Fact-checked by Grok 2 weeks ago

Divergence theorem

The Divergence Theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a fundamental result in vector calculus that establishes a relationship between a volume integral of the divergence of a vector field over a region and the flux of that field through the boundary surface of the region. In mathematical terms, for a sufficiently smooth vector field \mathbf{F} defined on a bounded region V in \mathbb{R}^3 with piecewise smooth boundary surface S oriented outward, the theorem states: \iiint_V \nabla \cdot \mathbf{F} \, dV = \iint_S \mathbf{F} \cdot d\mathbf{S}, where \nabla \cdot \mathbf{F} is the divergence of \mathbf{F} and d\mathbf{S} is the outward-pointing area element on S. This theorem generalizes the one-dimensional and the two-dimensional to three dimensions, forming part of the broader framework of the fundamental theorems of that connect differential forms to integrals over manifolds. Historically, the result was first noted without proof by in 1762 in the context of , but it was rigorously stated and proved in its modern form by in 1813; Mikhail Ostrogradsky independently derived and published a proof in 1828, particularly emphasizing its application to . The Divergence Theorem has profound applications across physics and , particularly in —where it underpins relating to enclosed charge—fluid mechanics for analyzing incompressible flows and continuity equations, and for deriving laws. It also facilitates the solution of partial differential equations, such as in , by transforming volume integrals into more tractable surface integrals or vice versa. In higher , the theorem extends to more general settings via differential forms on manifolds, playing a key role in and .

Intuitive Explanations

Flux Interpretation

The flux of a through an oriented surface quantifies the net of the field across that surface in the direction specified by the , often visualized as the amount of "stuff" passing through per unit area. For a closed surface enclosing a , the is typically taken with the outward-pointing normal vector, so the measures the total outflow from the interior region. This interpretation arises from considering the as representing a field of a , where the of the field with the normal indicates the component of perpendicular to the surface. The divergence of the vector field at a point captures the local behavior of expansion or contraction within the field, serving as a measure of how much the field is spreading out or converging at that location per unit volume. Positive divergence indicates a source-like region where field lines emanate outward, while negative divergence points to a sink where lines converge inward. Integrating the divergence over the entire volume enclosed by the surface thus accumulates these local source or sink contributions, providing a total measure of the net creation or destruction of the field inside the region. Visually, regions of positive divergence appear as areas where the vector field arrows radiate away from a point, leading to a net outflow through any surrounding closed surface, much like water emerging from a . Conversely, negative divergence regions show arrows pointing toward a central point, resulting in net inflow, akin to a pulling in fluid. This creates an intuitive link between the internal dynamics of the field and the boundary behavior observed as . To build before considering three dimensions, a two-dimensional considers the across a simple closed curve in the , which equals the total integrated over the enclosed area, highlighting how local expansions or compressions along the boundary relate to sources or sinks inside. This planar version, often encountered in the form of , mirrors the three-dimensional case by showing that net flow through the boundary depends solely on the internal , without regard to the specific shape of the enclosure.

Liquid Flow Analogy

One intuitive way to understand the divergence theorem is by envisioning a as the field of particles flowing within a closed , such as a three-dimensional region representing a volume of fluid. In this analogy, the vectors indicate the direction and speed at which the fluid moves at each point inside the container. For a steady, where the liquid does not accumulate or deplete within the , the total amount of leaving through the must exactly balance the net addition or removal of from internal sources or sinks, such as faucets adding liquid or drains removing it. This conservation principle ensures that any excess introduced by sources inside will result in a corresponding net outflow across the entire surface, while sinks would cause a net inflow. The divergence of the vector field captures the local "source strength" at each point, quantifying how much the fluid is expanding or contracting in that vicinity; a positive divergence indicates a local source where fluid is being created or expanding outward, whereas zero divergence signifies no net creation or destruction, allowing the fluid to simply pass through without accumulation. For instance, in a scenario of uniform flow—where the velocity is constant everywhere, like water streaming steadily through parallel pipes—the net flux across any closed boundary is zero, as the inflow matches the outflow precisely. Conversely, if internal sources are present, such as points where fluid is injected (creating positive divergence), the overall effect is a positive net flux out of the boundary, illustrating how localized expansions drive the global outflow.

Mathematical Formulation

Statement in Euclidean Space

The divergence theorem in \mathbb{R}^3 states that if \mathbf{F} = (P, Q, R) is a whose components P, Q, and R have continuous first-order partial derivatives on an containing a bounded domain V with piecewise smooth boundary \partial V, then the of \mathbf{F} through the oriented closed surface \partial V equals the volume of the of \mathbf{F} over V: \iint_{\partial V} \mathbf{F} \cdot \mathbf{n} \, dS = \iiint_V \left( \frac{\partial P}{\partial x} + \frac{\partial Q}{\partial y} + \frac{\partial R}{\partial z} \right) dV, where \mathbf{n} denotes the outward-pointing unit normal vector to \partial V. The domain V must be bounded and compactly contained in the region where \mathbf{F} is C^1, ensuring the integrals are well-defined, while the piecewise smooth boundary allows for the surface to consist of finitely many smooth pieces. This result extends naturally to \mathbb{R}^n for n \geq 1. Let U \subset \mathbb{R}^n be a bounded open set with piecewise smooth boundary \partial U, and let \mathbf{F} = (F_1, \dots, F_n) be a vector field that is continuously differentiable on U \cup \partial U. Then, \int_U \nabla \cdot \mathbf{F} \, dV = \int_{\partial U} \mathbf{F} \cdot \mathbf{n} \, dS, where the divergence is \nabla \cdot \mathbf{F} = \sum_{i=1}^n \frac{\partial F_i}{\partial x_i}, dV is the volume element in \mathbb{R}^n, dS is the surface element on \partial U, and \mathbf{n} is again the outward unit normal. The regularity assumptions on \mathbf{F} and \partial U mirror those in the three-dimensional case, guaranteeing the existence of the integrals and the validity of the orientation convention.

Notation and Assumptions

The divergence theorem in Euclidean space \mathbb{R}^n employs standard notation for integrals and geometric elements. The volume integral is denoted \int_V \nabla \cdot \mathbf{F} \, dV, where dV represents the volume element and \nabla \cdot \mathbf{F} is the divergence of the vector field \mathbf{F}. The surface integral over the boundary is \int_{\partial V} \mathbf{F} \cdot \mathbf{n} \, dS, with dS the surface area element and \mathbf{n} the outward-pointing unit normal vector to \partial V. The theorem requires the domain V to be a compact of \mathbb{R}^n, typically the of a bounded , ensuring the integrals are well-defined over finite measures. The \partial V must be piecewise C^1, consisting of finitely many smooth hypersurfaces meeting at edges of 2, to allow a consistent and into charts for . This piecewise smoothness facilitates the definition of the unit \mathbf{n} on \partial V. Topologically, V is a bounded whose \partial V is a compact orientable (n-1)-manifold without , possibly consisting of multiple connected components. This setup allows for domains with holes, where the includes all components with outward-pointing normals relative to V. In two dimensions (n=2), \partial V consists of finitely many simple closed curves, each guaranteed by the to bound a well-defined region, and the theorem corresponds to the divergence form of . For smoothness, the \mathbf{F}: \mathbb{R}^n \supset V \to \mathbb{R}^n must be continuously differentiable (C^1) on a neighborhood of the \overline{V}, ensuring the partial derivatives exist and the and surface integrals are continuous and finite. Extensions to less regular fields, such as those in L^1(\overline{V})^n with in the sense of distributions, hold via density arguments: smooth (C^\infty_c) approximations converge in L^1, preserving the equality under .

