Divergence theorem


In vector calculus, the divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a result that relates the flux of a vector field through a closed surface to the divergence of the field in the volume enclosed.
More precisely, the divergence theorem states that the surface integral of a vector field over a closed surface, which is called the flux through the surface, is equal to the volume integral of the divergence over the region inside the surface. Intuitively, it states that the sum of all sources of the field in a region gives the net flux out of the region.
The divergence theorem is an important result for the mathematics of physics and engineering, in particular in electrostatics and fluid dynamics.
In physics and engineering, the divergence theorem is usually applied in three dimensions. However, it generalizes to any number of dimensions. In one dimension, it is equivalent to integration by parts. In two dimensions, it is equivalent to Green's theorem.

Explanation using liquid flow

s are often illustrated using the example of the velocity field of a fluid, such as a gas or liquid. A moving liquid has a velocity—a speed and a direction—at each point, which can be represented by a vector, so that the velocity of the liquid forms a vector field. Consider an imaginary closed surface S inside a body of liquid, enclosing a volume of liquid. The flux of liquid out of the volume is equal to the volume rate of fluid crossing this surface, i.e., the surface integral of the velocity over the surface.
Since liquids are incompressible, the amount of liquid inside a closed volume is constant; if there are no sources or sinks inside the volume then the flux of liquid out of S is zero. If the liquid is moving, it may flow into the volume at some points on the surface S and out of the volume at other points, but the amounts flowing in and out at any moment are equal, so the net flux of liquid out of the volume is zero.
However if a source of liquid is inside the closed surface, such as a pipe through which liquid is introduced, the additional liquid will exert pressure on the surrounding liquid, causing an outward flow in all directions. This will cause a net outward flow through the surface S. The flux outward through S equals the volume rate of flow of fluid into S from the pipe. Similarly if there is a sink or drain inside S, such as a pipe which drains the liquid off, the external pressure of the liquid will cause a velocity throughout the liquid directed inward toward the location of the drain. The volume rate of flow of liquid inward through the surface S equals the rate of liquid removed by the sink.
If there are multiple sources and sinks of liquid inside S, the flux through the surface can be calculated by adding up the volume rate of liquid added by the sources and subtracting the rate of liquid drained off by the sinks. The volume rate of flow of liquid through a source or sink is equal to the divergence of the velocity field at the pipe mouth, so adding up the divergence of the liquid throughout the volume enclosed by S equals the volume rate of flux through S. This is the divergence theorem.
The divergence theorem is employed in any conservation law which states that the total volume of all sinks and sources, that is the volume integral of the divergence, is equal to the net flow across the volume's boundary.

Mathematical statement

Suppose is a subset of which is compact and has a piecewise smooth boundary . If is a continuously differentiable vector field defined on a neighborhood of, then:
The left side is a volume integral over the volume, the right side is the surface integral over the boundary of the volume. The closed manifold is oriented by outward-pointing normals, and is the outward pointing unit normal at each point on the boundary. In terms of the intuitive description above, the left-hand side of the equation represents the total of the sources in the volume, and the right-hand side represents the total flow across the boundary.

Informal derivation

The divergence theorem follows from the fact that if a volume is partitioned into separate parts, the flux out of the original volume is equal to the sum of the flux out of each component volume. This is true despite the fact that the new subvolumes have surfaces that were not part of the original volume's surface, because these surfaces are just partitions between two of the subvolumes and the flux through them just passes from one volume to the other and so cancels out when the flux out of the subvolumes is summed.

See the diagram. A closed, bounded volume is divided into two volumes and by a surface . The flux out of each component region is equal to the sum of the flux through its two faces, so the sum of the flux out of the two parts is
where and are the flux out of surfaces and, is the flux through out of volume 1, and is the flux through out of volume 2. The point is that surface is part of the surface of both volumes. The "outward" direction of the normal vector is opposite for each volume, so the flux out of one through is equal to the negative of the flux out of the other
so these two fluxes cancel in the sum. Therefore
Since the union of surfaces and is


This principle applies to a volume divided into any number of parts, as shown in the diagram. Since the integral over each internal partition appears with opposite signs in the flux of the two adjacent volumes they cancel out, and the only contribution to the flux is the integral over the external surfaces . Since the external surfaces of all the component volumes equal the original surface.


