \left( \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{j} \right) dV = 0 under suitable conditions on the time derivative. Since this holds for any arbitrary volume V, the integrand must vanish pointwise, establishing the link to the local . This proof highlights how the global conservation implies a local balance everywhere in space. In practical applications to closed systems, the integral form simplifies significantly. For instance, in incompressible flow within an with no or outlet (a closed ), the \rho is , and the \mathbf{j} = \rho \mathbf{v} where \mathbf{v} is the . The equation then reduces to \int_S \mathbf{v} \cdot d\mathbf{A} = 0, implying zero net volume flux across the , which ensures the volume of remains over time. This is commonly applied in analyzing rigid containers or sealed pipes under steady conditions, where it confirms the absence of net mass accumulation. Another example arises in electrostatics for charge conservation in an isolated region, where \rho is charge and \mathbf{j} is current ; for a closed conductor, the equation verifies that total charge is preserved if no external currents flow.

Differential Formulation

The differential formulation of the continuity equation expresses the local of a in continuous , relating the time of change of to the spatial of its at every point. This form is obtained by applying mathematical theorems to the global statement, assuming the underlying fields are smooth enough for such localization. The standard differential form is given by \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{j} = 0, where \rho(\mathbf{x}, t) represents the of the at position \mathbf{x} and time t, and \mathbf{j}(\mathbf{x}, t) is the associated . This enforces balance locally: any temporal increase in must be offset by a net outflow of through the surrounding space. To derive this from the integral formulation over a fixed V with \partial V, begin with the statement \frac{d}{dt} \int_V \rho \, dV + \oint_{\partial V} \mathbf{j} \cdot d\mathbf{A} = 0, where d\mathbf{A} is the outward-pointing area element. For a fixed , the Leibniz rule allows interchanging the time and : \int_V \frac{\partial \rho}{\partial t} \, dV + \oint_{\partial V} \mathbf{j} \cdot d\mathbf{A} = 0. The surface integral is then converted to a volume integral via the : \int_V \frac{\partial \rho}{\partial t} \, dV + \int_V \nabla \cdot \mathbf{j} \, dV = 0, yielding \int_V \left( \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{j} \right) dV = 0. Since this equality holds for any arbitrary fixed volume V, the integrand vanishes pointwise, resulting in the . This derivation relies on key assumptions: the density \rho and flux \mathbf{j} must be continuous functions within V and sufficiently differentiable (at least once continuously differentiable) to justify the application of the divergence theorem and Leibniz rule. These conditions ensure the integrals exist and the transition from global to local conservation is rigorous, excluding singularities or discontinuities in the fields. For steady-state scenarios, where the system does not evolve temporally (\frac{\partial \rho}{\partial t} = 0), the equation reduces to \nabla \cdot \mathbf{j} = 0, implying solenoidal with no net sources or sinks. In the case of incompressible media, where \rho is constant in space and time, and if the takes the convective form \mathbf{j} = \rho \mathbf{v} with \mathbf{v}, the equation simplifies further to \nabla \cdot \mathbf{v} = 0, indicating volume-preserving .

Conservation Principles

The continuity equation finds a profound physical interpretation through Noether's theorem, which establishes a deep connection between symmetries in physical laws and conservation principles. In 1918, Emmy Noether published her seminal paper "Invariante Variationsprobleme," demonstrating that every differentiable symmetry of the action of a physical system leads to a corresponding conservation law. This theorem, originally developed in the context of general relativity to address conservation laws in gravitational fields, applies broadly to Lagrangian field theories, revealing that the continuity equation mathematically encodes these conserved quantities as the local manifestation of underlying symmetries. Noether's theorem states that if the action integral is invariant under continuous transformations—such as rotations, translations, or scaling—then there exists a conserved current associated with that symmetry. The conserved current j^\mu satisfies a continuity equation of the form \partial_\mu j^\mu = 0, where \partial_\mu denotes the with respect to coordinates, and the equation holds in four-dimensional Minkowski (or its curved generalizations). This divergence-free condition implies global of the charge Q = \int j^0 \, d^3x over space, provided suitable boundary conditions are met, linking local flux balance to global invariants. The theorem thus positions the continuity equation not merely as an empirical statement of balance but as a necessary consequence of principles in variational formulations of physics. A canonical example arises from spacetime translation invariance, which corresponds to the homogeneity of and time in physical laws. This symmetry yields the energy-momentum tensor T^{\mu\nu} as the , satisfying \partial_\mu T^{\mu\nu} = 0, thereby conserving total energy and momentum in isolated systems. Such applications underscore how Noether's framework unifies diverse continuity equations across physics, from particle number conservation in quantum fields to mass preservation in fluids, all rooted in the invariance of the fundamental action.

Derivation from Symmetries

The derivation of the continuity equation from symmetries in Lagrangian field theory proceeds through Noether's theorem, which associates continuous symmetries of the action with conserved currents whose divergence vanishes on-shell. The procedure begins by identifying an infinitesimal symmetry transformation of the coordinates x^\mu \to x^\mu + \xi^\mu \epsilon and the fields \phi \to \phi + \delta \phi, where \epsilon is an infinitesimal parameter, \xi^\mu describes spacetime transformations, and \delta \phi includes both the intrinsic field variation and the Lie drag term \xi^\nu \partial_\nu \phi. For the action S = \int d^4x \, L(\phi, \partial \phi) to be invariant under this transformation (up to a boundary term), the variation of the Lagrangian must satisfy \delta L = \partial_\mu K^\mu for some K^\mu, ensuring \delta S = 0 when evaluated on solutions to the Euler-Lagrange equations. Computing the variation explicitly yields \delta L = \frac{\partial L}{\partial \phi} \delta \phi + \frac{\partial L}{\partial (\partial_\mu \phi)} \partial_\mu (\delta \phi) + L \partial_\mu \xi^\mu. Integrating by parts and using the Euler-Lagrange equations \frac{\partial L}{\partial \phi} - \partial_\mu \left( \frac{\partial L}{\partial (\partial_\mu \phi)} \right) = 0 on-shell, the off-shell variation of the action simplifies to a total divergence: \delta S = \int d^4x \, \partial_\mu \left[ \frac{\partial L}{\partial (\partial_\mu \phi)} \delta \phi - \xi^\mu L - K^\mu \right]. Thus, the symmetry implies the existence of a j^\mu = \frac{\partial L}{\partial (\partial_\mu \phi)} \delta \phi - \xi^\mu L - K^\mu, satisfying the \partial_\mu j^\mu = 0 when the fields obey the equations of motion. In the common case of internal symmetries where \xi^\mu = 0 and K^\mu = 0, this reduces to j^\mu = \frac{\partial L}{\partial (\partial_\mu \phi)} \delta \phi. A representative general expression for the Noether current in field theories is j^\mu = \frac{\partial L}{\partial (\partial_\mu \phi)} \delta \phi - \theta^\mu, where \theta^\mu encapsulates the contributions from spacetime transformations and any additional terms like \xi^\mu L + K^\mu. This current's four-divergence being zero encodes the local form of the derived from the . As a , consider time-translation invariance, where \xi^\mu = (\epsilon, 0, 0, 0) and \delta \phi = -\epsilon \partial_t \phi, assuming the Lagrangian is time-independent so K^\mu = 0. The resulting current components include j^0 = \mathcal{H} (the ) and spatial components forming the , with the full tensor T^\mu{}_\nu = \frac{\partial L}{\partial (\partial_\mu \phi)} \partial_\nu \phi - \delta^\mu_\nu L satisfying \partial_\mu T^\mu{}_\nu = 0, which is the continuity equation for energy-momentum. Noether's procedure applies primarily to local theories with a variational principle, such as those invariant under Lorentz transformations, where the symmetries are continuous and the action is differentiable; it does not hold for non-variational formulations or global topological constraints.