Derivations and Proofs

Informal Derivation

To derive the divergence theorem heuristically, consider a bounded volume V in \mathbb{R}^3 divided into a large number of tiny , each of side length \Delta x = \Delta y = \Delta z = h, where h is small. For a single such cube centered at a point \mathbf{x}, the net of a \mathbf{F} out of the cube's surface is approximately equal to the of \mathbf{F} at \mathbf{x} multiplied by the cube's volume h^3. This follows from the intuitive notion that the divergence measures the local "source" or "sink" strength of the field, so the net outflow through the should with that quantity times the enclosed volume. To see this more precisely, apply a first-order expansion to the components of \mathbf{F} = (F_x, F_y, F_z) around the cube's center. For the pair of faces perpendicular to the x-axis, the flux through the right face at x + h/2 is approximately F_x(\mathbf{x} + (h/2)\mathbf{\hat{x}}) \cdot h^2 \approx [F_x(\mathbf{x}) + \frac{\partial F_x}{\partial x} h/2] h^2, while through the left face at x - h/2 it is approximately -[F_x(\mathbf{x}) - \frac{\partial F_x}{\partial x} h/2] h^2 (negative due to the inward ). The net contribution from these faces is thus \frac{\partial F_x}{\partial x} h^3. Analogous expansions for the y- and z-faces yield \frac{\partial F_y}{\partial y} h^3 and \frac{\partial F_z}{\partial z} h^3, respectively, so the total net flux is (\nabla \cdot \mathbf{F})(\mathbf{x}) \, h^3. Heuristically, this suggests \nabla \cdot \mathbf{F} \approx \frac{\text{net flux out}}{h^3} for small h. Summing the net fluxes over all cubes in the partition, the contributions from interior faces shared between adjacent cubes cancel pairwise (outflow from one equals inflow to the other). The remaining uncancelled fluxes are those on the outer boundary of V, which collectively approximate the integral over \partial V. Taking the limit as h \to 0, the sum of volume contributions \sum (\nabla \cdot \mathbf{F}) h^3 becomes the triple \iiint_V \nabla \cdot \mathbf{F} \, dV, motivating the formal of the divergence theorem: \iiint_V \nabla \cdot \mathbf{F} \, dV = \iint_{\partial V} \mathbf{F} \cdot d\mathbf{S}.

Proof for Bounded Domains in R^n

The proof of the divergence theorem for a bounded domain \Omega \subset \mathbb{R}^n with piecewise smooth boundary \partial \Omega and a C^1 \mathbf{F} = (F_1, \dots, F_n) proceeds componentwise, showing that \int_{\partial \Omega} F_i \mathbf{n} \cdot \mathbf{e}_i \, dS = \int_\Omega \frac{\partial F_i}{\partial x_i} \, dV for each i = 1, \dots, n, where \mathbf{n} is the outward unit normal and \mathbf{e}_i the vector; summing over i yields \int_{\partial \Omega} \mathbf{F} \cdot \mathbf{n} \, dS = \int_\Omega \nabla \cdot \mathbf{F} \, dV. To establish the result in three dimensions first, consider the x-component with F_1 = P. Project \Omega onto the yz-plane to obtain a domain D \subset \mathbb{R}^2, and for each (y,z) \in D, let the slice of \Omega parallel to the x-axis extend from a(y,z) to b(y,z). By Fubini's theorem, the volume integral decomposes as \int_\Omega \frac{\partial P}{\partial x} \, dV = \int_D \left( \int_{a(y,z)}^{b(y,z)} \frac{\partial P}{\partial x}(x,y,z) \, dx \right) dy \, dz. Applying the fundamental theorem of calculus to the inner integral gives \int_{a(y,z)}^{b(y,z)} \frac{\partial P}{\partial x}(x,y,z) \, dx = P(b(y,z), y, z) - P(a(y,z), y, z). Thus, \int_\Omega \frac{\partial P}{\partial x} \, dV = \int_D P(b(y,z), y, z) \, dy \, dz - \int_D P(a(y,z), y, z) \, dy \, dz. The first integral represents the flux of P through the portion of \partial \Omega where the outward normal has positive x-component (corresponding to the "right" faces of the slices, with \mathbf{n} \cdot \mathbf{e}_1 = 1), while the second is the flux through the portions where \mathbf{n} \cdot \mathbf{e}_1 = -1 (left faces). For a general boundary, the surface integral \int_{\partial \Omega} P n_x \, dS matches this difference, as the projected area elements dy \, dz relate to dS via the cosine of the angle between \mathbf{n} and \mathbf{e}_1, which is n_x. The proofs for the y- and z-components follow analogously by projecting onto the other coordinate planes and using Fubini's theorem for the respective iterated integrals. Summing the three components completes the proof for \mathbb{R}^3. To generalize to \mathbb{R}^n, apply Fubini's theorem iteratively over the coordinates. For the i-th component, project \Omega onto the hyperplane orthogonal to \mathbf{e}_i and integrate slices along the x_i-direction, reducing the volume \int_\Omega \frac{\partial F_i}{\partial x_i} \, dV to a difference of (n-1)-dimensional integrals over the "end" faces of the slices, which correspond to the flux through the portions of \partial \Omega where the outward normal aligns positively or negatively with \mathbf{e}_i. The factor n_i = \mathbf{n} \cdot \mathbf{e}_i accounts for the orientation and projection in the surface measure, ensuring \int_{\partial \Omega} F_i n_i \, dS = \int_\Omega \frac{\partial F_i}{\partial x_i} \, dV. Summing over all components yields the full theorem, as the process is symmetric in the coordinates. For general bounded domains with C^1 boundary, approximate \Omega by a sequence of polyhedral domains \Omega_k (e.g., convex polyhedra inscribed or circumscribed around \Omega) such that \Omega_k \to \Omega in the Hausdorff metric and the boundaries converge appropriately. The divergence theorem holds exactly for polyhedra, as the boundary consists of flat faces where the flux is computed directly via the normal on each face, and internal faces cancel when decomposing into simplices. By continuity of the integrals under this approximation (using uniform convergence of \mathbf{F} and the boundaries), the result extends to \Omega. For less regular \mathbf{F}, approximate by smooth vector fields via mollification with compactly supported kernels, preserving the divergence in the limit due to the density of smooth functions in C^1(\overline{\Omega}).

