fbpx
Wikipedia

Divergence theorem

In vector calculus, the divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem,[1] is a theorem relating 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 enclosed by the surface. Intuitively, it states that "the sum of all sources of the field in a region (with sinks regarded as negative sources) gives the net flux out of the region".

The divergence theorem is an important result for the mathematics of physics and engineering, particularly in electrostatics and fluid dynamics. In these fields, it is usually applied in three dimensions. However, it generalizes to any number of dimensions. In one dimension, it is equivalent to the fundamental theorem of calculus. In two dimensions, it is equivalent to Green's theorem.

Explanation using liquid flow edit

Vector fields 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 at any moment 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 at any time 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 (with the flow through a sink given a negative sign) is equal to the divergence of the velocity field at the pipe mouth, so adding up (integrating) the divergence of the liquid throughout the volume enclosed by S equals the volume rate of flux through S. This is the divergence theorem.[2]

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.[3]

Mathematical statement edit

 
A region V bounded by the surface   with the surface normal n

Suppose V is a subset of   (in the case of n = 3, V represents a volume in three-dimensional space) which is compact and has a piecewise smooth boundary S (also indicated with  ). If F is a continuously differentiable vector field defined on a neighborhood of V, then:[4][5]

      

The left side is a volume integral over the volume V, the right side is the surface integral over the boundary of the volume V. The closed manifold   is oriented by outward-pointing normals, and   is the outward pointing unit normal at each point on the boundary  . (  may be used as a shorthand for  .) In terms of the intuitive description above, the left-hand side of the equation represents the total of the sources in the volume V, and the right-hand side represents the total flow across the boundary S.

Informal derivation edit

The divergence theorem follows from the fact that if a volume V 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.[6][7] 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.

 
A volume divided into two subvolumes. At right the two subvolumes are separated to show the flux out of the different surfaces.

See the diagram. A closed, bounded volume V is divided into two volumes V1 and V2 by a surface S3 (green). The flux Φ(Vi) out of each component region Vi 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 Φ1 and Φ2 are the flux out of surfaces S1 and S2, Φ31 is the flux through S3 out of volume 1, and Φ32 is the flux through S3 out of volume 2. The point is that surface S3 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 S3 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 S1 and S2 is S

 
 
The volume can be divided into any number of subvolumes and the flux out of V is equal to the sum of the flux out of each subvolume, because the flux through the green surfaces cancels out in the sum. In (b) the volumes are shown separated slightly, illustrating that each green partition is part of the boundary of two adjacent volumes

This principle applies to a volume divided into any number of parts, as shown in the diagram.[7] Since the integral over each internal partition (green surfaces) 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 (grey). Since the external surfaces of all the component volumes equal the original surface.

 
 
As the volume is subdivided into smaller parts, the ratio of the flux   out of each volume to the volume   approaches  

The flux Φ out of each volume is the surface integral of the vector field F(x) 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 S(Vi) 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 div F as the volume approaches zero.[7]

 

As long as the vector field F(x) 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 dV, the part in parentheses becomes the divergence, and the sum becomes a volume integral over V

 

Since this derivation is coordinate free, it shows that the divergence does not depend on the coordinates used.

Proofs edit

For bounded open subsets of Euclidean space edit

We are going to prove the following:

Theorem — Let   be open and bounded with   boundary. If   is   on an open neighborhood   of  , that is,  , then for each  ,

 
where   is the outward pointing unit normal vector to  . Equivalently,
 

Proof of Theorem. [8] (1) The first step is to reduce to the case where  . Pick   such that   on  . Note that   and   on  . Hence it suffices to prove the theorem for  . Hence we may assume that  .

(2) Let   be arbitrary. The assumption that   has   boundary means that there is an open neighborhood   of   in   such that   is the graph of a   function with   lying on one side of this graph. More precisely, this means that after a translation and rotation of  , there are   and   and a   function  , such that with the notation

 
it holds that
 
and for  ,
 
Since   is compact, we can cover   with finitely many neighborhoods   of the above form. Note that   is an open cover of  . By using a   partition of unity subordinate to this cover, it suffices to prove the theorem in the case where either   has compact support in   or   has compact support in some  . If   has compact support in  , then for all  ,   by the fundamental theorem of calculus, and   since   vanishes on a neighborhood of  . Thus the theorem holds for   with compact support in  . Thus we have reduced to the case where   has compact support in some  .

(3) So assume   has compact support in some  . The last step now is to show that the theorem is true by direct computation. Change notation to  , and bring in the notation from (2) used to describe  . Note that this means that we have rotated and translated  . This is a valid reduction since the theorem is invariant under rotations and translations of coordinates. Since   for   and for  , we have for each   that

 
For   we have by the fundamental theorem of calculus that
 
Now fix  . Note that
 
Define   by  . By the chain rule,
 
But since   has compact support, we can integrate out   first to deduce that
 
Thus
 
In summary, with   we have
 
Recall that the outward unit normal to the graph   of   at a point   is   and that the surface element   is given by  . Thus
 
This completes the proof.

For compact Riemannian manifolds with boundary edit

We are going to prove the following:

Theorem — Let   be a   compact manifold with boundary with   metric tensor  . Let   denote the manifold interior of   and let   denote the manifold boundary of  . Let   denote   inner products of functions and   denote inner products of vectors. Suppose   and   is a   vector field on  . Then

 
where   is the outward-pointing unit normal vector to  .

Proof of Theorem. [9] We use the Einstein summation convention. By using a partition of unity, we may assume that   and   have compact support in a coordinate patch  . First consider the case where the patch is disjoint from  . Then   is identified with an open subset of   and integration by parts produces no boundary terms:

 
In the last equality we used the Voss-Weyl coordinate formula for the divergence, although the preceding identity could be used to define   as the formal adjoint of  . Now suppose   intersects  . Then   is identified with an open set in  . We zero extend   and   to   and perform integration by parts to obtain
 
where  . By a variant of the straightening theorem for vector fields, we may choose   so that   is the inward unit normal   at  . In this case   is the volume element on   and the above formula reads
 
This completes the proof.

Corollaries edit

By replacing F in the divergence theorem with specific forms, other useful identities can be derived (cf. vector identities).[10]

  • With   for a scalar function g and a vector field F,
      
A special case of this is  , in which case the theorem is the basis for Green's identities.
  • With   for two vector fields F and G, where   denotes a cross product,
      
  • With   for two vector fields F and G, where   denotes a dot product,
      
  • With   for a scalar function f and vector field c:[11]
      
The last term on the right vanishes for constant   or any divergence free (solenoidal) vector field, e.g. Incompressible flows without sources or sinks such as phase change or chemical reactions etc. In particular, taking   to be constant:
      
  • With   for vector field F and constant vector c:[11]
      
By reordering the triple product on the right hand side and taking out the constant vector of the integral,
      
Hence,
      

Example edit

 
The vector field corresponding to the example shown. Vectors may point into or out of the sphere.
 