Classical Physics Applications

Fluid Dynamics

In fluid dynamics, the continuity equation expresses the principle of mass conservation for a fluid, stating that the rate of change of mass within a volume equals the net mass flux across its boundaries. This is particularly relevant in Eulerian coordinates, where the fluid is analyzed from a fixed reference frame, with properties like density varying as functions of position and time. The fluid-specific form of the equation is derived by considering an infinitesimal control volume and applying the conservation of mass: the time rate of change of density inside the volume balances the divergence of the mass flux. For a fluid with mass density \rho and velocity field \mathbf{v}, this yields the partial differential equation \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0, where the first term represents local storage changes and the second term accounts for advective transport. This derivation assumes no sources or sinks of mass, such as chemical reactions, and holds for both compressible and incompressible flows. In the incompressible limit, where density \rho is constant (valid for liquids or low-speed gases), the equation simplifies significantly: the partial derivative of \rho with respect to time vanishes, leaving \nabla \cdot \mathbf{v} = 0. This divergence-free condition implies that the fluid velocity field is solenoidal, meaning volume is preserved under flow, which is crucial for simplifying simulations in engineering applications like pipe flows. Practical applications of the continuity equation abound in modeling natural and engineered fluid systems. For instance, in river flow modeling, the equation underpins one-dimensional approximations like the Saint-Venant continuity form \frac{\partial A}{\partial t} + \frac{\partial Q}{\partial x} = 0, where A is the cross-sectional area and Q is the discharge, enabling predictions of flood propagation and water resource management. In compressible flows, such as those involving shock waves in high-speed gases, the equation enforces mass balance across discontinuities: for a normal shock, \rho_1 u_1 = \rho_2 u_2, where subscripts denote upstream and downstream states, linking density jumps to velocity changes in supersonic flows like those in jet engines or blast waves.

Electromagnetism

In , the continuity equation embodies the principle of conservation, stating that the rate of change of within a volume equals the negative of the density flowing out of that volume. This is mathematically expressed as \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{J} = 0, where \rho is the and \mathbf{J} is the . This local form ensures that charge cannot be created or destroyed arbitrarily, reflecting a fundamental derived from experimental observations and theoretical consistency in electromagnetic theory. The equation arises directly from , specifically through the interplay between for electricity and the Ampère-Maxwell law. states \nabla \cdot \mathbf{E} = \rho / \varepsilon_0, where \mathbf{E} is the and \varepsilon_0 is the . Taking the time derivative yields \nabla \cdot (\partial \mathbf{E}/\partial t) = (1/\varepsilon_0) \partial \rho / \partial t. The Ampère-Maxwell law is \nabla \times \mathbf{B} = \mu_0 \mathbf{J} + \mu_0 \varepsilon_0 \partial \mathbf{E}/\partial t, where \mathbf{B} is the and \mu_0 is the . Applying the operator to this equation gives zero on the left (due to the divergence of a being zero) and \mu_0 \nabla \cdot \mathbf{J} + \mu_0 \varepsilon_0 \nabla \cdot (\partial \mathbf{E}/\partial t) = 0, which simplifies to \partial \rho / \partial t + \nabla \cdot \mathbf{J} = 0 upon substitution and rearrangement. This derivation demonstrates that is not an independent postulate but a consequence of Maxwell's framework, ensuring consistency across the equations. Historically, the foundations trace to André-Marie Ampère's work in the 1820s, where he formulated the circuital law relating to steady electric currents, \oint \mathbf{B} \cdot d\mathbf{l} = \mu_0 I, based on experiments with current-carrying wires. However, Ampère's law alone was inconsistent with for time-varying fields, as its divergence implied \nabla \cdot \mathbf{J} = 0 even when charges could accumulate. James Clerk Maxwell resolved this in his 1865 paper by introducing the displacement current term \mu_0 \varepsilon_0 \partial \mathbf{E}/\partial t into Ampère's law, making the full set of equations compatible with the continuity equation and enabling the prediction of electromagnetic waves. In practical implications, for steady currents where charge density does not change over time (\partial \rho / \partial t = 0), the continuity equation reduces to \nabla \cdot \mathbf{J} = 0, indicating that current lines neither begin nor end within the medium, akin to . This condition holds in conductors with uniform charge distribution, such as in circuit analysis. For electromagnetic waves in , where \mathbf{J} = 0, the equation becomes \partial \rho / \partial t = 0, ensuring no charge accumulation during propagation and maintaining the transverse nature of the fields, as the wave equations derived from set rely on this consistency.