Proof on Riemannian Manifolds

The divergence theorem on a smooth, compact, oriented Riemannian manifold (M, g) of dimension n with boundary \partial M states that for any smooth vector field X on M, \int_M (\div_g X) \, \vol_g = \int_{\partial M} g(X, N) \, \vol_{\partial M}, where \div_g X denotes the divergence of X with respect to the metric g, \vol_g is the Riemannian volume form on M, N is the outward-pointing unit normal vector field to \partial M, and \vol_{\partial M} is the induced Riemannian volume form on the boundary with its induced orientation from M. The divergence \div_g X is defined globally as the unique on M satisfying \Lie_X \vol_g = (\div_g X) \vol_g, where \Lie_X is the along X. Equivalently, in local coordinates (x^1, \dots, x^n) on an U \subset M, if X = X^i \frac{\partial}{\partial x^i}, then \div_g X = \frac{1}{\sqrt{\det g}} \frac{\partial}{\partial x^i} \left( \sqrt{\det g} \, X^i \right), with the Einstein summation convention over i = 1, \dots, n, and g = (g_{jk}) the components with \det g > 0. This coordinate expression follows from computing the of the local expression \vol_g = \sqrt{\det g} \, dx^1 \wedge \cdots \wedge dx^n and is independent of the choice of coordinates due to the tensorial nature of the construction. To prove the theorem, consider the interior product i_X \vol_g, which is an (n-1)-form on M. A key identity is d(i_X \vol_g) = (\div_g X) \vol_g, where d is the . This holds globally but is verified locally: in coordinates, i_X \vol_g = \sum_{i=1}^n (-1)^{i-1} \sqrt{\det g} \, X^i \, dx^1 \wedge \cdots \wedge \widehat{dx^i} \wedge \cdots \wedge dx^n, and applying d yields the divergence term after expansion and using the on \sqrt{\det g} \, X^i. Applying the general Stokes' theorem on oriented manifolds with boundary, which states that \int_M d\alpha = \int_{\partial M} \alpha for any smooth (n-1)-form \alpha on M, take \alpha = i_X \vol_g. The left side becomes \int_M (\div_g X) \vol_g. For the right side, restrict to \partial M: since the induced volume form satisfies \vol_g (N, V_1, \dots, V_{n-1}) = \vol_{\partial M} (V_1, \dots, V_{n-1}) by the definition of the induced orientation and volume form (with N outward), evaluating i_X \vol_g on tangent vectors V_1, \dots, V_{n-1} to \partial M gives i_X \vol_g (V_1, \dots, V_{n-1}) = \vol_g (X, V_1, \dots, V_{n-1}) = g(X, N) \vol_{\partial M} (V_1, \dots, V_{n-1}), as the tangential components of X contribute zero to the pairing with N. Thus, i_X \vol_g|_{\partial M} = g(X, N) \vol_{\partial M}, completing the proof. This vector-field version relates directly to the coordinate-based proofs in , where the local chart computations mirror the flat case after accounting for the metric determinant, but the form approach leverages the general proved via simplicial approximations or partitions of unity on the manifold.

Corollaries and Identities

Vector Calculus Connections

The divergence theorem forms one of the four fundamental theorems of , alongside the , , and , collectively establishing a framework that relates integrals over domains to their boundaries in a manner analogous to the . These theorems unify the treatment of scalar and vector fields by connecting line integrals, surface integrals, and volume integrals through differential operators like , , and . Key vector identities underpin these connections, such as the identity \nabla \cdot (\nabla \times \mathbf{F}) = 0 for any sufficiently smooth \mathbf{F}, which implies that the flux of a field through any closed surface vanishes by the divergence theorem, characterizing solenoidal (divergence-free) fields. Similarly, \nabla \times (\nabla \phi) = 0 for any \phi, ensuring that the of a field is path-independent and the associated flux relates conservatively to values. In the Helmholtz decomposition, any vector field in \mathbb{R}^3 (with appropriate decay conditions) can be uniquely expressed as the sum of an irrotational part (gradient of a scalar), a solenoidal part (curl of a vector potential), and a harmonic part (both curl-free and divergence-free), with the divergence theorem essential for proving the orthogonality of these components in the L^2 sense over bounded domains. This decomposition highlights the theorem's role in structuring vector field analysis across Euclidean space.

Green's Identities

Green's identities are integral relations derived from the divergence theorem that connect volume integrals of scalar functions and their Laplacians to boundary integrals involving normal derivatives. These identities are particularly useful in the study of elliptic partial differential equations, such as Laplace's equation, by facilitating integration by parts in higher dimensions. Green's first identity arises directly from applying the divergence theorem to the vector field \mathbf{F} = \phi \nabla \psi, where \phi and \psi are sufficiently smooth scalar functions defined on a bounded domain V \subset \mathbb{R}^n with piecewise smooth boundary \partial V. The divergence of \mathbf{F} is given by \nabla \cdot (\phi \nabla \psi) = \phi \Delta \psi + \nabla \phi \cdot \nabla \psi, where \Delta denotes the Laplacian. Integrating over V and applying the divergence theorem yields \int_V (\phi \Delta \psi + \nabla \phi \cdot \nabla \psi) \, dV = \int_{\partial V} \phi \frac{\partial \psi}{\partial n} \, dS, where \frac{\partial \psi}{\partial n} = \nabla \psi \cdot \mathbf{n} is the outward normal derivative on \partial V. This identity generalizes the one-dimensional integration by parts formula and holds under the assumptions that \phi, \psi \in C^2(V) and the domain satisfies the necessary regularity for the divergence theorem. To obtain Green's second identity, apply the first identity with the roles of \phi and \psi interchanged: \int_V (\psi \Delta \phi + \nabla \psi \cdot \nabla \phi) \, dV = \int_{\partial V} \psi \frac{\partial \phi}{\partial n} \, dS. Subtracting this from the original first identity gives \int_V (\phi \Delta \psi - \psi \Delta \phi) \, dV = \int_{\partial V} \left( \phi \frac{\partial \psi}{\partial n} - \psi \frac{\partial \phi}{\partial n} \right) dS. This form is symmetric and plays a key role in uniqueness proofs for boundary value problems. The derivation relies solely on the first identity and the linearity of the integrals, preserving the same smoothness and domain assumptions. A notable application of Green's second identity occurs when one function is , satisfying \Delta u = 0. Setting \psi to the fundamental solution of leads to the mean value property: for a u in a B_r(x_0), u(x_0) equals the average of u over the sphere \partial B_r(x_0). This property characterizes among continuous solutions to .

Special Cases like Zero Divergence

A fundamental of the divergence theorem arises when the vector field \mathbf{F} is divergence-free, meaning \nabla \cdot \mathbf{F} = 0 throughout a V. , the theorem implies that the through the closed surface \partial V vanishes: \int_{\partial V} \mathbf{F} \cdot \mathbf{n} \, dS = 0. This indicates no net flux emanating from or entering the , as there are no sources or sinks within V. Such fields are termed solenoidal or source-free, reflecting their incompressible nature. This property has significant implications across mathematical modeling. In , divergence-free fields describe incompressible flows, where the net volume through any closed surface is zero, conserving without expansion or contraction. Similarly, in the mathematical of , \nabla \cdot \mathbf{B} = 0 ensures no net magnetic through closed surfaces, consistent with the absence of magnetic monopoles. In , Kirchhoff's current law—stating that the algebraic sum of currents at any node is zero—analogizes this zero-divergence condition in , derived via the divergence theorem applied to fields. For more complex domains, such as those with holes or voids (multiply connected regions), the divergence theorem extends by considering the outer boundary and inner boundary components separately. If \nabla \cdot \mathbf{F} = 0 in the domain between the outer surface S_0 and inner surfaces S_i, the flux through S_0 equals the sum of the fluxes through the S_i (with inward orientation for the holes). Thus, while the overall net flux remains zero, the fluxes through inner boundaries can be nonzero and are linked to the domain's topology, such as the number of holes, influencing global field behavior. A special class of divergence-free fields are harmonic vector fields, defined by both \nabla \cdot \mathbf{F} = 0 and \nabla \times \mathbf{F} = \mathbf{0}. In simply connected domains, such fields admit a representation as the of a harmonic \phi, where \mathbf{F} = \nabla \phi and \Delta \phi = 0, satisfying . This underscores their in and solutions to boundary value problems.