The divergence theorem can be used to calculate a flux through a closed surface that fully encloses a volume, like any of the surfaces on the left. It can not directly be used to calculate the flux through surfaces with boundaries, like those on the right. (Surfaces are blue, boundaries are red.)

Suppose we wish to evaluate

    

where S is the unit sphere defined by

 

and F 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 W is the unit ball:

 

Since the function y is positive in one hemisphere of W and negative in the other, in an equal and opposite way, its total integral over W is zero. The same is true for z:

 

Therefore,

    

because the unit ball W has volume 4π/3.

Applications edit

Differential and integral forms of physical laws edit

As a result of the divergence theorem, a host of physical laws can be written in both a differential form (where one quantity is the divergence of another) and an integral form (where the flux of one quantity through a closed surface is equal to another quantity). Three examples are Gauss's law (in electrostatics), Gauss's law for magnetism, and Gauss's law for gravity.

Continuity equations edit

Continuity equations 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 (in terms of a divergence) and an integral form (in terms of a flux).[12]

Inverse-square laws edit

Any inverse-square law can instead be written in a Gauss's law-type form (with a differential and integral form, as described above). Two examples are Gauss's law (in electrostatics), 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.[12]

History edit

Joseph-Louis Lagrange introduced the notion of surface integrals in 1760 and again in more general terms in 1811, in the second edition of his Mécanique Analytique. Lagrange employed surface integrals in his work on fluid mechanics.[13] He discovered the divergence theorem in 1762.[14]

Carl Friedrich Gauss was also using surface integrals while working on the gravitational attraction of an elliptical spheroid in 1813, when he proved special cases of the divergence theorem.[15][13] He proved additional special cases in 1833 and 1839.[16] But it was Mikhail Ostrogradsky, who gave the first proof of the general theorem, in 1826, as part of his investigation of heat flow.[17] Special cases were proven by George Green in 1828 in An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism,[18][16] Siméon Denis Poisson in 1824 in a paper on elasticity, and Frédéric Sarrus in 1828 in his work on floating bodies.[19][16]

Worked examples edit

Example 1 edit

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  , we can evaluate  , and because  ,  . Thus

 

Example 2 edit

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 flux integral        as follows:

 

Now that we have set up the integral, we can evaluate it.

 

Generalizations edit

Multiple dimensions edit

One can use the generalised Stokes' theorem to equate the n-dimensional volume integral of the divergence of a vector field F over a region U to the (n − 1)-dimensional surface integral of F over the boundary of U:

 

This equation is also known as the divergence theorem.

When n = 2, this is equivalent to Green's theorem.

When n = 1, it reduces to the fundamental theorem of calculus, part 2.

Tensor fields edit

Writing the theorem in Einstein notation:

      

suggestively, replacing the vector field F with a rank-n tensor field T, this can be generalized to:[20]

      

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 (or lower) dimensions (for example to 4d spacetime in general relativity[21]).

See also edit

References edit

  1. ^ Katz, Victor J. (1979). "The history of Stokes's theorem". Mathematics Magazine. 52 (3): 146–156. doi:10.2307/2690275. JSTOR 2690275. reprinted in Anderson, Marlow (2009). Who Gave You the Epsilon?: And Other Tales of Mathematical History. Mathematical Association of America. pp. 78–79. ISBN 978-0-88385-569-0.
  2. ^ R. G. Lerner; G. L. Trigg (1994). Encyclopaedia of Physics (2nd ed.). VHC. ISBN 978-3-527-26954-9.
  3. ^ Byron, Frederick; Fuller, Robert (1992), Mathematics of Classical and Quantum Physics, Dover Publications, p. 22, ISBN 978-0-486-67164-2
  4. ^ Wiley, C. Ray Jr. Advanced Engineering Mathematics, 3rd Ed. McGraw-Hill. pp. 372–373.
  5. ^ Kreyszig, Erwin; Kreyszig, Herbert; Norminton, Edward J. (2011). Advanced Engineering Mathematics (10 ed.). John Wiley and Sons. pp. 453–456. ISBN 978-0-470-45836-5.
  6. ^ Benford, Frank A. (May 2007). "Notes on Vector Calculus" (PDF). Course materials for Math 105: Multivariable Calculus. Prof. Steven Miller's webpage, Williams College. Retrieved 14 March 2022.
  7. ^ a b c Purcell, Edward M.; David J. Morin (2013). Electricity and Magnetism. Cambridge Univ. Press. pp. 56–58. ISBN 978-1-107-01402-2.
  8. ^ Alt, Hans Wilhelm (2016). "Linear Functional Analysis". Universitext. London: Springer London. pp. 259–261, 270–272. doi:10.1007/978-1-4471-7280-2. ISBN 978-1-4471-7279-6. ISSN 0172-5939.
  9. ^ Taylor, Michael E. (2011). "Partial Differential Equations I". Applied Mathematical Sciences. New York, NY: Springer New York. pp. 178–179. doi:10.1007/978-1-4419-7055-8. ISBN 978-1-4419-7054-1. ISSN 0066-5452.
  10. ^ M. R. Spiegel; S. Lipschutz; D. Spellman (2009). Vector Analysis. Schaum's Outlines (2nd ed.). USA: McGraw Hill. ISBN 978-0-07-161545-7.
  11. ^ a b MathWorld
  12. ^ a b C.B. Parker (1994). McGraw Hill Encyclopaedia of Physics (2nd ed.). McGraw Hill. ISBN 978-0-07-051400-3.
  13. ^ a b Katz, Victor (2009). "Chapter 22: Vector Analysis". A History of Mathematics: An Introduction. Addison-Wesley. pp. 808–9. ISBN 978-0-321-38700-4.
  14. ^ In his 1762 paper on sound, Lagrange treats a special case of the divergence theorem: Lagrange (1762) "Nouvelles recherches sur la nature et la propagation du son" (New researches on the nature and propagation of sound), Miscellanea Taurinensia (also known as: Mélanges de Turin ), 2: 11 – 172. This article is reprinted as: "Nouvelles recherches sur la nature et la propagation du son" in: J.A. Serret, ed., Oeuvres de Lagrange, (Paris, France: Gauthier-Villars, 1867), vol. 1, pages 151–316; on pages 263–265, Lagrange transforms triple integrals into double integrals using integration by parts.
  15. ^ C. F. Gauss (1813) "Theoria attractionis corporum sphaeroidicorum ellipticorum homogeneorum methodo nova tractata," Commentationes societatis regiae scientiarium Gottingensis recentiores, 2: 355–378; Gauss considered a special case of the theorem; see the 4th, 5th, and 6th pages of his article.
  16. ^ a b c Katz, Victor (May 1979). "A History of Stokes' Theorem". Mathematics Magazine. 52 (3): 146–156. doi:10.1080/0025570X.1979.11976770. JSTOR 2690275.
  17. ^ Mikhail Ostragradsky presented his proof of the divergence theorem to the Paris Academy in 1826; however, his work was not published by the Academy. He returned to St. Petersburg, Russia, where in 1828–1829 he read the work that he'd done in France, to the St. Petersburg Academy, which published his work in abbreviated form in 1831.
    • His proof of the divergence theorem – "Démonstration d'un théorème du calcul intégral" (Proof of a theorem in integral calculus) – which he had read to the Paris Academy on February 13, 1826, was translated, in 1965, into Russian together with another article by him. See: Юшкевич А.П. (Yushkevich A.P.) and Антропова В.И. (Antropov V.I.) (1965) "Неопубликованные работы М.В. Остроградского" (Unpublished works of MV Ostrogradskii), Историко-математические исследования (Istoriko-Matematicheskie Issledovaniya / Historical-Mathematical Studies), 16: 49–96; see the section titled: "Остроградский М.В. Доказательство одной теоремы интегрального исчисления" (Ostrogradskii M. V. Dokazatelstvo odnoy teoremy integralnogo ischislenia / Ostragradsky M.V. Proof of a theorem in integral calculus).
    • M. Ostrogradsky (presented: November 5, 1828; published: 1831) "Première note sur la théorie de la chaleur" (First note on the theory of heat) Mémoires de l'Académie impériale des sciences de St. Pétersbourg, series 6, 1: 129–133; for an abbreviated version of his proof of the divergence theorem, see pages 130–131.
    • Victor J. Katz (May1979) "The history of Stokes' theorem," April 2, 2015, at the Wayback Machine Mathematics Magazine, 52(3): 146–156; for Ostragradsky's proof of the divergence theorem, see pages 147–148.
  18. ^ George Green, An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism (Nottingham, England: T. Wheelhouse, 1838). A form of the "divergence theorem" appears on pages 10–12.
  19. ^ Other early investigators who used some form of the divergence theorem include:
    • Poisson (presented: February 2, 1824; published: 1826) "Mémoire sur la théorie du magnétisme" (Memoir on the theory of magnetism), Mémoires de l'Académie des sciences de l'Institut de France, 5: 247–338; on pages 294–296, Poisson transforms a volume integral (which is used to evaluate a quantity Q) into a surface integral. To make this transformation, Poisson follows the same procedure that is used to prove the divergence theorem.
    • Frédéric Sarrus (1828) "Mémoire sur les oscillations des corps flottans" (Memoir on the oscillations of floating bodies), Annales de mathématiques pures et appliquées (Nismes), 19: 185–211.
  20. ^ K.F. Riley; M.P. Hobson; S.J. Bence (2010). Mathematical methods for physics and engineering. Cambridge University Press. ISBN 978-0-521-86153-3.
  21. ^ see for example:
    J.A. Wheeler; C. Misner; K.S. Thorne (1973). Gravitation. W.H. Freeman & Co. pp. 85–86, §3.5. ISBN 978-0-7167-0344-0., and
    R. Penrose (2007). The Road to Reality. Vintage books. ISBN 978-0-679-77631-4.