Thermal and Statistical Applications

Heat and Energy Transfer

The continuity equation for and energy transfer embodies the conservation of within a medium, for conduction and as primary mechanisms of energy redistribution. It states that the local rate of change of balances the divergence of the vector plus any volumetric sources or sinks, expressed mathematically as \frac{\partial u}{\partial t} + \nabla \cdot \mathbf{q} = f, where u denotes the volumetric energy density (typically \rho c_p T, with \rho as density, c_p as at constant pressure, and T as ), \mathbf{q} is the vector, and f represents heat generation per unit volume, such as from chemical reactions or electrical dissipation. This formulation ensures no unaccounted creation or destruction of , fundamental to thermodynamic consistency in thermal systems. Fourier's law supplies the relation for conductive heat flux, \mathbf{q} = -k \nabla T, where k is the material's thermal conductivity, capturing how heat flows down the temperature gradient from hotter to cooler regions. Substituting this into the continuity equation produces the heat conduction equation \frac{\partial u}{\partial t} = \nabla \cdot (k \nabla T) + f, which, for constant properties and u = \rho c_p T, simplifies to the standard parabolic form \rho c_p \frac{\partial T}{\partial t} = \nabla \cdot (k \nabla T) + f. Convection enters when fluid motion advects , modifying \mathbf{q} to include a convective term \mathbf{q} = -k \nabla T + \rho c_p \mathbf{v} T, where \mathbf{v} is , though pure conduction dominates in . This equation derives directly from of applied to a fixed , which equates the time rate of change of total energy within the volume to the net rate of energy transfer across its boundaries plus any internal generation. For a V with surface S, the balance is \frac{d}{dt} \int_V u \, dV = -\oint_S \mathbf{q} \cdot d\mathbf{A} + \int_V f \, dV, neglecting mechanical work and assuming no across boundaries; 's core assertion—that \delta Q = dU + \delta W in differential form—manifests here as heat input driving energy accumulation when work is absent or balanced. Applying the and shrinking the volume to infinitesimal size yields the differential continuity equation, rigorously linking macroscopic thermodynamic principles to local transport behavior. In steady-state scenarios, where temperatures remain time-invariant (\frac{\partial u}{\partial t} = 0) and sources vanish (f = 0), the equation reduces to \nabla \cdot (k \nabla T) = 0, an elliptic solved for equilibrium temperature profiles under fixed boundary conditions like prescribed temperatures or fluxes. Transient cases retain the time derivative, modeling diffusive spreading of perturbations, often requiring numerical methods like finite differences for irregular geometries or variable k. For isotropic constant k, this becomes \nabla^2 T = 0 in steady state () or \frac{\partial T}{\partial t} = \alpha \nabla^2 T transiently, with \alpha = k / (\rho c_p) as . Building insulation models leverage the steady-state form to quantify loss through envelopes like walls, where low-k materials (e.g., with k \approx 0.03 W/m·K) minimize and ensure \nabla T across layers yields acceptable thermal resistance R = L/k per unit area. Solving \nabla \cdot (k \nabla T) = 0 for composite slabs predicts overall U-values (inverse of total R), guiding designs to cut use by up to 50% in temperate climates via optimized layering. In stellar interiors, the continuity equation governs transport in radiative zones, approximated as conductive with an effective k = \frac{1}{3} \bar{v} \ell \rho c_v from ( \ell, speed \bar{v}, specific heat c_v), enabling models of profiles where sources (f > 0) balance outward to sustain . The integral form aids in specifying boundary conditions, such as Neumann flux \mathbf{q} \cdot \mathbf{n} = h (T - T_\infty) for convective surfaces.

Probability Distributions

The Fokker-Planck equation provides a key application of the continuity equation to the dynamics of probability densities in stochastic processes, describing how the density p(\mathbf{x}, t) evolves under the influence of both systematic drift and random diffusive fluctuations. This takes the form \frac{\partial p}{\partial t} + \nabla \cdot (\mathbf{v} p - D \nabla p) = 0, where \mathbf{v} represents the drift velocity field capturing deterministic motion, and D is the diffusion tensor accounting for stochastic spreading. The term \mathbf{v} p - D \nabla p defines the probability current, analogous to the flux in classical continuity equations, ensuring that local changes in density arise solely from the divergence of this current. This structure enforces the conservation of total probability, with \int p(\mathbf{x}, t) \, d\mathbf{x} = 1 holding for all times, provided appropriate boundary conditions are met. In , which encompasses both position and momentum coordinates for the system, the Fokker-Planck equation interprets probability mass as an incompressible fluid conserved along trajectories, preventing creation or annihilation of probabilistic measure except through boundaries. This conservation principle extends the differential formulation's divergence term to settings, where randomness introduces diffusive corrections to the advective flow. The equation thus models the flow as a balance between drift-induced or and diffusion-induced spreading, maintaining the total probability. The Fokker-Planck framework originated in the study of , where in 1905 derived a as the continuum limit of random particle displacements due to molecular collisions, laying the groundwork for probabilistic transport descriptions. further refined this in 1906 by incorporating velocity-dependent friction, yielding the overdamped limit of the Fokker-Planck equation for position distributions in viscous media. These links highlight its role in processes, such as solute transport in fluids, where drift from external fields combines with random walks to evolve the density toward equilibrium distributions like the Boltzmann form. In , the Fokker-Planck equation models the evolution of trait or size distributions under birth-death rates and environmental variability, as explored in analyses of autocatalytic where amplifies or stabilizes variance. For instance, it captures how fluctuating selection pressures lead to broadening or narrowing of density profiles, conserving the total expected while quantifying uncertainty propagation. Similarly, in simulations, the equation serves as the governing continuum model for particle-based methods, enabling validation of discrete paths against analytical density evolutions, such as in test-particle approximations for . These applications underscore its utility in bridging microscopic to macroscopic probabilistic flows.

Imaging and Computational Applications

Computer Vision

In , the continuity equation manifests through the constraint, which models the conservation of image across consecutive frames to estimate motion. This approach treats the apparent motion of brightness patterns in image sequences as a field, analogous in structure to continuity principles in but applied to discrete perceptual data. The core equation, derived under the brightness constancy assumption, states that the I(x, y, t) at a point remains unchanged under small displacements, leading to: \frac{\partial I}{\partial t} + u \frac{\partial I}{\partial x} + v \frac{\partial I}{\partial y} = 0 where \frac{\partial I}{\partial t} is the temporal gradient, \frac{\partial I}{\partial x} and \frac{\partial I}{\partial y} are spatial gradients, and (u, v) represent the optical flow velocity components. This constraint enforces local conservation of intensity, assuming that brightness is constant along motion paths and that motions are small enough for linear approximations to hold. Solving the underconstrained optical flow equation requires additional regularization. The Horn-Schunck algorithm imposes a global smoothness constraint on the field, minimizing an functional that balances data fidelity to the constraint with a penalty on velocity gradients, typically solved iteratively via Euler-Lagrange equations. In contrast, the Lucas-Kanade method adopts a local approach, assuming constant flow within small spatial windows and solving the using to estimate velocities at feature points. These seminal differential methods, both introduced in 1981, address the problem—where motion perpendicular to edges is ambiguous—by incorporating spatial coherence, enabling robust estimation in grayscale image sequences. Optical flow continuity has broad applications in video processing, particularly for video stabilization, where estimated motion fields compensate for camera shake by warping frames to a stable reference path, reducing jitter in handheld footage. In object tracking for systems, it facilitates real-time motion prediction by propagating feature trajectories across frames, enhancing detection in dynamic scenes such as autonomous driving or . These techniques underpin modern pipelines, with extensions like pyramidal implementations handling larger displacements and methods (as of 2024) such as enabling end-to-end supervised or self-supervised estimation while preserving the underlying continuity principle.