Applications in Physics and Mathematics

Continuity and Conservation Laws

The continuity equation expresses the local conservation of a quantity, such as mass or charge, in the absence of sources or sinks, in differential form as \partial \rho / \partial t + \nabla \cdot \mathbf{J} = 0, where \rho is the density of the conserved quantity and \mathbf{J} is the flux density vector. This equation states that the rate of change of density at a point equals the negative divergence of the flux, indicating that any local increase in density must be balanced by a net influx from surrounding regions. Applying the divergence theorem to the over a fixed V with \partial V yields the integral form: \frac{d}{dt} \int_V \rho \, dV + \int_{\partial V} \mathbf{J} \cdot \mathbf{n} \, dS = 0, where \mathbf{n} is the outward unit normal to the . This equation, derived by integrating the differential form and interchanging the time derivative with the (valid under suitable smoothness assumptions), shows that the rate of change of the total quantity inside V equals the negative of the through the . Physically, it means the decrease in the enclosed quantity is exactly accounted for by the outflow across the surface, embodying global conservation. In steady-state conditions, where \partial \rho / \partial t = 0, the simplifies to \nabla \cdot \mathbf{J} = 0, implying that the flux is divergence-free and the total quantity is conserved without temporal variation. The integral form then reduces to \int_{\partial V} \mathbf{J} \cdot \mathbf{n} \, dS = 0, confirming zero net flux through any closed surface. This framework applies to mass conservation in , where \rho is mass density and \mathbf{J} = \rho \mathbf{v} with \mathbf{v} the velocity field, ensuring that mass neither accumulates nor depletes in a except through boundary flows. In , it governs , with \rho as and \mathbf{J} the , leading to steady-state conditions where \nabla \cdot \mathbf{J} = 0 prevents charge buildup.

Electromagnetic and Gravitational Fields

The divergence theorem plays a central role in by connecting the differential form of to their integral counterparts, particularly for the electric and magnetic fields. In , the differential equation \nabla \cdot \mathbf{E} = \frac{\rho}{\epsilon_0} describes how the divergence of the \mathbf{E} relates to the \rho, with \epsilon_0 as the . Applying the divergence theorem to a volume V bounded by surface \partial V yields \int_{\partial V} \mathbf{E} \cdot d\mathbf{S} = \frac{1}{\epsilon_0} \int_V \rho \, dV = \frac{Q_{\text{encl}}}{\epsilon_0}, where Q_{\text{encl}} is the total charge enclosed in V. This integral form, known as for electricity, quantifies the net through the surface as determined solely by the enclosed charge, enabling the analysis of field behavior from internal sources. For magnetostatics, Maxwell's equation \nabla \cdot \mathbf{B} = 0 indicates the absence of magnetic monopoles, implying that magnetic field lines form closed loops. The applied to this equation over the same and surface gives \int_{\partial V} \mathbf{B} \cdot d\mathbf{S} = 0, or , meaning the net through any closed surface is zero regardless of enclosed currents or magnets. This result underscores that magnetic fields originate from sources like currents or spins but do not have isolated "north" or "south" poles as sinks or sources of flux. In Newtonian gravity, the gravitational field \mathbf{g} satisfies \nabla \cdot \mathbf{g} = -4\pi G \rho_m, where \rho_m is the mass density and G is the gravitational constant. By the divergence theorem, this leads to \int_{\partial V} \mathbf{g} \cdot d\mathbf{S} = -4\pi G \int_V \rho_m \, dV = -4\pi G M_{\text{encl}}, with M_{\text{encl}} the enclosed mass. Known as Gauss's law for gravity, this integral form reveals that the net gravitational flux through a closed surface depends only on the mass within, analogous to the electric case but with an attractive sign convention, facilitating derivations of field strengths from mass distributions.

Inverse-Square Laws and Potential Theory

The vector field associated with inverse-square laws, such as \mathbf{F} = \frac{\mathbf{r}}{|\mathbf{r}|^3}, has a divergence that vanishes everywhere except at the origin, where it equals $4\pi \delta(\mathbf{r}), with \delta(\mathbf{r}) denoting the three-dimensional Dirac delta function. This distributional divergence arises because direct computation yields zero for \mathbf{r} \neq 0, but the divergence theorem reveals a nonzero flux through any closed surface enclosing the origin, necessitating the delta function to reconcile the singularity. Applying the divergence theorem to such a field over a volume V enclosing sources yields \oint_{\partial V} \mathbf{F} \cdot d\mathbf{S} = \int_V \nabla \cdot \mathbf{F} \, dV = 4\pi (up to scaling constants for physical charges or masses), interpreting the flux as proportional to the enclosed "charge." Away from sources, the is irrotational (\nabla \times \mathbf{F} = 0), allowing representation as the of a \phi via \mathbf{F} = -\nabla \phi, where \phi(\mathbf{r}) = -\int_{\infty}^{\mathbf{r}} \mathbf{F} \cdot d\mathbf{l} defines the potential through a along a path from . In potential theory, the governing equation outside sources is Laplace's equation \Delta \phi = 0, derived from the source-free divergence \nabla \cdot \mathbf{F} = 0 and \mathbf{F} = -\nabla \phi. Solutions to this equation satisfy boundary value problems, where —obtained by applying the divergence theorem to products of harmonic functions—express volume integrals in terms of surface integrals over boundaries, facilitating the representation of potentials via boundary data. For far-field approximations, multipole expansions of the potential \phi(\mathbf{r}) decompose it into , , and higher-order terms, with coefficients determined by moments of the source distribution; these expansions employ surface integrals derived from to approximate distant behavior without full volume computation.

Illustrative Examples

Sphere Flux Calculation

To illustrate the divergence theorem numerically, consider the radial \mathbf{F} = \frac{(x, y, [z](/page/Z))}{[r](/page/R)^3}, where r = \sqrt{x^2 + y^2 + [z](/page/Z)^2}, and compute its through S defined by x^2 + y^2 + [z](/page/Z)^2 = [1](/page/1), oriented outward. On the surface of , r = [1](/page/1), so \mathbf{F} = (x, y, [z](/page/Z)) = \hat{\mathbf{[r](/page/R)}}, the unit radial . Parametrize the sphere using spherical coordinates: \mathbf{r}(\theta, \phi) = (\sin\theta \cos\phi, \sin\theta \sin\phi, \cos\theta), with $0 \leq \theta \leq \pi and $0 \leq \phi \leq 2\pi. The outward-pointing surface element is d\mathbf{S} = \hat{\mathbf{r}} \sin\theta \, d\theta \, d\phi. Thus, the flux integral is \iint_S \mathbf{F} \cdot d\mathbf{S} = \int_0^{2\pi} \int_0^\pi \hat{\mathbf{r}} \cdot \hat{\mathbf{r}} \sin\theta \, d\theta \, d\phi = \int_0^{2\pi} d\phi \int_0^\pi \sin\theta \, d\theta = 2\pi [-\cos\theta]_0^\pi = 2\pi (1 - (-1)) = 4\pi. This result holds by direct symmetry, as \mathbf{F} is everywhere normal to the sphere and of unit magnitude on its surface. For the volume integral over the unit ball V enclosed by S, the divergence \nabla \cdot \mathbf{F} = 0 at all points where r > 0. However, \mathbf{F} has a singularity at the origin, and the divergence theorem implies \iiint_V \nabla \cdot \mathbf{F} \, dV = 4\pi. This equality is reconciled by interpreting the divergence in the distributional sense: \nabla \cdot \mathbf{F} = 4\pi \delta^3(\mathbf{r}), where \delta^3(\mathbf{r}) is the three-dimensional Dirac delta function concentrated at the origin. Alternatively, excluding a small ball of radius \epsilon > 0 around the origin and taking the limit as \epsilon \to 0^+ yields the same $4\pi, confirming the theorem's validity despite the singularity. The equality of the surface and volume integrals to $4\pi verifies the divergence theorem for this field, where the nonzero flux arises from the enclosed singularity at the center, analogous to a point source.