External links edit

divergence, theorem, gauss, theorem, redirects, here, theorem, concerning, electric, field, gauss, ostrogradsky, theorem, redirects, here, theorem, concerning, linear, instability, hamiltonian, associated, with, lagrangian, dependent, higher, time, derivatives. Gauss s theorem redirects here For the theorem concerning the electric field see Gauss s law Ostrogradsky s theorem redirects here For the theorem concerning the linear instability of the Hamiltonian associated with a Lagrangian dependent on higher time derivatives than the first see Ostrogradsky instability In vector calculus the divergence theorem also known as Gauss s theorem or Ostrogradsky s theorem 1 is a theorem relating 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 enclosed by the surface Intuitively it states that the sum of all sources of the field in a region with sinks regarded as negative sources gives the net flux out of the region The divergence theorem is an important result for the mathematics of physics and engineering particularly in electrostatics and fluid dynamics In these fields it is usually applied in three dimensions However it generalizes to any number of dimensions In one dimension it is equivalent to the fundamental theorem of calculus In two dimensions it is equivalent to Green s theorem Contents 1 Explanation using liquid flow 2 Mathematical statement 3 Informal derivation 4 Proofs 4 1 For bounded open subsets of Euclidean space 4 2 For compact Riemannian manifolds with boundary 5 Corollaries 6 Example 7 Applications 7 1 Differential and integral forms of physical laws 7 1 1 Continuity equations 7 2 Inverse square laws 8 History 9 Worked examples 9 1 Example 1 9 2 Example 2 10 Generalizations 10 1 Multiple dimensions 10 2 Tensor fields 11 See also 12 References 13 External linksExplanation using liquid flow editVector fields 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 at any moment 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 at any time 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 with the flow through a sink given a negative sign is equal to the divergence of the velocity field at the pipe mouth so adding up integrating the divergence of the liquid throughout the volume enclosed by S equals the volume rate of flux through S This is the divergence theorem 2 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 3 Mathematical statement edit nbsp A region V bounded by the surface S V displaystyle S partial V nbsp with the surface normal n Suppose V is a subset of R n displaystyle mathbb R n nbsp in the case of n 3 V represents a volume in three dimensional space which is compact and has a piecewise smooth boundary S also indicated with V S displaystyle partial V S nbsp If F is a continuously differentiable vector field defined on a neighborhood of V then 4 5 V F d V displaystyle iiint V left mathbf nabla cdot mathbf F right mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp F n d S displaystyle mathbf F cdot mathbf hat n mathrm d S nbsp The left side is a volume integral over the volume V the right side is the surface integral over the boundary of the volume V The closed manifold V displaystyle partial V nbsp is oriented by outward pointing normals and n displaystyle mathbf hat n nbsp is the outward pointing unit normal at each point on the boundary V displaystyle partial V nbsp d S displaystyle mathrm d mathbf S nbsp may be used as a shorthand for n d S displaystyle mathbf n mathrm d S nbsp In terms of the intuitive description above the left hand side of the equation represents the total of the sources in the volume V and the right hand side represents the total flow across the boundary S Informal derivation editThe divergence theorem follows from the fact that if a volume V 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 6 7 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 nbsp A volume divided into two subvolumes At right the two subvolumes are separated to show the flux out of the different surfaces See the diagram A closed bounded volume V is divided into two volumes V1 and V2 by a surface S3 green The flux F Vi out of each component region Vi is equal to the sum of the flux through its two faces so the sum of the flux out of the two parts is F V 1 F V 2 F 1 F 31 F 2 F 32 displaystyle Phi V text 1 Phi V text 2 Phi text 1 Phi text 31 Phi text 2 Phi text 32 nbsp where F1 and F2 are the flux out of surfaces S1 and S2 F31 is the flux through S3 out of volume 1 and F32 is the flux through S3 out of volume 2 The point is that surface S3 is part of the surface of both volumes The outward direction of the normal vector n displaystyle mathbf hat n nbsp is opposite for each volume so the flux out of one through S3 is equal to the negative of the flux out of the other F 31 S 3 F n d S S 3 F n d S F 32 displaystyle Phi text 31 iint S 3 mathbf F cdot mathbf hat n mathrm d S iint S 3 mathbf F cdot mathbf hat n mathrm d S Phi text 32 nbsp so these two fluxes cancel in the sum Therefore F V 1 F V 2 F 1 F 2 displaystyle Phi V text 1 Phi V text 2 Phi text 1 Phi text 2 nbsp Since the union of surfaces S1 and S2 is S F V 1 F V 2 F V displaystyle Phi V text 1 Phi V text 2 Phi V nbsp nbsp The volume can be divided into any number of subvolumes and the flux out of V is equal to the sum of the flux out of each subvolume because the flux through the green surfaces cancels out in the sum In b the volumes are shown separated slightly illustrating that each green partition is part of the boundary of two adjacent volumes This principle applies to a volume divided into any number of parts as shown in the diagram 7 Since the integral over each internal partition green surfaces 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 grey Since the external surfaces of all the component volumes equal the original surface F V V i V F V i displaystyle Phi V sum V text i subset V Phi V text i nbsp nbsp As the volume is subdivided