Traffic and Network Flow

The continuity equation finds application in modeling as a macroscopic for vehicle , treating the stream of vehicles analogously to a compressible . In one-dimensional streams, the Lighthill-Whitham-Richards (LWR) model formulates this as \frac{\partial \rho}{\partial t} + \frac{\partial (\rho v)}{\partial x} = 0, where \rho(x,t) denotes the vehicle at position x and time t, and v(\rho) is the equilibrium speed as a decreasing of . This ensures that changes in arise solely from the of the \rho v, without sources or sinks, capturing the of kinematic along highways. In traffic networks, the continuity equation manifests at junctions and links as a balance of inflows, outflows, and accumulation. For a , the discrete form states that the inflow the of change in : \sum_{\text{in}} q_{\text{in}} - \sum_{\text{out}} q_{\text{out}} = \frac{dS}{dt}, where q represents flow and S is the length or density integral over the . This nodal condition, combined with link-level conservation laws, enables simulation of route choices and congestion spillover across interconnected roads. The LWR model predicts shock wave formation when traffic transitions from free-flow to congested states, such as behind a slowdown, creating discontinuities in density that propagate backward at speed \frac{q_2 - q_1}{\rho_2 - \rho_1}, where subscripts denote upstream and downstream states. These shocks represent abrupt queue buildups, with resolution requiring entropy conditions to select physically admissible solutions amid nonlinear wave steepening. Post-2000 developments have integrated the continuity equation into simulations for predicting congestion in heterogeneous networks, as in multi-class extensions of the LWR model that account for vehicle types and signals. Recent advances as of 2025 incorporate to enhance real-time predictions in complex scenarios. In , continuum-discrete hybrids apply laws to model flows across nodes, treating goods as with discontinuous fluxes at buffers to optimize against disruptions.

Quantum and Semiconductor Contexts

Quantum Mechanics

In quantum mechanics, the continuity equation expresses the conservation of probability for a non-relativistic particle described by the wave function \psi(\mathbf{r}, t). The probability density is defined as \rho = |\psi|^2, which gives the likelihood of detecting the particle at position \mathbf{r} at time t. The equation takes the form \frac{\partial \rho}{\partial t} + \nabla \cdot \mathbf{j} = 0, where \mathbf{j} is the probability current density, given by \mathbf{j} = \frac{\hbar}{2mi} \left( \psi^* \nabla \psi - \psi \nabla \psi^* \right). This form ensures that the total probability \int \rho \, dV over all space remains constant, reflecting the unitary evolution of the quantum state./01%3A_Introduction/1.04%3A_Continuity_Equation) The continuity equation arises directly from the time-dependent Schrödinger equation, i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where \hat{H} = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) is the Hamiltonian operator for a particle in potential V. To derive it, compute the time derivative of the density: \frac{\partial \rho}{\partial t} = \frac{\partial}{\partial t} (\psi^* \psi) = \psi^* \frac{\partial \psi}{\partial t} + \psi \frac{\partial \psi^*}{\partial t}. Substituting from the Schrödinger equation and its complex conjugate yields \frac{\partial \psi}{\partial t} = -\frac{i}{\hbar} \hat{H} \psi, \quad \frac{\partial \psi^*}{\partial t} = \frac{i}{\hbar} \hat{H}^* \psi^*. For a real potential V, the potential terms cancel, leaving the kinetic contribution: \frac{\partial \rho}{\partial t} = \frac{i \hbar}{2m} \left( \psi \nabla^2 \psi^* - \psi^* \nabla^2 \psi \right). Applying the vector identity \nabla \cdot (\psi^* \nabla \psi) = \psi^* \nabla^2 \psi + \nabla \psi^* \cdot \nabla \psi (and its conjugate) and rearranging gives \frac{\partial \rho}{\partial t} = -\nabla \cdot \mathbf{j}, confirming the continuity equation. This derivation holds for the standard non-relativistic case without magnetic fields or spin./01%3A_Introduction/1.04%3A_Continuity_Equation) The probability density \rho = |\psi|^2 interprets |\psi|^2 dV as the probability of finding the particle in volume dV, a statistical postulate introduced to resolve the physical meaning of the wave function. The continuity equation thus guarantees that probability is conserved, akin to mass or in classical systems, ensuring no probability is created or destroyed during evolution. This framework underpins particle detection probabilities in experiments, where repeated measurements yield outcomes distributed according to \rho. The association with phase invariance follows from , linking the global U(1) symmetry of the to probability conservation. In quantum tunneling through a potential barrier, the continuity equation maintains a constant across regions of classically forbidden penetration, where the wave function decays exponentially but the non-zero \mathbf{j} allows net probability , enabling transmission probabilities that defy classical . Similarly, in interference phenomena like the , superpositions of wave functions from multiple paths lead to oscillatory patterns in both \rho and \mathbf{j}, manifesting as modulated probability currents that constructively reinforce in bright fringes and destructively cancel in dark ones, highlighting the wave-like of probability.