Charge Distribution in a Volume

To illustrate the divergence theorem for a volume with non-constant divergence arising from a distributed source, consider a ball of radius R filled with uniform charge density \rho, modeled mathematically by a vector field \mathbf{F} satisfying \nabla \cdot \mathbf{F} = \rho inside the ball and \nabla \cdot \mathbf{F} = 0 outside./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/06%3A_Gauss's_Law/6.04%3A_Applying_Gausss_Law) By spherical symmetry, \mathbf{F} is radial, \mathbf{F}(r) = F(r) \hat{r}, where r = |\mathbf{r}| and \mathbf{r} is the position vector from the center. To solve for \mathbf{F}, apply the in integral form to Gaussian spheres of radius r. For r < R (inside the ball), the enclosed "charge" is \rho \cdot \frac{4}{3} \pi r^3. The flux through the Gaussian surface is F(r) \cdot 4 \pi r^2 = \rho \cdot \frac{4}{3} \pi r^3, yielding F(r) = \frac{\rho r}{3}./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/06%3A_Gauss's_Law/6.04%3A_Applying_Gausss_Law) For r > R (outside), the enclosed charge is the total Q = \rho \cdot \frac{4}{3} \pi R^3, so F(r) = \frac{Q}{4 \pi r^2} = \frac{\rho R^3}{3 r^2}./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/06%3A_Gauss's_Law/6.04%3A_Applying_Gausss_Law) Now evaluate the surface integral over the boundary \partial V of the ball (sphere of radius R): \iint_{\partial V} \mathbf{F} \cdot d\mathbf{S} = F(R) \cdot 4 \pi R^2 = \frac{\rho R}{3} \cdot 4 \pi R^2 = \rho \cdot \frac{4}{3} \pi R^3./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/06%3A_Gauss's_Law/6.04%3A_Applying_Gausss_Law) By the divergence theorem, this equals the volume integral \iiint_V (\nabla \cdot \mathbf{F}) \, dV = \iiint_V \rho \, dV = \rho \cdot \frac{4}{3} \pi R^3, confirming the equality. Outside the ball, the zero divergence ensures no additional contribution from the exterior region./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/06%3A_Gauss's_Law/6.04%3A_Applying_Gausss_Law) This distributed source example extends the simpler point-charge case (uniform flux over a sphere enclosing a delta-function divergence), where the field is inversely quadratic everywhere outside but here varies linearly inside due to the volume distribution./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/06%3A_Gauss's_Law/6.04%3A_Applying_Gausss_Law)

Historical Development

Early Precursors

The foundations of the divergence theorem were laid in the through investigations into and variational principles, particularly by Leonhard Euler and . Euler's work in the mid-1700s on the provided essential tools for analyzing extremal problems in , including those involving fluid motion and the flow across boundaries. These efforts emphasized geometric and intuitive approaches to integrals over surfaces enclosing fluid volumes, setting the stage for later formalizations without yet establishing a general relation between volume and surface integrals. Joseph-Louis Lagrange built directly on Euler's framework, advancing the study of through surfaces in the context of acoustics and during the . In 1762, Lagrange derived a form of the between flux across a closed surface and sources within the enclosed volume, applied specifically to the propagation of sound waves. His approach, detailed in papers presented to the Academy, focused on variational methods in wave propagation, where surface integrals represented aspects of wave . However, Lagrange provided no general proof, limiting his results to particular cases and assuming idealized conditions. Lagrange extended these ideas in the 1770s through his contributions to within , where he explored gravitational and fluid potentials to describe forces and fluxes in continuous media. In works addressing the of bodies and fluid equilibrium, he employed integral expressions linking distributed sources in a volume to their effects on bounding surfaces, further anticipating the theorem's structure. These developments, while innovative, remained tied to specific mechanical contexts and lacked the abstract generality that would emerge later. A significant self-published work appeared in 1828 with George Green's An Essay on the Application of to the Theories of and . In this privately circulated essay, Green formulated an connecting the of a over a volume to its through the boundary surface, motivated by problems in and magnetostatics. Although not identical to the modern divergence theorem—focusing instead on potential functions and boundary values—this essay encapsulated the core idea of transforming volume integrals of sources into surface integrals, yet it included no rigorous proof and went largely unnoticed until its wider recognition in the 1840s. Green's applied to rectangular coordinates and specific physical distributions, such as charge densities, without broader geometric proofs. Around the same time, Mikhail Ostrogradsky independently derived and published a general proof of the theorem. Ostrogradsky presented his work to the Paris Academy in 1827 and to the St. Petersburg Academy in 1828 (published in 1831), emphasizing its applications to and providing the first rigorous general proof in three dimensions. His contributions highlighted the theorem's utility in physics, particularly for inverse-square laws.

Gauss's Formulation and Proof

In his 1813 publication Theoria attractionis corporum sphaeroidicorum ellipticorum homogeneorum methodo nova tractata, presented a fundamental result relating the of a gravitational attraction force through a closed surface to the enclosed within the volume bounded by that surface. Specifically, he stated that the surface of the normal component of the attraction force over any closed surface equals four times π times the total inside the volume, assuming unit gravitational constant for simplicity in his notation. This formulation, originally derived in the context of for homogeneous ellipsoids, marked Gauss's key contribution to theorems in three dimensions. Gauss's proof sketch relied on transforming the general case to rectangular coordinates, where the volume could be decomposed into elementary rectangular prisms. For each coordinate direction, he applied the to relate the difference in the field components across opposite faces of the prism to the integral of the field's "" (in his scalar notation) over the prism's interior. Summing these contributions over the entire yielded the surface flux, establishing the equality component-wise before generalizing. This approach paralleled later developments but was tailored to the gravitational force derived from the V, where the force components were partial derivatives of V. The original text, written in Latin, employed scalar integrals without modern , denoting the surface flux as ∫ P ds (with P the normal force component) and the volume term involving sums of second partial derivatives of V. Although Gauss's work built briefly on precursors like Lagrange's 1762 flux results for specific potentials, it advanced a more general and rigorous framework applicable beyond isolated cases. Initially recognized as "Gauss's theorem" in continental European mathematical circles for its role in attraction theory, the result was later abstracted as the divergence theorem in the emergence of vector analysis during the late 19th century.

Modern Extensions and Generalizations

In the early , advanced the generalization of the divergence theorem through his development of the calculus of differential forms, culminating in the unified formulation of in 1945, which encompasses the divergence theorem as the case for 3-forms on oriented manifolds. This framework allowed the theorem to be expressed in terms of the and integration over chains, providing a coordinate-free approach applicable to curved spaces. By the mid-20th century, Michael Spivak's 1965 text Calculus on Manifolds formalized these ideas rigorously for smooth manifolds with boundary, proving the generalized divergence theorem as a corollary of and emphasizing its role in on spaces. Spivak's treatment extended the theorem to n-dimensional settings, integrating it with to handle oriented manifolds embedded in . In the post-2000 era, computational extensions have leveraged the divergence theorem in finite element methods (FEM) for approximating solutions on irregular domains, particularly through error estimates that bound discrepancies using volume-to-boundary flux relations. For instance, in H(div)-conforming FEM for elliptic problems, the theorem underpins and analyses on polygonal or curved domains, enabling adaptive mesh refinement with optimal error rates in the L^2-norm. Recent research in the 2020s has explored analogs, such as extended Itô formulas for stochastic differential equations on manifolds, which generalize the divergence theorem to incorporate and provide integration-by-parts rules for anticipating processes in . These developments, rooted in , facilitate applications in and theory on non-Euclidean spaces.