into smaller parts the ratio of the flux F V i displaystyle Phi V text i nbsp out of each volume to the volume V i displaystyle V text i nbsp approaches div F displaystyle operatorname div mathbf F nbsp The flux F out of each volume is the surface integral of the vector field F x over the surface S V F n d S V i V S V i F n d S displaystyle iint S V mathbf F cdot mathbf hat n mathrm d S sum V text i subset V iint S V text i mathbf F cdot mathbf hat n mathrm d S nbsp 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 S Vi approaches zero However from the definition of divergence the ratio of flux to volume F V i V i 1 V i S V i F n d S displaystyle frac Phi V text i V text i frac 1 V text i iint S V text i mathbf F cdot mathbf hat n mathrm d S nbsp the part in parentheses below does not in general vanish but approaches the divergence div F as the volume approaches zero 7 S V F n d S V i V 1 V i S V i F n d S V i displaystyle iint S V mathbf F cdot mathbf hat n mathrm d S sum V text i subset V left frac 1 V text i iint S V text i mathbf F cdot mathbf hat n mathrm d S right V text i nbsp As long as the vector field F x has continuous derivatives the sum above holds even in the limit when the volume is divided into infinitely small increments S V F n d S lim V i 0 V i V 1 V i S V i F n d S V i displaystyle iint S V mathbf F cdot mathbf hat n mathrm d S lim V text i to 0 sum V text i subset V left frac 1 V text i iint S V text i mathbf F cdot mathbf hat n mathrm d S right V text i nbsp As V i displaystyle V text i nbsp approaches zero volume it becomes the infinitesimal dV the part in parentheses becomes the divergence and the sum becomes a volume integral over V S V F n d S V div F d V displaystyle iint S V mathbf F cdot mathbf hat n mathrm d S iiint V operatorname div mathbf F mathrm d V nbsp Since this derivation is coordinate free it shows that the divergence does not depend on the coordinates used Proofs editFor bounded open subsets of Euclidean space edit We are going to prove the following Theorem Let W R n displaystyle Omega subset mathbb R n nbsp be open and bounded with C 1 displaystyle C 1 nbsp boundary If u displaystyle u nbsp is C 1 displaystyle C 1 nbsp on an open neighborhood O displaystyle O nbsp of W displaystyle overline Omega nbsp that is u C 1 O displaystyle u in C 1 O nbsp then for each i 1 n displaystyle i in 1 dots n nbsp W u x i d V W u n i d S displaystyle int Omega u x i dV int partial Omega u nu i dS nbsp where n W R n displaystyle nu partial Omega to mathbb R n nbsp is the outward pointing unit normal vector to W displaystyle partial Omega nbsp Equivalently W u d V W u n d S displaystyle int Omega nabla u dV int partial Omega u nu dS nbsp Proof of Theorem 8 1 The first step is to reduce to the case where u C c 1 R n displaystyle u in C c 1 mathbb R n nbsp Pick ϕ C c O displaystyle phi in C c infty O nbsp such that ϕ 1 displaystyle phi 1 nbsp on W displaystyle overline Omega nbsp Note that ϕ u C c 1 O C c 1 R n displaystyle phi u in C c 1 O subset C c 1 mathbb R n nbsp and ϕ u u displaystyle phi u u nbsp on W displaystyle overline Omega nbsp Hence it suffices to prove the theorem for ϕ u displaystyle phi u nbsp Hence we may assume that u C c 1 R n displaystyle u in C c 1 mathbb R n nbsp 2 Let x 0 W displaystyle x 0 in partial Omega nbsp be arbitrary The assumption that W displaystyle overline Omega nbsp has C 1 displaystyle C 1 nbsp boundary means that there is an open neighborhood U displaystyle U nbsp of x 0 displaystyle x 0 nbsp in R n displaystyle mathbb R n nbsp such that W U displaystyle partial Omega cap U nbsp is the graph of a C 1 displaystyle C 1 nbsp function with W U displaystyle Omega cap U nbsp lying on one side of this graph More precisely this means that after a translation and rotation of W displaystyle Omega nbsp there are r gt 0 displaystyle r gt 0 nbsp and h gt 0 displaystyle h gt 0 nbsp and a C 1 displaystyle C 1 nbsp function g R n 1 R displaystyle g mathbb R n 1 to mathbb R nbsp such that with the notationx x 1 x n 1 displaystyle x x 1 dots x n 1 nbsp it holds that U x R n x lt r and x n g x lt h displaystyle U x in mathbb R n x lt r text and x n g x lt h nbsp and for x U displaystyle x in U nbsp x n g x x W h lt x n g x lt 0 x W 0 lt x n g x lt h x W displaystyle begin aligned x n g x amp implies x in partial Omega h lt x n g x lt 0 amp implies x in Omega 0 lt x n g x lt h amp implies x notin Omega end aligned nbsp Since W displaystyle partial Omega nbsp is compact we can cover W displaystyle partial Omega nbsp with finitely many neighborhoods U 1 U N displaystyle U 1 dots U N nbsp of the above form Note that W U 1 U N displaystyle Omega U 1 dots U N nbsp is an open cover of W W W displaystyle overline Omega Omega cup partial Omega nbsp By using a C displaystyle C infty nbsp partition of unity subordinate to this cover it suffices to prove the theorem in the case where either u displaystyle u nbsp has compact support in W displaystyle Omega nbsp or u displaystyle u nbsp has compact support in some U j displaystyle U j nbsp If u displaystyle u nbsp has compact support in W displaystyle Omega nbsp then for all i 1 n displaystyle i in 1 dots n nbsp W u x i d V R n u x i d V R n 1 u x i x d x i d x 0 displaystyle int Omega u x i dV int mathbb R n u x i dV int mathbb R n 1 int infty infty u x i x dx i dx 0 nbsp by the fundamental theorem of calculus and W u n i d S 0 displaystyle int partial Omega u nu i dS 0 nbsp since u displaystyle u nbsp vanishes on a neighborhood of W displaystyle partial Omega nbsp Thus the theorem holds for u displaystyle u nbsp with compact support in W displaystyle Omega nbsp Thus we have reduced to the case where u displaystyle u nbsp has compact support in some U j displaystyle U j nbsp 3 So assume u displaystyle u nbsp has compact support in some U j displaystyle U j nbsp The last step now is to show that the theorem is true by direct computation Change notation to U U j displaystyle U U