Você está na página 1de 23

Submitted by,

V KARTHIK REDDY 12BEC0440


V BHARGAV CHOWDARY 12BEI0015
SHIVENDRE PRATAP SINGH 12BEC0230
SHIANGH SINGH PUNDHIR 12BEC0334



Mini Project
Applications of Divergence

This project is about Divergence and its applications in
various fields.






APPLICATIONS OF DIVERGENCE
Divergence theorem was first discovered
by Lagrange in 1762, then later independently
rediscovered by Gauss in 1813, by Green in
1828, and in 1826 by Ostrogradsky, who also
gave the first proof of the theorem.
Subsequently, variations on the divergence
theorem are correctly called Ostrogradsky's
theorem, but also commonly Gauss's theorem,
or Green's theorem.







INTRODUCTION :
The divergence theorem, more commonly known especially in older literature as Gauss's
theorem (e.g., Arfken 1985) and also known as the Gauss-Ostrogradsky theorem, is a
theorem in vector calculus that can be stated as follows. Let be a region in space with
boundary . Then the volume integral of the divergence of over and the surface
integral of over the boundary of are related by

The divergence theorem is a mathematical statement of the physical fact that, in the
absence of the creation or destruction of matter, the density within a region of space
can change only by having it flow into or away from the region through its boundary.
A special case of the divergence theorem follows by specializing to the plane. Letting
be a region in the plane with boundary , above equation then collapses to

If the vector field satisfies certain constraints, simplified forms can be used. For
example, if where is a constant vector , then

But

so





and






But , and must vary with so that cannot always equal zero. Therefore,

Similarly, if , where is a constant vector , then



Required to Prove: The Divergence theorem. If V is the volume bounded by a
closed surface S and A is a vector function of position with continuous derivatives, then




where n is the positive (outward drawn) normal to S.






Proof: The Divergence theorem in the full generality in which it is stated is not easy to
prove. However given a sufficiently simple region it is quite easily proved.

Let S be a closed surface so shaped that any line parallel to any coordinate axis cuts the
surface in at most two points. We will now proceed to prove the following assertion:



Denote the projection of the surface on the xy plane by R. A line erected from within R
perpendicular to the xy plane intersects the surface at two points, at a lower surface and
an upper surface. Denote the lower surface by S
1
and the upper surface by S
2
. Let the
equation of S
1
be z = f
1
(x, y) and the equation of S
2
be z = f
2
(x, y). See Fig. 1. Now



Now by the Fundamental Theorem of Integral Calculus



Substituting 3) into the right member of 2) we get











For the upper surface S
2
, dy dx = cos
2
dS
2
= kn
2
dS
2
since the normal n
2
to S
2
makes
an acute angle
2
with k.

For the lower surface S
1
, dy dx = -cos
1
dS
1
= -kn
1
dS
1
since the normal n
1
to S
1
makes
an obtuse angle
1
with k.

Consequently





Substituting 5) and 6) into 4) we get







or

which is what we wished to prove.

In the same way, by projecting S on the other coordinate planes, we can obtain



Adding 8), 9) and 10) we get

or

SOME EXAMPLES ON DIVERGENCE THEOREM:
Example 1:
Compute SFdS where





F = (3x+z77, y2sinx2z, xz+yex5)
and S is surface of box
0x1 ,0y3, 0z2.
Use outward normal n.
For F = (xy2,yz2,x2z), use the divergence theorem to evaluate
SFdS
solution: Given the ugly nature of the vector field, it would be hard to compute this
integral directly. However, the divergence of F is nice:
divF = 3+2y+x.

We use the divergence theorem to convert the surface integral into a triple integral

SFDs = BdivFdV
Where B is the box
0x1 ,0y3, 0z2.