Generalizations Beyond Vectors

Tensor Field Versions

The divergence theorem extends naturally to tensor fields of higher rank, where the divergence operator is defined via contraction with the covariant derivative. For a contravariant tensor field T^i_j of type (1,1), the divergence is a vector field given componentwise by (\operatorname{div} T)^k = \nabla_i T^{i k}, where \nabla_i denotes the covariant derivative. This generalizes the vector case by applying the divergence to each "row" of the tensor, effectively contracting the first index. The integral form of the theorem for such tensor fields states that for a compact oriented manifold V with boundary \partial V, \int_{\partial V} (T \cdot \mathbf{n}) \, dS = \int_V \operatorname{div} T \, dV, where T \cdot \mathbf{n} denotes the tensor contraction T^{i k} n_i (yielding a vector tangent to the boundary), and \mathbf{n} is the outward-pointing unit normal. This equates the flux of the tensor through the boundary surface to the integral of its divergence over the volume. In applications, this tensor version underpins conservation laws in physics. In general relativity, the stress-energy tensor T^{\mu\nu} satisfies \nabla_\mu T^{\mu\nu} = 0, and integrating this over a volume via the divergence theorem yields the conservation of energy-momentum across the . Similarly, in , the equilibrium equation is \operatorname{div} \sigma + \mathbf{f} = 0, where \sigma is the second-order stress tensor and \mathbf{f} is the density; the theorem relates the surface tractions \sigma \cdot \mathbf{n} to the volume integral of internal forces. In component form for a second-order contravariant tensor T^{ij}, the divergence is (\operatorname{div} T)^j = \nabla_i T^{ij}, and the trace of this divergence vector is \operatorname{tr}(\operatorname{div} T) = \nabla_i T^{i i}, which equals the divergence of the tensor's trace scalar \operatorname{tr} T = T^{i i}. This relation connects the vectorial divergence to scalar invariants, preserving their coordinate-independent properties under tensor transformations. The theorem's proof in curved spaces relies on the Riemannian metric to handle index contractions consistently.

Differential Forms Approach

The differential forms approach reformulates the divergence theorem as a special case of the general , which relates the of a over a manifold to the of its over the manifold's interior. This perspective, developed through the theory of exterior calculus, provides a coordinate-free framework that extends naturally beyond . In three dimensions, a \mathbf{F} = F_x \frac{\partial}{\partial x} + F_y \frac{\partial}{\partial y} + F_z \frac{\partial}{\partial z} corresponds to the 2-form \omega = F_x \, dy \wedge dz + F_y \, dz \wedge dx + F_z \, dx \wedge dy. The d\omega is then the 3-form d\omega = \left( \frac{\partial F_x}{\partial x} + \frac{\partial F_y}{\partial y} + \frac{\partial F_z}{\partial z} \right) dx \wedge dy \wedge dz = (\nabla \cdot \mathbf{F}) \, dx \wedge dy \wedge dz, where dx \wedge dy \wedge dz is the oriented . For an oriented compact V with \partial V, yields \int_{\partial V} \omega = \int_V d\omega, which is precisely the divergence theorem equating the of \mathbf{F} through \partial V to the of its over V. This formulation generalizes to n dimensions, where a on \mathbb{R}^n associates to an (n-1)-form \omega via the metric-induced between vectors and covectors, and the d\omega equals (\operatorname{div} \mathbf{F}) times the volume form. The resulting identity \int_{\partial M} \omega = \int_M d\omega for an oriented n-manifold M with \partial M captures the divergence theorem in this setting. The primary advantages of this approach lie in its unification of vector calculus identities: the gradient corresponds to the exterior derivative on 0-forms, the curl to the derivative on 1-forms, and the divergence to the derivative on (n-1)-forms, all under the single Stokes' theorem. Moreover, it applies seamlessly to oriented manifolds without requiring a flat metric, facilitating extensions to curved spaces such as Riemannian manifolds.

Higher-Dimensional and Abstract Settings

The divergence theorem extends to infinite-dimensional settings within functional analysis, particularly in Hilbert and Banach spaces, where classical notions of divergence are replaced by weak formulations to handle the lack of a natural inner product or trace in infinite dimensions. In abstract Wiener spaces, which consist of a Banach space B with a densely embedded Hilbert space H and a Gaussian measure, a divergence theorem relates the integral of the weak divergence of a vector field to a boundary term via Wiener measure. Specifically, for a measurable, H-differentiable function F: V \cup \partial V \to H on a set V with H-C¹ boundary, the theorem states that the expected value of the trace of the derivative operator adjusted by a test operator equals the surface integral over the boundary using normal surface measure under the Wiener process. This weak divergence is defined through integration by parts in the Hilbert directions, enabling applications to stochastic processes like the Ornstein-Uhlenbeck semigroup. A variant on Wiener spaces W = C([0,1], \mathbb{R}) with Gaussian measure provides a Gauss-type formula for subsets with boundaries, linking the inner product of the derivative to a boundary correction involving the time of minimum and Dirac measure at the boundary point. On Lie groups, the divergence theorem adapts to the group structure using left-invariant s, which are unique up to scaling and preserve the group's multiplication. For a G equipped with a left-invariant \mathrm{vol}_G, the of a X is defined such that the L_X \mathrm{dvol}_G = (\mathrm{div}_G X) \mathrm{dvol}_G. For left-invariant vector fields X \in \mathfrak{g}, this equals the of the \mathrm{trace}(\mathrm{ad} X), modulated by the modular function \mu_G of the group, which is 1 for unimodular groups like compact or semisimple s. The theorem then asserts that for compactly supported smooth functions, the integral of \mathrm{div}_G X over an \Omega \subset G equals the flux through the boundary, with the ensuring left-invariance. This formulation unifies on groups like SO(3) or Heisenberg groups, where sub-Laplacians \Delta_G u = \mathrm{div}_G (\nabla_G u) arise naturally from horizontal gradients on bracket-generating subspaces. In non-compact spaces, such as complete s without , the divergence theorem holds under or integrability conditions to compensate for the absence of compactness. For instance, on a non-compact complete M with metric g, if a C^1 X satisfies uniform |X| \to 0 at , then \int_M \mathrm{div} X \, \mathrm{d}\nu_g = 0, where \nu_g is the volume measure. More generally, without but with recurrent flow on the unit and integrability of the function f_X(p,v) = g(\nabla_v X, v), the of \mathrm{div} X vanishes, extending the theorem to infinite-volume cases like . For fields with compact support, the classical flux equality persists, while conditions at ensure the term at "infinity" vanishes, as in \mathbb{R}^n where \int_{\mathbb{R}^n} \mathrm{div} F = 0 for compactly supported F. Recent developments in the 2020s explore analogs of the divergence theorem in on curved spacetimes, where anomalies link to violations of classical laws derived from the theorem. In interacting quantum field theories, Weyl anomalies induce non-vanishing traces in the stress-energy tensor, analogous to how divergence-free conditions fail due to quantum effects, with the anomaly coefficients determined by local curvature invariants. These analogs appear in renormalization schemes for fields coupled to external two-forms, where nonlocal terms in the preserve invariance but reveal anomaly-induced divergences in curved backgrounds, connecting to chiral and gravitational anomalies. Such frameworks, built on heat-kernel methods, extend the theorem's role in deriving identities to settings in asymptotically flat or Anti-de Sitter spacetimes.