j nbsp and bring in the notation from 2 used to describe U displaystyle U nbsp Note that this means that we have rotated and translated W displaystyle Omega nbsp This is a valid reduction since the theorem is invariant under rotations and translations of coordinates Since u x 0 displaystyle u x 0 nbsp for x r displaystyle x geq r nbsp and for x n g x h displaystyle x n g x geq h nbsp we have for each i 1 n displaystyle i in 1 dots n nbsp that W u x i d V x lt r g x h g x u x i x x n d x n d x R n 1 g x u x i x x n d x n d x displaystyle begin aligned int Omega u x i dV amp int x lt r int g x h g x u x i x x n dx n dx amp int mathbb R n 1 int infty g x u x i x x n dx n dx end aligned nbsp For i n displaystyle i n nbsp we have by the fundamental theorem of calculus that R n 1 g x u x n x x n d x n d x R n 1 u x g x d x displaystyle int mathbb R n 1 int infty g x u x n x x n dx n dx int mathbb R n 1 u x g x dx nbsp Now fix i 1 n 1 displaystyle i in 1 dots n 1 nbsp Note that R n 1 g x u x i x x n d x n d x R n 1 0 u x i x g x s d s d x displaystyle int mathbb R n 1 int infty g x u x i x x n dx n dx int mathbb R n 1 int infty 0 u x i x g x s ds dx nbsp Define v R n R displaystyle v mathbb R n to mathbb R nbsp by v x s u x g x s displaystyle v x s u x g x s nbsp By the chain rule v x i x s u x i x g x s u x n x g x s g x i x displaystyle v x i x s u x i x g x s u x n x g x s g x i x nbsp But since v displaystyle v nbsp has compact support we can integrate out d x i displaystyle dx i nbsp first to deduce that R n 1 0 v x i x s d s d x 0 displaystyle int mathbb R n 1 int infty 0 v x i x s ds dx 0 nbsp Thus R n 1 0 u x i x g x s d s d x R n 1 0 u x n x g x s g x i x d s d x R n 1 u x g x g x i x d x displaystyle begin aligned int mathbb R n 1 int infty 0 u x i x g x s ds dx amp int mathbb R n 1 int infty 0 u x n x g x s g x i x ds dx amp int mathbb R n 1 u x g x g x i x dx end aligned nbsp In summary with u u x 1 u x n displaystyle nabla u u x 1 dots u x n nbsp we have W u d V R n 1 g x u d V R n 1 u x g x g x 1 d x displaystyle int Omega nabla u dV int mathbb R n 1 int infty g x nabla u dV int mathbb R n 1 u x g x nabla g x 1 dx nbsp Recall that the outward unit normal to the graph G displaystyle Gamma nbsp of g displaystyle g nbsp at a point x g x G displaystyle x g x in Gamma nbsp is n x g x 1 1 g x 2 g x 1 displaystyle nu x g x frac 1 sqrt 1 nabla g x 2 nabla g x 1 nbsp and that the surface element d S displaystyle dS nbsp is given by d S 1 g x 2 d x displaystyle dS sqrt 1 nabla g x 2 dx nbsp Thus W u d V W u n d S displaystyle int Omega nabla u dV int partial Omega u nu dS nbsp This completes the proof For compact Riemannian manifolds with boundary edit We are going to prove the following Theorem Let W displaystyle overline Omega nbsp be a C 2 displaystyle C 2 nbsp compact manifold with boundary with C 1 displaystyle C 1 nbsp metric tensor g displaystyle g nbsp Let W displaystyle Omega nbsp denote the manifold interior of W displaystyle overline Omega nbsp and let W displaystyle partial Omega nbsp denote the manifold boundary of W displaystyle overline Omega nbsp Let displaystyle cdot cdot nbsp denote L 2 W displaystyle L 2 overline Omega nbsp inner products of functions and displaystyle langle cdot cdot rangle nbsp denote inner products of vectors Suppose u C 1 W R displaystyle u in C 1 overline Omega mathbb R nbsp and X displaystyle X nbsp is a C 1 displaystyle C 1 nbsp vector field on W displaystyle overline Omega nbsp Then grad u X u div X W u X N d S displaystyle operatorname grad u X u operatorname div X int partial Omega u langle X N rangle dS nbsp where N displaystyle N nbsp is the outward pointing unit normal vector to W displaystyle partial Omega nbsp Proof of Theorem 9 We use the Einstein summation convention By using a partition of unity we may assume that u displaystyle u nbsp and X displaystyle X nbsp have compact support in a coordinate patch O W displaystyle O subset overline Omega nbsp First consider the case where the patch is disjoint from W displaystyle partial Omega nbsp Then O displaystyle O nbsp is identified with an open subset of R n displaystyle mathbb R n nbsp and integration by parts produces no boundary terms grad u X O grad u X g d x O j u X j g d x O u j g X j d x O u 1 g j g X j g d x u 1 g j g X j u div X displaystyle begin aligned operatorname grad u X amp int O langle operatorname grad u X rangle sqrt g dx amp int O partial j uX j sqrt g dx amp int O u partial j sqrt g X j dx amp int O u frac 1 sqrt g partial j sqrt g X j sqrt g dx amp u frac 1 sqrt g partial j sqrt g X j amp u operatorname div X end aligned nbsp In the last equality we used the Voss Weyl coordinate formula for the divergence although the preceding identity could be used to define div displaystyle operatorname div nbsp as the formal adjoint of grad displaystyle operatorname grad nbsp Now suppose O displaystyle O nbsp intersects W displaystyle partial Omega nbsp Then O displaystyle O nbsp is identified with an open set in R n x R n x n 0 displaystyle mathbb R n x in mathbb R n x n geq 0 nbsp We zero extend u displaystyle u nbsp and X displaystyle X nbsp to R n displaystyle mathbb R n nbsp and perform integration by parts to obtain grad u X O grad u X g d x R n j u X j g d x u div X R n 1 u x 0 X n x 0 g x 0 d x displaystyle begin aligned operatorname grad u X amp int O langle operatorname grad u X rangle sqrt g dx amp int mathbb R n partial j uX j sqrt g dx amp u operatorname div X int mathbb R n 1 u x 0 X n x 0 sqrt g x 0 dx end aligned nbsp where d x d x 1 d x n 1 displaystyle dx dx 1 dots dx n 1 nbsp By a variant of the straightening theorem for vector fields we may choose O displaystyle O nbsp so that x n displaystyle frac partial partial x n nbsp is the inward unit normal N displaystyle N nbsp at W displaystyle partial Omega nbsp In this case g x 0 d x g W x d x d S displaystyle sqrt g x 0 dx sqrt g partial Omega x dx dS nbsp is the volume element on W