Semiconductor Physics

In semiconductor physics, the continuity equation governs the conservation of charge carriers, accounting for their generation, recombination, and transport via drift and diffusion currents. For electrons, the equation is expressed as \frac{\partial n}{\partial t} = G_n - R_n + \frac{1}{q} \nabla \cdot \mathbf{J}_n, where n is the electron density, G_n is the generation rate, R_n is the recombination rate, q is the elementary charge, and \mathbf{J}_n is the electron current density. Similarly, for holes, \frac{\partial p}{\partial t} = G_p - R_p - \frac{1}{q} \nabla \cdot \mathbf{J}_p, with p the hole density, G_p and R_p the corresponding generation and recombination rates, and \mathbf{J}_p the hole current density. These equations arise from the balance of carrier influx and outflux across a differential volume, incorporating microscopic processes like thermal generation, radiative recombination, and Auger effects, which are prominent in doped materials under band theory. The derivation of these continuity equations stems from the Boltzmann transport equation within the framework of semiconductor band theory, where carriers occupy states in the conduction and valence bands. Starting from the , which describes the evolution of the carrier distribution function f(\mathbf{r}, \mathbf{k}, t) under band dispersion E(\mathbf{k}), moments are taken to obtain macroscopic densities and currents. Integrating over the yields the continuity form after applying the relaxation-time approximation for scattering, linking particle conservation to the divergence of the current \mathbf{J}_n = -q \int v(\mathbf{k}) f(\mathbf{k}) d\mathbf{k}, where v(\mathbf{k}) is the . This approach captures band-structure effects, such as effective masses in parabolic approximations, distinguishing it from classical fluids. Seminal work in the formalized these derivations for device modeling, emphasizing quasi-Fermi levels in non-equilibrium conditions. In steady-state conditions (\partial n / \partial t = 0), the continuity equations simplify to G_n - R_n + (1/q) \nabla \cdot \mathbf{J}_n = 0, enabling analytical solutions for device structures like p-n junctions. For a forward-biased p-n junction under low injection, the minority carrier (electron) density in the p-region decays exponentially as n_p(x) = n_{p0} + \Delta n_p(0) \exp(-x / L_n), where L_n = \sqrt{D_n \tau_n} is the diffusion length, D_n the diffusion coefficient, \tau_n the lifetime, and boundary conditions reflect injection at the junction edge. This solution, derived assuming quasi-neutrality and neglecting drift in the neutral base, predicts the diffusion current dominating saturation current. In solar cells, steady-state analysis of the continuity equation in the base region yields the short-circuit current density J_{sc} \propto q \int G(x) \exp(-x / L_n) dx, quantifying collection efficiency for photogenerated carriers, with examples in silicon cells showing L_n \approx 100-300 \, \mum under AM1.5 illumination. Numerical solutions to the coupled continuity, , and drift-diffusion equations became feasible in the 1970s with the advent of drift-diffusion models, which approximate \mathbf{J}_n = q \mu_n n \mathbf{E} + q D_n \nabla n using Einstein relations. Early advancements, such as Mock's steady-state analysis and Gummel's 1964 decoupling extended into multidimensional simulations by the late 1970s, enabled device optimization for bipolar transistors and diodes via finite-difference methods. By the 1980s, Scharfetter-Gummel stabilization improved accuracy for high-field regions, forming the basis of tools like Sentaurus TCAD, with ongoing refinements for submicron scales. These models prioritize computational efficiency over full Boltzmann solutions, establishing scale for carrier dynamics in integrated circuits.

Relativistic and Particle Extensions

Special Relativistic Form

In , the continuity equation for the conservation of a quantity such as charge, , or particle number takes a covariant form in four-dimensional Minkowski , ensuring Lorentz invariance. The conserved quantity is carried by the four-current j^\mu, a contravariant whose components are j^\mu = (\gamma \rho c, \gamma \rho \mathbf{v}), where \rho is the proper density in the fluid's , \gamma = (1 - v^2/c^2)^{-1/2} is the , \mathbf{v} is the three-velocity, and c is the . The relativistic continuity equation is then the four-divergence vanishing: \partial_\mu j^\mu = 0, where \partial_\mu = (\frac{1}{c} \partial_t, \nabla) in the mostly-minus (+, -, -, -). This equation expresses local conservation without sources or sinks. The four-current transforms as a under Lorentz boosts, preserving the structure of the equation. For a boost along the x-direction with velocity \beta c, the components mix time and space parts: j'^0 = \gamma_\beta (j^0 - \beta j^1) and j'^1 = \gamma_\beta (j^1 - \beta j^0), with transverse components unchanged, while \partial'_\mu transforms inversely to maintain the scalar nature of \partial_\mu j^\mu. This invariance arises because the four-divergence is a , guaranteeing that conservation holds equally in all inertial frames without explicit coordinate adjustments. In relativistic hydrodynamics, the continuity equation applies to the particle four-current N^\mu = n u^\mu, where n is the proper and u^\mu = \gamma (c, \mathbf{v}) is the satisfying u^\mu u_\mu = c^2; follows as \partial_\mu N^\mu = 0. For and , the analogous is the vanishing four-divergence of the energy-momentum tensor T^{\mu\nu}, \partial_\mu T^{\mu\nu} = 0, which for an ideal fluid is T^{\mu\nu} = (\varepsilon + p) u^\mu u^\nu / c^2 + p g^{\mu\nu}, with \varepsilon the proper and p the ; this tensor encodes both continuity-like mass and relativistic - flow. These equations, first systematically derived in the context of irreversible , form the basis for describing relativistic fluids such as those in high-energy astrophysics or heavy-ion collisions. In the non-relativistic limit where v \ll c, \gamma \approx 1 and the time component dominates, so j^0 \approx \rho c and the spatial part \mathbf{j} \approx \rho \mathbf{v}, reducing the equation to the three-dimensional form \partial_t \rho + \nabla \cdot (\rho \mathbf{v}) = 0, where \rho approximates the lab-frame density; higher-order relativistic corrections, such as and effects on density, vanish. This limit highlights how the relativistic form unifies space and time derivatives while recovering classical at low speeds.

General Relativistic Form

In , the continuity equation generalizes to the covariant conservation of the stress-energy tensor T^{\mu\nu}, expressed as \nabla_\mu T^{\mu\nu} = 0, where \nabla_\mu denotes the compatible with the g_{\mu\nu}. This form encapsulates the local and momentum in curved , accounting for gravitational effects through the connection terms in the . In the flat-space limit of , this reduces to the ordinary divergence of the flux being zero. The equation \nabla_\mu T^{\mu\nu} = 0 arises as a consequence of the (EFEs), G_{\mu\nu} = 8\pi G T_{\mu\nu}, where G_{\mu\nu} is the . The twice-contracted Bianchi identity, \nabla^\mu G_{\mu\nu} = 0, which holds identically due to the antisymmetry of the Riemann tensor, ensures the compatibility of the EFEs with stress-energy conservation. Thus, invariance of the gravitational action implies, via the Bianchi identities, that matter fields must satisfy this covariant continuity equation without additional assumptions. A prominent application occurs in cosmological models, such as the Friedmann-Lemaître-Robertson-Walker (FLRW) metric for a universe filled with dust (pressureless matter), where T^{\mu\nu} = \rho u^\mu u^\nu and \rho is the proper energy density. The \nu = 0 component of the continuity equation yields \dot{\rho} + 3H(\rho + p) = 0, with H = \dot{a}/a the Hubble parameter and p = 0 for dust, leading to \rho \propto a^{-3} and integrating into the Friedmann equations that govern cosmic expansion. This conservation law also reflects the gauge invariance of general relativity under diffeomorphisms, the smooth coordinate transformations preserving the metric structure. Noether's second theorem associates these spacetime symmetries with the vanishing covariant divergence of the stress-energy tensor, ensuring that local conservation holds without global conserved quantities in generic curved spacetimes.