The flux out of each volume is the surface integral of the vector field over the surface
The goal is to divide the original volume into infinitely many infinitesimal volumes. As the volume is divided into smaller and smaller parts, the surface integral on the right, the flux out of each subvolume, approaches zero because the surface area approaches zero. However, from the definition of divergence, the ratio of flux to volume,, the part in parentheses below, does not in general vanish but approaches the divergence as the volume approaches zero.
As long as the vector field has continuous derivatives, the sum above holds even in the limit when the volume is divided into infinitely small increments
As approaches zero volume, it becomes the infinitesimal, the part in parentheses becomes the divergence, and the sum becomes a volume integral over
Since this derivation is coordinate free, it shows that the divergence does not depend on the coordinates used.

Corollaries

By replacing in the divergence theorem with specific forms, other useful identities can be derived.
Suppose we wish to evaluate
where is the unit sphere defined by
and is the vector field
The direct computation of this integral is quite difficult, but we can simplify the derivation of the result using the divergence theorem, because the divergence theorem says that the integral is equal to:
where is the unit ball:
Since the function is positive in one hemisphere of and negative in the other, in an equal and opposite way, its total integral over is zero. The same is true for :
Therefore,
because the unit ball has volume.

Applications

''Differential form'' and ''integral form'' of physical laws

As a result of the divergence theorem, a host of physical laws can be written in both a differential form and an integral form. Three examples are Gauss's law, Gauss's law for magnetism, and Gauss's law for gravity.

Continuity equations

s offer more examples of laws with both differential and integral forms, related to each other by the divergence theorem. In fluid dynamics, electromagnetism, quantum mechanics, relativity theory, and a number of other fields, there are continuity equations that describe the conservation of mass, momentum, energy, probability, or other quantities. Generically, these equations state that the divergence of the flow of the conserved quantity is equal to the distribution of sources or sinks of that quantity. The divergence theorem states that any such continuity equation can be written in a differential form and an integral form.

Inverse-square laws

Any inverse-square law can instead be written in a Gauss's law-type form. Two examples are Gauss's law, which follows from the inverse-square Coulomb's law, and Gauss's law for gravity, which follows from the inverse-square Newton's law of universal gravitation. The derivation of the Gauss's law-type equation from the inverse-square formulation or vice versa is exactly the same in both cases; see either of those articles for details.

History

The theorem was first discovered by Lagrange in 1762, then later independently rediscovered by Gauss in 1813, by Ostrogradsky, who also gave the first proof of the general theorem, in 1826, by Green in 1828, Simeon-Denis Poisson in 1824 and Frédéric Sarrus in 1828.

Worked examples

Example 1

To verify the planar variant of the divergence theorem for a region :
and the vector field:
The boundary of is the unit circle,, that can be represented parametrically by:
such that where units is the length arc from the point to the point on. Then a vector equation of is
At a point on :
Therefore,
Because, and because. Thus

Example 2

Let's say we wanted to evaluate the flux of the following vector field defined by bounded by the following inequalities:
By the divergence theorem,
We now need to determine the divergence of. If is a three-dimensional vector field, then the divergence of is given by.
Thus, we can set up the following the flux integral
as follows:
Now that we have set up the integral, we can evaluate it.

Generalizations

Multiple dimensions

One can use the general Stokes' Theorem to equate the -dimensional volume integral of the divergence of a vector field over a region to the -dimensional surface integral of over the boundary of :
This equation is also known as the divergence theorem.
When, this is equivalent to Green's theorem.
When, it reduces to integration by parts.

Tensor fields

Writing the theorem in Einstein notation:
suggestively, replacing the vector field with a rank- tensor field, this can be generalized to:
where on each side, tensor contraction occurs for at least one index. This form of the theorem is still in 3d, each index takes values 1, 2, and 3. It can be generalized further still to higher dimensions.