displaystyle partial Omega nbsp and the above formula reads grad u X u div X W u X N d S displaystyle operatorname grad u X u operatorname div X int partial Omega u langle X N rangle dS nbsp This completes the proof Corollaries editBy replacing F in the divergence theorem with specific forms other useful identities can be derived cf vector identities 10 With F F g displaystyle mathbf F rightarrow mathbf F g nbsp for a scalar function g and a vector field F V F g g F d V displaystyle iiint V left mathbf F cdot left nabla g right g left nabla cdot mathbf F right right mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp g F n d S displaystyle g mathbf F cdot mathbf n mathrm d S nbsp dd A special case of this is F f displaystyle mathbf F nabla f nbsp in which case the theorem is the basis for Green s identities With F F G displaystyle mathbf F rightarrow mathbf F times mathbf G nbsp for two vector fields F and G where displaystyle times nbsp denotes a cross product V F G d V V G F F G d V displaystyle iiint V nabla cdot left mathbf F times mathbf G right mathrm d V iiint V left mathbf G cdot left nabla times mathbf F right mathbf F cdot left nabla times mathbf G right right mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp F G n d S displaystyle mathbf F times mathbf G cdot mathbf n mathrm d S nbsp dd With F F G displaystyle mathbf F rightarrow mathbf F cdot mathbf G nbsp for two vector fields F and G where displaystyle cdot nbsp denotes a dot product V F G d V V G F F G d V displaystyle iiint V nabla left mathbf F cdot mathbf G right mathrm d V iiint V left left nabla mathbf G right cdot mathbf F left nabla mathbf F right cdot mathbf G right mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp F G n d S displaystyle mathbf F cdot mathbf G mathbf n mathrm d S nbsp dd With F f c displaystyle mathbf F rightarrow f mathbf c nbsp for a scalar function f and vector field c 11 V c f d V displaystyle iiint V mathbf c cdot nabla f mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp c f n d S V f c d V displaystyle mathbf c f cdot mathbf n mathrm d S iiint V f nabla cdot mathbf c mathrm d V nbsp dd The last term on the right vanishes for constant c displaystyle mathbf c nbsp or any divergence free solenoidal vector field e g Incompressible flows without sources or sinks such as phase change or chemical reactions etc In particular taking c displaystyle mathbf c nbsp to be constant V f d V displaystyle iiint V nabla f mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp f n d S displaystyle f mathbf n mathrm d S nbsp dd With F c F displaystyle mathbf F rightarrow mathbf c times mathbf F nbsp for vector field F and constant vector c 11 V c F d V displaystyle iiint V mathbf c cdot nabla times mathbf F mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp F c n d S displaystyle mathbf F times mathbf c cdot mathbf n mathrm d S nbsp dd By reordering the triple product on the right hand side and taking out the constant vector of the integral V F d V c displaystyle iiint V nabla times mathbf F mathrm d V cdot mathbf c nbsp nbsp S displaystyle scriptstyle S nbsp d S F c displaystyle mathrm d mathbf S times mathbf F cdot mathbf c nbsp dd Hence V F d V displaystyle iiint V nabla times mathbf F mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp n F d S displaystyle mathbf n times mathbf F mathrm d S nbsp dd Example edit nbsp The vector field corresponding to the example shown Vectors may point into or out of the sphere nbsp The divergence theorem can be used to calculate a flux through a closed surface that fully encloses a volume like any of the surfaces on the left It can not directly be used to calculate the flux through surfaces with boundaries like those on the right Surfaces are blue boundaries are red Suppose we wish to evaluate nbsp S displaystyle scriptstyle S nbsp F n d S displaystyle mathbf F cdot mathbf n mathrm d S nbsp where S is the unit sphere defined by S x y z R 3 x 2 y 2 z 2 1 displaystyle S left x y z in mathbb R 3 x 2 y 2 z 2 1 right nbsp and F is the vector field F 2 x i y 2 j z 2 k displaystyle mathbf F 2x mathbf i y 2 mathbf j z 2 mathbf k nbsp 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 W F d V 2 W 1 y z d V 2 W d V 2 W y d V 2 W z d V displaystyle iiint W nabla cdot mathbf F mathrm d V 2 iiint W 1 y z mathrm d V 2 iiint W mathrm d V 2 iiint W y mathrm d V 2 iiint W z mathrm d V nbsp where W is the unit ball W x y z R 3 x 2 y 2 z 2 1 displaystyle W left x y z in mathbb R 3 x 2 y 2 z 2 leq 1 right nbsp Since the function y is positive in one hemisphere of W and negative in the other in an equal and opposite way its total integral over W is zero The same is true for z W y d V W z d V 0 displaystyle iiint W y mathrm d V iiint W z mathrm d V 0 nbsp Therefore nbsp S displaystyle scriptstyle S nbsp F n d S 2 W d V 8 p 3 displaystyle mathbf F cdot mathbf n mathrm d S 2 iiint W dV frac 8 pi 3 nbsp because the unit ball W has volume 4p 3 Applications editDifferential and integral forms of physical laws edit As a result of the divergence theorem a host of physical laws can be written in both a differential form where one quantity is the divergence of another and an integral form where the flux of one quantity through a closed surface is equal to another quantity Three examples are Gauss s law in electrostatics Gauss s law for magnetism and Gauss s law for gravity Continuity equations edit Main article continuity equation Continuity equations 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 in terms of a divergence and an integral form in terms of a flux 12 Inverse square laws edit Any inverse square law can instead be written in a Gauss s law type form with a differential and integral form as described above Two examples are Gauss s law in electrostatics 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 12 History editJoseph Louis Lagrange introduced the notion of surface integrals in 1760 and again in more general terms in 1811 