Particle Physics Applications

In quantum field theory, the continuity equation expresses the local conservation of charges associated with global symmetries via Noether's theorem. For a global U(1) symmetry, such as phase invariance under \psi \to e^{i\alpha} \psi for a fermion field \psi, the Noether current j^\mu = \bar{\psi} \gamma^\mu \psi satisfies the classical continuity equation \partial_\mu j^\mu = 0, ensuring the charge Q = \int d^3x \, j^0 is time-independent. This form underpins particle number conservation in interacting field theories, building on the relativistic structure where currents transform as four-vectors. In the Standard Model of particle physics, separate continuity equations govern baryon number and lepton number conservation. The baryon current j_B^\mu = \frac{1}{3} \sum_{f=u,d,s,c,b,t} \bar{q}_f \gamma^\mu q_f, summing over quark flavors q_f, obeys \partial_\mu j_B^\mu = 0 exactly, as the U(1)_B symmetry is anomaly-free and preserved across electromagnetic, strong, and weak interactions at all orders. Similarly, the total lepton current j_L^\mu = \sum_{\ell=e,\mu,\tau} (\bar{\ell} \gamma^\mu \ell + \bar{\nu}_\ell \gamma^\mu P_L \nu_\ell) satisfies \partial_\mu j_L^\mu = 0 perturbatively, reflecting an accidental global U(1)_L symmetry, though non-perturbative weak effects like sphalerons can violate B + L while conserving B - L. Weak interactions introduce subtle violations of individual flavor lepton numbers, as evidenced by neutrino oscillations discovered in 1998 by the Super-Kamiokande experiment, which observed muon neutrino disappearance in atmospheric data, indicating nonzero neutrino masses and mixing angles on the order of \sin^2 2\theta \approx 1. This mixing implies \Delta L = \pm 1 transitions between flavors, though total lepton number remains conserved in the minimal seesaw extension of the Standard Model; full \Delta L = 2 processes, like neutrinoless double beta decay, remain unobserved but constrained to lifetimes exceeding $10^{26} years. Lattice QCD simulations provide a non-perturbative framework to verify continuity equations by computing current divergences numerically. For vector currents, improved lattice actions ensure \partial_\mu j^\mu \approx 0 up to discretization errors of order a^2 (lattice spacing a \approx 0.05 fm), allowing precise matrix elements for hadronic processes. Axial currents exhibit nonzero divergences due to the quantum chiral anomaly, reproduced universally in the continuum limit across formulations like Wilson or overlap fermions, matching continuum QFT predictions such as \partial_\mu j_5^\mu = \frac{g^2}{16\pi^2} \mathrm{tr}(F_{\mu\nu} \tilde{F}^{\mu\nu}).