References

  1. [1]
    Calculus III - Divergence Theorem - Pauls Online Math Notes
    Nov 16, 2022 · In this section we are going to relate surface integrals to triple integrals. We will do this with the Divergence Theorem. Divergence Theorem.
  2. [2]
    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 ...
  3. [3]
    16.9 The Divergence Theorem
    1 (Divergence Theorem) Under suitable conditions, if E is a region of three dimensional space and D is its boundary surface, oriented outward, then ∫∫DF⋅NdS=∫∫∫ ...Missing: definition | Show results with:definition
  4. [4]
    4.2 The Divergence Theorem
    The divergence theorem expresses the integral of a derivative of a function (in this case a vector-valued function) over a region in terms of the values of the ...
  5. [5]
    The fundamental theorems of vector calculus - Math Insight
    The divergence theorem then says that the total expansion of the fluid in W is equal to the total flux of the fluid out of the boundary S=∂W. Thread navigation.
  6. [6]
    [PDF] Divergence-measure fields: Gauss-Green formulas and Normal Traces
    The formula that would be later known as the divergence theorem was first discovered by. Lagrange2 in 1762 (see Fig. 2), but he did not provide a proof of the ...
  7. [7]
    [PDF] The History of Stokes' Theorem - Harvard Mathematics Department
    The first of these theorems to be stated and proved in essentially its present form was the one known today as Gauss' theorem or the divergence theorem. In ...
  8. [8]
    [PDF] 11–Applications of the Divergence Theorem - UC Davis Mathematics
    Derivation of the Compressible Euler Equations. In this section we use the divergence theorem to derive a physical inter- pretation of the compressible ...
  9. [9]
    [PDF] Lesson 10. Applications of Divergence Theorem - Purdue Math
    Applications of Divergence Theorem. Consider incompressible fluid of constant density ρ in a region T with the boundary S. We take ρ ≡ 1. Let v be the ...
  10. [10]
    14.1 Application of the Divergence Theorem: Laplace's Equation
    The method used in this argument is a very important and general one that is used in dealing with many differential equations. In fact the use of the divergence ...
  11. [11]
    15.7 The Divergence Theorem and Stokes' Theorem
    It states, in words, that the flux across a closed surface equals the sum of the divergences over the domain enclosed by the surface.
  12. [12]
    6.8 The Divergence Theorem - Calculus Volume 3 | OpenStax
    Mar 30, 2016 · Use the divergence theorem to calculate flux ... We can now use the divergence theorem to justify the physical interpretation of divergence that ...
  13. [13]
    3D divergence theorem (article) | Khan Academy
    ### Summary of 3D Divergence Theorem (Khan Academy)
  14. [14]
    The idea behind the divergence theorem - Math Insight
    Recall that if a vector field F represents the flow of a fluid, then the divergence of F represents the expansion or compression of the fluid. The divergence ...<|control11|><|separator|>
  15. [15]
  16. [16]
    Divergence - Richard Fitzpatrick
    ... fluid flow out of $ S$ . If $ {\bf A}$ is constant in space then it is ... divergence theorem. This states that for any volume $ V$ surrounded by a ...
  17. [17]
  18. [18]
    [PDF] M342 PDE: THE DIVERGENCE THEOREM Let R be a bounded ...
    STATEMENT OF THE DIVERGENCE THEOREM. Let R be a bounded open subset of Rn with smooth (or piecewise smooth) boundary. ∂R. Let X = (X1,...,Xn) be a smooth ...
  19. [19]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    The divergence of a vector field X = (X1,X2,...,Xn):Ω → Rn is div X = ∂X1 ... quire that the boundary of the domain is smooth. A solution of an ...
  20. [20]
    [PDF] 0 Introduction - Penn Math
    We also suppose that the boundary is a piecewise smooth curve; that is, the boundary ∂B consists of a finite number of smooth curves. Such regions are called ...Missing: compact | Show results with:compact
  21. [21]
    [PDF] Line Integrals and Green's Theorem - MIT OpenCourseWare
    For an arbitrary simple closed curves, showing that it has a well-defined interior is more subtle. The theorem that proves this is called the Jordan curve ...<|control11|><|separator|>
  22. [22]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    These notes cover preliminaries, Laplace's equation, Sobolev spaces, elliptic PDEs, and the heat equation.
  23. [23]
    [PDF] 4 The Integral Theorems - DAMTP
    The 2d divergence theorem is the statement that the left-hand sides of (4.13) and (4.14) are equal; Green's theorem in the plane is the statement that the right ...Missing: Euclidean | Show results with:Euclidean
  24. [24]
    [PDF] Unit 24: Divergence Theorem
    The theorem explains what divergence means. If we integrate the divergence over a small cube, it is equal the flux of the field through the boundary of the cube ...
  25. [25]
    4.2 The Divergence Theorem
    The divergence theorem is going to relate a volume integral over a solid V to a flux integral over the surface of . V . First we need a couple of definitions ...
  26. [26]
    [PDF] REAL ANALYSIS LECTURE NOTES 5. Divergence theorem and ...
    Theorem 5.1 (Divergence theorem). Let Ω be a bounded domain in Rn with the boundary of class. C1. Let ~F = (F1, ...., Fn) be a a vector field of class C(Ω) ...
  27. [27]
    [PDF] Topics in Integration - John McCuan
    Mar 19, 2021 · This is accomplished by using Fubini's theorem in a different way: Taking. Jx = {y : 0 <y<. √ x} for 0 <x< 1, we have. ∫Uf = ∫x∈(0,1) ∫y ...
  28. [28]
    [PDF] Math 865, Topics in Riemannian Geometry - UCI Mathematics
    4.4 The divergence theorem for a Riemannian manifold. We begin with a useful formula for the divergence of a vector field. Proposition 4.7. For a vector ...
  29. [29]
  30. [30]
    [PDF] MATH 332: Vector Analysis Formulas
    MATH 332: Vector Analysis. Formulas. Vector Algebra x1 = x, x2 = y, x3 = z e1 = i ... Vector identities: ∇×∇ϕ = 0, curl grad = 0. ∇ · (∇ × F)=0, div curl = 0.Missing: calculus | Show results with:calculus
  31. [31]
    [PDF] 8 Vector Analysis and EM Waves - UNCW
    + µ0∇ · J = 0. As you can see, the vector identity, div curl = 0, does not work here! Maxwell argued that we need to account for a changing charge.
  32. [32]
    Helmholtz's theorem - Richard Fitzpatrick
    This strongly suggests that if we know the divergence and the curl of a vector field then we know everything there is to know about the field. In fact, this is ...
  33. [33]
    [PDF] Helmholtz Decomposition of Vector Fields
    We can use the divergence theorem on the first term to turn it into a surface integral over the surface S bounding volume V. If we let V be all space, then ...
  34. [34]
    [PDF] 12 Green's first identity Having studied Laplace's equation in regions ...
    Starting from the divergence theorem we derived Green's first identity (2), which can be thought of as integration by parts in higher dimensions. Using this ...
  35. [35]
    [PDF] The Divergence Theorem - Purdue Math
    Theorem from the Divergence Theorem). Green's First Identity. We can use use the Divergece Theorem to derive the following useful formula. Let E be a domain.
  36. [36]
    [PDF] 13 Green's second identity, Green's functions - UCSB Math
    We derived Green's second identity from Green's first identity, which was subsequently applied to the pair of a harmonic function and the fundamental radial ...
  37. [37]
    [PDF] 7 Green's Functions and Nonhomogeneous Problems - UNCW
    green's functions and nonhomogeneous problems 283 c. Use Green's first identity to prove Green's second identity,. Z. D. (u∇2v − v∇2u) dA = Z. C. (u∇v − v ...
  38. [38]
    [PDF] HARMONIC FUNCTIONS, GREEN'S FUNCTIONS and POTENTIALS.
    Definition. Green's function in Rn with pole x0 is defined as: Gx0 (x) = 1 (n − 2)ωn−1|x − x0|n−2 (n ≥ 3), Gx0 (x) = − 1 2π log |x−x0|(n = 2).
  39. [39]