in the second edition of his Mecanique Analytique Lagrange employed surface integrals in his work on fluid mechanics 13 He discovered the divergence theorem in 1762 14 Carl Friedrich Gauss was also using surface integrals while working on the gravitational attraction of an elliptical spheroid in 1813 when he proved special cases of the divergence theorem 15 13 He proved additional special cases in 1833 and 1839 16 But it was Mikhail Ostrogradsky who gave the first proof of the general theorem in 1826 as part of his investigation of heat flow 17 Special cases were proven by George Green in 1828 in An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism 18 16 Simeon Denis Poisson in 1824 in a paper on elasticity and Frederic Sarrus in 1828 in his work on floating bodies 19 16 Worked examples editExample 1 edit To verify the planar variant of the divergence theorem for a region R displaystyle R nbsp R x y R 2 x 2 y 2 1 displaystyle R left x y in mathbb R 2 x 2 y 2 leq 1 right nbsp and the vector field F x y 2 y i 5 x j displaystyle mathbf F x y 2y mathbf i 5x mathbf j nbsp The boundary of R displaystyle R nbsp is the unit circle C displaystyle C nbsp that can be represented parametrically by x cos s y sin s displaystyle x cos s quad y sin s nbsp such that 0 s 2 p displaystyle 0 leq s leq 2 pi nbsp where s displaystyle s nbsp units is the length arc from the point s 0 displaystyle s 0 nbsp to the point P displaystyle P nbsp on C displaystyle C nbsp Then a vector equation of C displaystyle C nbsp is C s cos s i sin s j displaystyle C s cos s mathbf i sin s mathbf j nbsp At a point P displaystyle P nbsp on C displaystyle C nbsp P cos s sin s F 2 sin s i 5 cos s j displaystyle P cos s sin s Rightarrow mathbf F 2 sin s mathbf i 5 cos s mathbf j nbsp Therefore C F n d s 0 2 p 2 sin s i 5 cos s j cos s i sin s j d s 0 2 p 2 sin s cos s 5 sin s cos s d s 7 0 2 p sin s cos s d s 0 displaystyle begin aligned oint C mathbf F cdot mathbf n mathrm d s amp int 0 2 pi 2 sin s mathbf i 5 cos s mathbf j cdot cos s mathbf i sin s mathbf j mathrm d s amp int 0 2 pi 2 sin s cos s 5 sin s cos s mathrm d s amp 7 int 0 2 pi sin s cos s mathrm d s amp 0 end aligned nbsp Because M 2 y displaystyle M 2y nbsp we can evaluate M x 0 displaystyle frac partial M partial x 0 nbsp and because N 5 x displaystyle N 5x nbsp N y 0 displaystyle frac partial N partial y 0 nbsp Thus R F d A R M x N y d A 0 displaystyle iint R mathbf nabla cdot mathbf F mathrm d A iint R left frac partial M partial x frac partial N partial y right mathrm d A 0 nbsp Example 2 edit Let s say we wanted to evaluate the flux of the following vector field defined by F 2 x 2 i 2 y 2 j 2 z 2 k displaystyle mathbf F 2x 2 textbf i 2y 2 textbf j 2z 2 textbf k nbsp bounded by the following inequalities 0 x 3 2 y 2 0 z 2 p displaystyle left 0 leq x leq 3 right left 2 leq y leq 2 right left 0 leq z leq 2 pi right nbsp By the divergence theorem V F d V displaystyle iiint V left mathbf nabla cdot mathbf F right mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp F n d S displaystyle mathbf F cdot mathbf n mathrm d S nbsp We now need to determine the divergence of F displaystyle textbf F nbsp If F displaystyle mathbf F nbsp is a three dimensional vector field then the divergence of F displaystyle textbf F nbsp is given by F x i y j z k F textstyle nabla cdot textbf F left frac partial partial x textbf i frac partial partial y textbf j frac partial partial z textbf k right cdot textbf F nbsp Thus we can set up the following flux integral I displaystyle I nbsp nbsp S displaystyle scriptstyle S nbsp F n d S displaystyle mathbf F cdot mathbf n mathrm d S nbsp as follows I V F d V V F x x F y y F z z d V V 4 x 4 y 4 z d V 0 3 2 2 0 2 p 4 x 4 y 4 z d V displaystyle begin aligned I amp iiint V nabla cdot mathbf F mathrm d V 6pt amp iiint V left frac partial mathbf F x partial x frac partial mathbf F y partial y frac partial mathbf F z partial z right mathrm d V 6pt amp iiint V 4x 4y 4z mathrm d V 6pt amp int 0 3 int 2 2 int 0 2 pi 4x 4y 4z mathrm d V end aligned nbsp Now that we have set up the integral we can evaluate it 0 3 2 2 0 2 p 4 x 4 y 4 z d V 2 2 0 2 p 12 y 12 z 18 d y d z 0 2 p 24 2 z 3 d z 48 p 2 p 3 displaystyle begin aligned int 0 3 int 2 2 int 0 2 pi 4x 4y 4z mathrm d V amp int 2 2 int 0 2 pi 12y 12z 18 mathrm d y mathrm d z 6pt amp int 0 2 pi 24 2z 3 mathrm d z 6pt amp 48 pi 2 pi 3 end aligned nbsp Generalizations editMultiple dimensions edit One can use the generalised Stokes theorem to equate the n dimensional volume integral of the divergence of a vector field F over a region U to the n 1 dimensional surface integral of F over the boundary of U U n F d V U n 1 F n d S displaystyle underbrace int cdots int U n nabla cdot mathbf F mathrm d V underbrace oint cdots oint partial U n 1 mathbf F cdot mathbf n mathrm d S nbsp This equation is also known as the divergence theorem When n 2 this is equivalent to Green s theorem When n 1 it reduces to the fundamental theorem of calculus part 2 Tensor fields edit Main article Tensor field Writing the theorem in Einstein notation V F i x i d V displaystyle iiint V dfrac partial mathbf F i partial x i mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp F i n i d S displaystyle mathbf F i n i mathrm d S nbsp suggestively replacing the vector field F with a rank n tensor field T this can be generalized to 20 V T i 1 i 2 i q i n x i q d V displaystyle iiint V dfrac partial T i 1 i 2 cdots i q cdots i n partial x i q mathrm d V nbsp nbsp S displaystyle scriptstyle S nbsp T i 1 i 2 i q i n n i q d S displaystyle T i 1 i 2 cdots i q cdots i n n i q mathrm d S nbsp 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 or lower dimensions for example to 4d spacetime in general relativity 21 See also editKelvin Stokes theoremReferences edit Katz Victor J 1979 The history of Stokes s theorem Mathematics Magazine 52 3 146 156 doi 10 2307 2690275 JSTOR 2690275 reprinted in Anderson Marlow 