References

  1. [1]
    [PDF] The Continuity Equation
    The continuity equation expresses conservation of mass, stating that mass in a volume changes locally through flux convergence or divergence.
  2. [2]
    7.7: Conservation of Mass: The Continuity Equation
    Nov 11, 2024 · Let's derive the equation for the conservation of mass in a fluid. We begin by writing down the flow of mass into and out of a small box.
  3. [3]
    Continuity Equation – Introduction to Aerospace Flight Vehicles
    Applying the principle of the conservation of mass to fluids results in a governing “star” equation called the continuity equation. This equation applies to all ...
  4. [4]
    Navier-Stokes Equations
    The Navier-Stokes equations consists of a time-dependent continuity equation for conservation of mass, three time-dependent conservation of momentum ...
  5. [5]
    [PDF] Current, continuity equation, resistance, Ohm's law. - MIT
    Feb 24, 2005 · This result is known as the continuity equation: it guarantees conservation of electric charge in the presence of currents. It also tells us ...
  6. [6]
    4.1: Continuity Equation and the Kirchhoff Laws - Physics LibreTexts
    Mar 5, 2022 · The electric field causes the motion of charge carriers in the wire – from the conductor with a higher electrostatic potential toward that with lower potential.
  7. [7]
    Equation of Continuity - an overview | ScienceDirect Topics
    The elementary physics of continuity equations is the conservation of mass principle, which is defined as the conservation law the rate at which mass changes ...
  8. [8]
    Calculus III - Surface Integrals of Vector Fields
    Nov 16, 2022 · In order to work with surface integrals of vector fields we will need to be able to write down a formula for the unit normal vector corresponding to the ...
  9. [9]
    Flux (Surface Integrals of Vector Fields) - Oregon State University
    The flux across S is the volume of fluid crossing S per unit time. The figure below shows a surface S and the vector field F at various points on the surface.
  10. [10]
    Mikhail Vasilevich Ostrogradski (1801 - 1862) - Biography - MacTutor
    Gauss, nor knowing about Ostrogradski's paper, proved special cases of the divergence theorem in 1833 and 1839 and the theorem is now often named after Gauss.
  11. [11]
    [PDF] Integral approach to the continuity equation
    However, the various forms of the continuity equation developed along the way are all useful in the variety of applications explored in this text.
  12. [12]
    [PDF] Governing Equations of Fluid Dynamics
    The governing equations of fluid dynamics are the continuity, momentum, and energy equations, based on mass conservation, F=ma, and energy conservation.
  13. [13]
    [PDF] AA210A Fundamentals of Compressible Flow
    Jan 19, 2022 · The integral form of the continuity equation on a Mixed Eulerian-. Lagrangian control volume is derived as follows. Let F in Liebniz rule be the ...
  14. [14]
    [PDF] 11–Applications of the Divergence Theorem - UC Davis Math
    Conclude: The continuity equation implies that the total rate of change of mass inside any volume is equal to minus the flux of mass out through the boundary, a ...
  15. [15]
    Divergence Theorem - Department of Mathematics at UTSA
    Nov 10, 2021 · The divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a theorem which relates the flux of a vector field through ...
  16. [16]
    [PDF] Derivation of the Continuity Equation (Section 9-2, Çengel and ...
    We summarize the second derivation in the text – the one that uses a differential control volume. First, we approximate the mass flow rate into or out of each ...
  17. [17]
    [PDF] CHAPTER 5. THE CONTINUITY EQUATION - Projects at Harvard
    In the present chaper we derive the continuity equation in its Eulerian form (fixed coordinate system) and in its Lagrangian form (moving coordinate system). ...
  18. [18]
    [PDF] ( ) − 1 ( ) 1 ( )+
    Continuity Equation (aka COM for a differential CV ). The continuity equation, which is simply conservation of mass for a differential fluid element or control.
  19. [19]
    [PDF] Lecture 4: Noether's Theorem Sept. 16, 2013 - Rutgers Physics
    Sep 16, 2013 · This completes our alternate derivation that S′ − S = RR d4xδL, and Eq. (12). Note that δL is a divergence plus a piece which vanishes if ...
  20. [20]
    [PDF] Differential form of continuity
    The differential form of continuity is ∂ρ/∂t + ∂(ρu)/∂x + ∂(ρv)/∂y + ∂(ρw)/∂z = 0, or ∂ρ/∂t + ∇.(ρu) = 0.
  21. [21]
    Euler Equations
    Continuity: partial u/ partial x + partial v/ partial y = 0. We can then factor the momentum equations and use the continuity equation to simplify them: X ...
  22. [22]
    Normal Shock Wave Equations
    On this slide we have listed the equations which describe the change in flow variables for flow across a normal shock. The equations presented here were derived ...Missing: continuity | Show results with:continuity
  23. [23]
    [PDF] Chapter 1 Maxwell's Equations
    Eq. (7) is a statement of current conservation, that is, currents cannot be generated or destroyed, the net flux through a closed surface is zero.
  24. [24]
    [PDF] Section 1: Maxwell Equations
    The continuity equation says that the total charge in any infinitesimal volume is constant unless there is a net flow of a pre-existing charge into or out of ...
  25. [25]
    [PDF] Electromagnetism - Lecture 8 Maxwell's Equations
    This conservation law is written as a continuity equation: I = IAJ.dS ... The displacement current is the time derivative of the electric field.
  26. [26]
    [PDF] The Continuity Equation and Poynting's Theorem
    The continuity equation is the mathematical statement of local charge conservation. Poynting's theorem is the work-energy theorem of electrodynamics.
  27. [27]
    Ampere-Maxwell — Electromagnetic Geophysics - EM GeoSci
    In 1861 [Max61], James Clerk Maxwell extended Ampere's law by introducing the displacement current into the electric current term to satisfy the continuity ...
  28. [28]
    VIII. A dynamical theory of the electromagnetic field - Journals
    Mathematical theories of statical electricity, of magnetism, of the mechanical action between conductors carrying currents, and of the induction of currents ...
  29. [29]
    18 The Maxwell Equations - Feynman Lectures - Caltech
    In particular, the equation for the magnetic field of steady currents was known only as ∇×B=jϵ0c2. Maxwell began by considering these known laws and expressing ...Missing: continuity | Show results with:continuity<|control11|><|separator|>
  30. [30]
    [PDF] Chapter 5: The heat equation - UC Davis Math
    The heat equation, ut = ∆u, models heat flow. It describes how a function evolves towards its mean, with the rate of change proportional to the difference ...
  31. [31]
    5.13 The First Law of Thermodynamics – Douglas College Physics ...
    The first law of thermodynamics states that the change in internal energy of a system equals the net heat transfer into the system minus the net work done by ...
  32. [32]
    [PDF] The analytical theory of heat
    JOSEPH FOURIER. TRANSLATED, WITH NOTES,. BY. ALEXANDER FREEMAN, M.A.,. FJIlLLOW O.
  33. [33]
    [PDF] The Heat Equation - Rose-Hulman
    ... heat energy flows past the point x at time t; the dimension of q is simply energy per time. If heat energy is conserved in the bar then the continuity equation.
  34. [34]
    2.5 Control volume form of the conservation laws - MIT
    The rate of change of mass inside the volume is given by the difference between the mass flow rate in and the mass flow rate out.
  35. [35]
    [PDF] HEAT CONDUCTION EQUATION
    Heat conduction is proportional to the temperature gradient, and is a vector quantity with both direction and magnitude. It can be steady or transient.
  36. [36]
    Conductive Heat Transfer - The Engineering ToolBox
    Conduction as heat transfer takes place if there is a temperature gradient in a solid or stationary fluid medium.
  37. [37]
    [PDF] Chapter 5 - Energy transport in stellar interiors - Astrophysics
    CV VT, we can write this as an equation for heat conduction,. F = −K VT with. K = 1. 3. ¯υℓCV,. (5.11) where K is the conductivity. This description is valid ...
  38. [38]
    [PDF] The continuity equation and its applications in phase space ... - arXiv
    Jul 22, 2023 · The continuity equation applied to the spatial density of probability expresses the conser- vation of probability. It plays a crucial role ...
  39. [39]
    [PDF] Simulation of the fokker-planck equation by random walks of test ...
    The structure of an analytical expression obtained from a random-walk process was compared with that of the Fokker-Planck equation.
  40. [40]
    Determining optical flow - ScienceDirect.com
    A method for finding the optical flow pattern is presented which assumes that the apparent velocity of the brightness pattern varies smoothly almost everywhere ...
  41. [41]
    Optical Flow - OpenCV Documentation