    [PDF] Divergence theorem - Harvard Mathematics Department
    4) For divergence free vector fields F, the flux through a closed surface is zero. Such fields F are also called incompressible or source free. Example. Compute ...
  40. [40]
    [PDF] Lecture 27: More about the divergence theorem - Faculty Web Pages
    The flux of such a vector field across the boundary of any solid region is 0, because the divergence theorem turns that flux integral into an integral of 0 over ...
  41. [41]
    [PDF] Applications of Mathematical Theorems and laws in Electrical ...
    ... Kirchhoff's laws are derived by the applications of Mathematical Theorem and laws, Divergence Theorem, Gauss's law. Moreover, laws, that is a relation.
  42. [42]
    [PDF] Vector Field Analysis Other Features
    Gauss Theorem. • Also known as divergence theorem, that relates the vectors on the boundary ! = " of a region ! to the divergence in the region. # V ! = ! # V ...<|control11|><|separator|>
  43. [43]
    [PDF] The Continuity Equation
    This is the velocity divergence form of the continuity equation that states the fractional rate of change of mass per unit volume following the motion is equal ...
  44. [44]
    [PDF] Electric Charge Conservation and the Continuity Equation
    This continuity equation is the local form of the Law of Electric Charge Conservation, and it always holds true, for every physical system of charges and ...
  45. [45]
    Continuity Equation – Introduction to Aerospace Flight Vehicles
    Therefore, this latter equation states that to satisfy the conservation of mass, the divergence of the local velocity field must be identically zero. If not, ...
  46. [46]
    [PDF] Gauss's Law - UTK-EECS
    Gauss's law: The divergence of electric field at each point is proportional to the local charge density. holds for any arbitrary S. Page 9. Gauss's theorem in ...
  47. [47]
    [PDF] Magnetostatics – Magnetic Flux Density
    Point form of Gauss's Law​​ The divergence theorem states that the net outflow of flux from a closed surface is equal to the sum of flux outflow (and inflow) ...
  48. [48]
    [PDF] Unit 34: Gauss theorem
    34.2. Gauss law div(F) = f = 4πGρ describes the gravitational field induced from a mass density ρ and gravitational constant G. The picture is that mass is a ...Missing: magnetism | Show results with:magnetism
  49. [49]
    [PDF] gauss-law.pdf
    The divergence theorem discovered by Gauss links the integral version of Gauss's law (Physics 212) with this differential version. We will frequently use ...
  50. [50]
    [PDF] 1.4 Potential theory
    POTENTIAL THEORY. 13. 1.4 Potential theory. The central result of Newton's gravitational theory is the inverse square law for the force between two point ...
  51. [51]
    Green's Representation Theorem — The Bempp Book
    Green's identities¶. Based on the divergence theorem, we can now derive the Green's identities. ... y). Finally, the third term is the Newton potential operator.
  52. [52]
    [PDF] Multipole Expansion of the Electrostatic Potential - UT Physics
    (3). Multipole Expansion of the Coulomb Potential. Now consider the Coulomb potential of some continuous charge distribution ρ(r),.
  53. [53]
    [PDF] 15.10 Gauss/Divergence Theorem The ... - Contemporary Calculus
    Example 2: Calculate the flux across the sphere x. 2 + y2 + z2 = R2 for the ... field F = ax, by, cz . Example 3: Calculate the outward flux across the boundary D ...
  54. [54]
    [PDF] The Divergence Theorem
    May 3, 2018 · Example. Find the flux of the unit radial vector field ~F = r |r| outward through the surface of the sphere x2 + y2 + z2 ≤ a2. Let R be the ...
  55. [55]
    [PDF] Lectures on Electromagnetic Field Theory
    Feb 8, 2023 · This set of lecture notes is from my teaching of ECE 604, Electromagnetic Field Theory, at. ECE, Purdue University, West Lafayette.
  56. [56]
    [PDF] The Divergence Theorem Based on lecture notes by James ...
    The Divergence Theorem is also known as Gauss's Theorem, and as Ostrograd- sky's Theorem. It was first discovered by Lagrange in 1762, and then independently.
  57. [57]
    Joseph-Louis Lagrange - Biography
    ### Summary of Lagrange's Contributions
  58. [58]
    Why did George Green Write His Essay of 1828 on Electricity ... - jstor
    Secondly was Green's type of divergence theorem, expressed entirely within the rectangular co-ordinate system (x, y, z) rather than with surface ...
  59. [59]
  60. [60]
    A history of the divergence theorem
    ### Summary of Early History of the Divergence Theorem
  61. [61]
    [PDF] the generalized stokes' theorem
    It was then vastly generalized in 1945 by Élie Cartan into its modern form, the generalized Stokes' Theo- rem, a result that spans several theorems of vector ...
  62. [62]
    [PDF] spivak-calculus-on-manifolds.pdf - Cimat
    Spivak's book should be a help to those who wish to see. Stoke's Theorem as the modern working mathematician sees it. A student with a good course in calculus ...
  63. [63]
    [PDF] Calculus on Manifolds - Strange beautiful grass of green
    Generalize the divergence theorem to the case of an n-manifold with boundary in nn. 5-35. Applying the generalized divergence theorem to the set M = {x E Rn ...
  64. [64]
    [PDF] Convergence analysis of Finite Element Methods for H(div;Ω)
    May 10, 2010 · Optimal error estimates in the H(div; Ω)-norms are obtained for the first time. The analysis is based on a so-called δ-strip argument, a new ...
  65. [65]
    [PDF] hexahedral h(div) and h(curl) finite elements
    The finite element spaces studied are defined on irregular ... and then use these results in § 8 to obtain error estimates for the finite element H(div; Ω).
  66. [66]
    Intrinsic Stochastic Differential Equations and Extended Ito Formula ...
    Nov 16, 2022 · Abstract:A general way of representing Stochastic Differential Equations (SDEs) on smooth manifold is based on Schwartz morphism.
  67. [67]
    Ito's formula for Gaussian processes with stochastic discontinuities
    Ys = Ys − Ys−. The analogous notation is used for the jumps of deterministic regulated functions. Weakly regulated processes can be characterized via the S- ...<|control11|><|separator|>
  68. [68]
    [PDF] 1.2 The divergence theorem
    Tensor Fields Let Ω ⊂ IRn be an open domain. • ϕ : Ω → IR is a scalar field;. • v : Ω → IRm is a vector field;. • T : Ω → Lin(IRm, IRk) is a tensor field.
  69. [69]
    [PDF] 1.14 Tensor Calculus I: Tensor Fields
    1.14.5 The Divergence Theorem. The divergence theorem 1.7.12 can be extended to the case of higher-order tensors. Consider an arbitrary differentiable tensor ...
  70. [70]
    [PDF] Stress Energy Tensor
    This shows that the divergence free stress energy tensor T together with any Killing field. X leads to a divergence free vector fields V = T · X, i.e. to vector ...
  71. [71]
    [PDF] Chapter 11: Elastostatics [version 1211.1.K] - Caltech PMA
    divergence of its elastic stress tensor is equal to the gravitational force ... – Bulk and shear moduli K, µ; elastic stress tensor T = −KΘg − 2µΣ, Sec.
  72. [72]
    tensor fields on a semi-Riemannian manifold with boundary
    We prove in this paper a divergence theorem for symmetric (0,2)-tensors on a semi-Riemannian manifold with boundary. We obtain a generalization of results ...
  73. [73]
  74. [74]
    None
    ### Summary of Divergence Theorem or Divergence for Vector Fields on Lie Groups Using Haar Measure
  75. [75]
    [PDF] arXiv:1612.08189v1 [math.DG] 24 Dec 2016
    Dec 24, 2016 · In this paper we use a dynamical approach to prove some new divergence theorems on complete non-compact Riemannian manifolds. 1. Introduction ...
  76. [76]
    [2504.17854] The Weyl anomaly in interacting quantum field theory ...
    Apr 24, 2025 · We discuss some general properties of Weyl anomalies, such as their relation to the trace anomaly. We give a criterion for a theory to be ...Missing: divergence analogs spacetime 2020s