2009 Who Gave You the Epsilon And Other Tales of Mathematical History Mathematical Association of America pp 78 79 ISBN 978 0 88385 569 0 R G Lerner G L Trigg 1994 Encyclopaedia of Physics 2nd ed VHC ISBN 978 3 527 26954 9 Byron Frederick Fuller Robert 1992 Mathematics of Classical and Quantum Physics Dover Publications p 22 ISBN 978 0 486 67164 2 Wiley C Ray Jr Advanced Engineering Mathematics 3rd Ed McGraw Hill pp 372 373 Kreyszig Erwin Kreyszig Herbert Norminton Edward J 2011 Advanced Engineering Mathematics 10 ed John Wiley and Sons pp 453 456 ISBN 978 0 470 45836 5 Benford Frank A May 2007 Notes on Vector Calculus PDF Course materials for Math 105 Multivariable Calculus Prof Steven Miller s webpage Williams College Retrieved 14 March 2022 a b c Purcell Edward M David J Morin 2013 Electricity and Magnetism Cambridge Univ Press pp 56 58 ISBN 978 1 107 01402 2 Alt Hans Wilhelm 2016 Linear Functional Analysis Universitext London Springer London pp 259 261 270 272 doi 10 1007 978 1 4471 7280 2 ISBN 978 1 4471 7279 6 ISSN 0172 5939 Taylor Michael E 2011 Partial Differential Equations I Applied Mathematical Sciences New York NY Springer New York pp 178 179 doi 10 1007 978 1 4419 7055 8 ISBN 978 1 4419 7054 1 ISSN 0066 5452 M R Spiegel S Lipschutz D Spellman 2009 Vector Analysis Schaum s Outlines 2nd ed USA McGraw Hill ISBN 978 0 07 161545 7 a b MathWorld a b C B Parker 1994 McGraw Hill Encyclopaedia of Physics 2nd ed McGraw Hill ISBN 978 0 07 051400 3 a b Katz Victor 2009 Chapter 22 Vector Analysis A History of Mathematics An Introduction Addison Wesley pp 808 9 ISBN 978 0 321 38700 4 In his 1762 paper on sound Lagrange treats a special case of the divergence theorem Lagrange 1762 Nouvelles recherches sur la nature et la propagation du son New researches on the nature and propagation of sound Miscellanea Taurinensia also known as Melanges de Turin 2 11 172 This article is reprinted as Nouvelles recherches sur la nature et la propagation du son in J A Serret ed Oeuvres de Lagrange Paris France Gauthier Villars 1867 vol 1 pages 151 316 on pages 263 265 Lagrange transforms triple integrals into double integrals using integration by parts C F Gauss 1813 Theoria attractionis corporum sphaeroidicorum ellipticorum homogeneorum methodo nova tractata Commentationes societatis regiae scientiarium Gottingensis recentiores 2 355 378 Gauss considered a special case of the theorem see the 4th 5th and 6th pages of his article a b c Katz Victor May 1979 A History of Stokes Theorem Mathematics Magazine 52 3 146 156 doi 10 1080 0025570X 1979 11976770 JSTOR 2690275 Mikhail Ostragradsky presented his proof of the divergence theorem to the Paris Academy in 1826 however his work was not published by the Academy He returned to St Petersburg Russia where in 1828 1829 he read the work that he d done in France to the St Petersburg Academy which published his work in abbreviated form in 1831 His proof of the divergence theorem Demonstration d un theoreme du calcul integral Proof of a theorem in integral calculus which he had read to the Paris Academy on February 13 1826 was translated in 1965 into Russian together with another article by him See Yushkevich A P Yushkevich A P and Antropova V I Antropov V I 1965 Neopublikovannye raboty M V Ostrogradskogo Unpublished works of MV Ostrogradskii Istoriko matematicheskie issledovaniya Istoriko Matematicheskie Issledovaniya Historical Mathematical Studies 16 49 96 see the section titled Ostrogradskij M V Dokazatelstvo odnoj teoremy integralnogo ischisleniya Ostrogradskii M V Dokazatelstvo odnoy teoremy integralnogo ischislenia Ostragradsky M V Proof of a theorem in integral calculus M Ostrogradsky presented November 5 1828 published 1831 Premiere note sur la theorie de la chaleur First note on the theory of heat Memoires de l Academie imperiale des sciences de St Petersbourg series 6 1 129 133 for an abbreviated version of his proof of the divergence theorem see pages 130 131 Victor J Katz May1979 The history of Stokes theorem Archived April 2 2015 at the Wayback Machine Mathematics Magazine 52 3 146 156 for Ostragradsky s proof of the divergence theorem see pages 147 148 George Green An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism Nottingham England T Wheelhouse 1838 A form of the divergence theorem appears on pages 10 12 Other early investigators who used some form of the divergence theorem include Poisson presented February 2 1824 published 1826 Memoire sur la theorie du magnetisme Memoir on the theory of magnetism Memoires de l Academie des sciences de l Institut de France 5 247 338 on pages 294 296 Poisson transforms a volume integral which is used to evaluate a quantity Q into a surface integral To make this transformation Poisson follows the same procedure that is used to prove the divergence theorem Frederic Sarrus 1828 Memoire sur les oscillations des corps flottans Memoir on the oscillations of floating bodies Annales de mathematiques pures et appliquees Nismes 19 185 211 K F Riley M P Hobson S J Bence 2010 Mathematical methods for physics and engineering Cambridge University Press ISBN 978 0 521 86153 3 see for example J A Wheeler C Misner K S Thorne 1973 Gravitation W H Freeman amp Co pp 85 86 3 5 ISBN 978 0 7167 0344 0 and R Penrose 2007 The Road to Reality Vintage books ISBN 978 0 679 77631 4 External links edit nbsp Wikiversity has a lesson on Divergence theorem Ostrogradski formula Encyclopedia of Mathematics EMS Press 2001 1994 Differential Operators and the Divergence Theorem at MathPages The Divergence Gauss Theorem by Nick Bykov Wolfram Demonstrations Project Weisstein Eric W Divergence Theorem MathWorld This article was originally based on the GFDL article from PlanetMath at https web archive org web 20021029094728 http planetmath org encyclopedia Divergence html Retrieved from https en wikipedia org w index php title Divergence theorem amp oldid 1220375367, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.