    Goal. In this chapter,. We will understand the concepts of optical flow and its estimation using Lucas-Kanade method. We will use functions ...Missing: original | Show results with:original<|separator|>
  42. [42]
    Optical Flow: Revolutionizing Motion Detection - Viso Suite
    Optical flow quantifies the motion of objects between consecutive frames captured by a camera. These algorithms attempt to capture the apparent motion of ...Techniques and Algorithms for... · Optical Flow in Action – Use...
  43. [43]
    On kinematic waves II. A theory of traffic flow on long crowded roads
    This paper uses the method of kinematic waves, developed in part I, but may be read independently. A functional relationship between flow and concentration ...
  44. [44]
    Shock Waves on the Highway | Semantic Scholar
    A simple theory of traffic flow is developed by replacing individual vehicles with a continuous “fluid” density and applying an empirical relation between ...
  45. [45]
    [PDF] continuum flow models - Federal Highway Administration
    Solution of the conservation equation as it applies to traffic flow was first proposed by Lighthill and Whitham (1955) and by. Richards (1956). Recently ...
  46. [46]
    Simulating vehicular traffic in a network using dynamic routing
    The continuity equation, the equation of vehicle number conservation, reads ... A general traffic network consists of road sections (streets) and junctions.
  47. [47]
    [PDF] A Mathematical Introduction to Traffic Flow Theory - UCLA
    Sep 11, 2015 · Note: The continuity equation merely states that no vehicles are lost or created. There is no modeling in it. In particular, it is not a closed ...
  48. [48]
    [PDF] The Lighthill-Whitham-Richards (LWR) Model - Traffic Flow Dynamics
    as traffic stream relation into the continuity equation gives (for homogeneous roads) the class of first-order models known as LWR models: ∂ρ. ∂t. +. ∂. ∂x.
  49. [49]
    Analysis and comparison of traffic flow models: a new hybrid traffic ...
    Nov 24, 2021 · This method is based on an exact solution of the continuity equation for one time step, assuming stepwise initial conditions given by the actual ...<|separator|>
  50. [50]
    Continuum modeling of supply chain networks using discontinuous ...
    Using a connectivity matrix, we establish a continuum modeling approach with partial differential equations of conservation laws for simulating materials flow ...Missing: continuity | Show results with:continuity
  51. [51]
    [PDF] Continuum-Discrete Models for Supply Chains and Networks
    Apr 26, 2011 · Introduction. A supply network consists of suppliers, manufacturers, warehouses, and stores, that perform.
  52. [52]
    [PDF] Quantum Physics I, Lecture Note 6 - MIT OpenCourseWare
    Feb 23, 2016 · We will try to understand how this equation is compatible with the time evolution prescribed by the Schrödinger equation.
  53. [53]
    [PDF] 1926-Schrodinger.pdf
    The paper gives an account of the author's work on a new form of quantum theory. §1. The Hamiltonian analogy between mechanics and optics. §2. The analogy is to ...
  54. [54]
    [PDF] Quantum mechanics of collision processes
    Introduction. Collision processes not only yield the most convincing experimental proof of the basic assumptions of quantum theory, but also seem suitable ...
  55. [55]
    [PDF] Quantum Tunneling
    The continuity equation which relates the time change of the charge density to the divergence of the current density, provides the departure point for the.
  56. [56]
    Continuity Equation in Semiconductor - EEEGUIDE.COM
    The Continuity Equation in Semiconductor states a condition of dynamic equilibrium for the concentration of mobile carriers in any elementary.
  57. [57]
    2.9 Continuity equation
    The continuity equation describes a basic concept, namely that a change in carrier density over time is due to the difference between the incoming and outgoing ...
  58. [58]
    Basic Equations | PVEducation
    In semiconductors we divide the charge up into four components: hole density, p, electron density, n, acceptor atom density, NA and donor atom density, ND.
  59. [59]
    Boltzmann approach to high-order transport: The nonlinear and ...
    Jun 20, 2017 · We provide a rigorous derivation of transport equations by solving the Boltzmann equation in the relaxation-time approximation and show that the ...
  60. [60]
    Transport theory of semiconductor energy conversion - AIP Publishing
    Dec 1, 1982 · Starting from the quasi‐classical particle Boltzmann equations, a continuity equation for total energy is derived containing terms which can be ...
  61. [61]
    Solving for Quasi Neutral Regions - PVEducation
    Equations (3a) and (3b) are general under steady state conditions as long as the depletion approximation holds (which assumes low injection) and as long as ...
  62. [62]
    [PDF] ECE 340 Lecture 24 : Quantitative Current Flow in a P-N Junction
    Oct 17, 2011 · Quantitative PN Junction Current. There are two ways to calculate current… Slopes of minority carrier concentrations. Steady state charge ...
  63. [63]
    [PDF] SOLAR CELL OPERATIONAL PRINCIPLES
    The continuity equations (Eq. 3.54) for electrons and holes in ... (4.38) describe the steady-state situation for minority carriers when the junction is.
  64. [64]
    [PDF] PN and Metal–Semiconductor Junctions
    We conclude that n = p = 0 at the junction edge under reverse bias. FIGURE 4–16 In steady state, the number of holes flowing into the box per second is equal to.
  65. [65]
    [PDF] The Mathematical Study and Approximation of Semiconductor Models
    The writer first became aware of the semiconductor drift-diffusion model in 1980, during a visit to Yale University, through the courtesy of Martin Schultz.
  66. [66]
    [PDF] Drift-diffusion and computational electronics - SciSpace
    Oct 8, 2015 · Abstract—The field of computational electronics began in a serious way when the so-called semiconductor equations were.
  67. [67]
    [PDF] Special Relativity
    Oct 29, 2010 · We'll call it the current density 4-vector. J. µ. = (cρ,Jx,Jy,Jz) ... so the continuity equation states that the current density. 4-vector ...
  68. [68]
    Electrodynamics in Relativistic Notation - Feynman Lectures - Caltech
    Now in the case of special relativity, time and space are inextricably mixed, and we must do the analogous things for four dimensions. We want our equations to ...
  69. [69]
    The Thermodynamics of Irreversible Processes. III. Relativistic ...
    The Thermodynamics of Irreversible Processes. III. Relativistic Theory of the Simple Fluid. Carl Eckart.Missing: hydrodynamics | Show results with:hydrodynamics
  70. [70]
    Relativistic fluid dynamics: physics for many different scales
    Jun 24, 2021 · recovering the usual continuity equation by introducing the mass density \rho = mn, with m the mass per particle. In order to work out the ...
  71. [71]
    [PDF] Physics 225A: General Relativity
    Dec 18, 2024 · So this equation is compatible with covariant conservation of stress energy tensor, and (we'll see later that it) reduces properly to the ...
  72. [72]
    [PDF] Revisiting the Bianchi identity 12.2 Making a theory of gravity - MIT
    Energy density, however, we now recognize as not a scalar, but rather as one component of the stress energy tensor: ρ = T00. To make this tensorial, one ...
  73. [73]
    [PDF] General Relativity Fall 2018 Lecture 9: Einstein's field equation
    Oct 4, 2018 · From the contracted Bianchi identity ∇µGµν, this equation is consistent with the conservation of stress-energy tensor. This is Einstein's ...
  74. [74]
    [PDF] Lecture XXX: Dynamics of cosmic expansion - Caltech (Tapir)
    Apr 11, 2012 · Continuity provides a strong constraint on the possible behavior of matter in cosmology. We derive the continuity equation in the FRW metric in ...
  75. [75]
    [2103.17160] Noether's Theorems and Energy in General Relativity
    Mar 31, 2021 · This paper has three main aims: first, to give a pedagogical introduction to Noether's two theorems and their implications for energy conservation in general ...
  76. [76]
    [2203.08771] Searches for Baryon Number Violation in Neutrino ...
    Mar 16, 2022 · Baryon number conservation is not guaranteed by any fundamental symmetry within the Standard Model, and therefore has been a subject of ...
  77. [77]
    [hep-ex/9807003] Evidence for oscillation of atmospheric neutrinos
    Jul 3, 1998 · We present an analysis of atmospheric neutrino data from a 33.0 kiloton-year (535-day) exposure of the Super-Kamiokande detector.
  78. [78]
    [1903.06557] Probing nonstandard lepton number violating ... - arXiv
    Mar 15, 2019 · Abstract:We discuss lepton number violating processes in the context of long-baseline neutrino oscillations. We summarise and compare ...
  79. [79]
    [hep-lat/0004005] Universality of the Axial Anomaly in Lattice QCD
    Apr 5, 2000 · We prove that lattice QCD generates the axial anomaly in the continuum limit under very general conditions on the lattice action, which includes ...