We compute the triple integral of divF = 3+2y+x over the box B:

SFdS=103020(3+2y+x)dzdydx=1030(6+4y+2x)dydx=10(18+18+6x)dx=36+3=39
.
Example 2: Use the divergence theorem to calculate , where S is the surface of the
box B with vertices (1, 2, 3) with outwards pointing normal vector and F(x, y, z) = (x
2
z
3
,
2xyz
3
, xz
4
).
Solution: Note that the surface integral will be difficult to compute, since there are six different
components to parameterize (corresponding to the six sides of the box) and so one would have
to compute six different integrals. Instead, using Gauss Theorem, it is easier to compute the
integral (F) of B.
First, we compute (F) = 2xz
3
+ 2xz
3
+ 4xz
3
= 8xz
3
. Now we integrate this function over the
region B bounded by S













APPLICATIONS OF DIVERGENCE THEOREM:

Divergence theorem has a wide applications on differential form and integral
form of physical law and continuity equations.

Divergence theorem has a wide applications on inverse square laws.

Divergence theorem of gauss plays a central role in deviation of differential
equation in fluid dynamics, electrodynamics, gravitational fields and optics.

There are some physical applications like fluid flow and heat transfer.

Divergence theorem is also used in Archimedes principle and Laplace equation.

APPLICATION OF DIVERGENCE THEOREM IN ARCHIMEDES PRINCIPLE:

Gauss divergence theorem may be used to derive Archimedes principle for the buoyant
force on a body totally immersed in a fluid of constant density (independent of depth).





Examine an elementary section of the surface S of the immersed body, at a depth z < 0
below the surface of the fluid:

The pressure at any depth z is the weight of fluid per unit area from the column of fluid
above that area. Therefore

pressure = p = g z g is the weight of the column
z is the height of the column (note z < 0).

The normal vector N
v
to S is directed outward, but the hydrostatic force on the surface
(due to the pressure p) acts inward. The element of hydrostatic force on AS is

( ) ( ) ( ) ( ) ( )
( )
( )

pressure area direction g z S g z S = A = + A N N


The element of buoyant force on AS is the component of the hydrostatic force in the
direction of k (vertically upwards):






( )

g z S + A N k g

Define

g z = F k
v
and

dS = dS N
v
.
Summing over all such elements AS, the total buoyant force on the immersed object is


S S V
g z dS dV = =
}} }} }}}
k N F dS F
v v v v
g g g

V (by the Gauss Divergence Theorem)
( )

, , 0, 0,
V V
g z dV g z dV
x y z

c c c
= =
c c c
}}} }}}
k
v
g g V
( ) provided 0
V
g dV g
z

c | |
=
|
c
\ .
}}}

=weight of the fluid displaced.
Therefore the total buoyant force on an object fully immersed in a fluid equals the weight of
the fluid displaced by the immersed object (Archimedes principle).

APPLICATION OF DIVERGENCE THEOREM : LAPLACE EQUATION

The combination div grad , ( ) or is called the "Laplacian" differential
operator,
The equation ( ) f = 0 is called Laplace's equation. Static electric and steady state magnetic
fields obey this equation where there are no charges or current.
Any solution to this equation in R has the property that its value at the center of a sphere within
R is the average of its value on the sphere's surface. If g(r) obeys Laplace's equation inside a
spherical surface, S, of radius a, centered at r', we have






Laplace's Equation

Thus solutions to Laplace's equation are very smooth they have no bumps maxima or
minima in R and essentially "interpolate" smoothly between their values on the
boundaries of R. We prove this important fact as an application of the divergence
theorem.
This result also implies that if we know the divergence of a vector v and its curl
everywhere, these are differentiable everywhere, and v vanishes at infinity, then v is
uniquely determined. proof box( if there were two solutions v and v' with the same
divergence and curl, then on applying the double cross product identity we find that
every component of their difference obeys Laplace's equation everywhere. Its value
anywhere is then its average value on a huge circle at infinity, which is 0 by assumption.
The same conclusion holds if v and v' are required to behave at infinity in the same way,
so that v - v' must approach 0 for large arguments.)
PROOF:
Suppose our function, f(r), obeys Laplace's equation within some sphere S centered
at r':
div grad f = ( ) f = 0 inside S
We apply the divergence theorem to the vector f g - g f in the sphere with surface S
excluding a tiny sphere of radius b with surface S' having the same center. We obtain

the latter being obtained by substituting for g. The second integral on the second line
vanishes here as can be seen by applying the divergence theorem again within S and
noticing that the Laplacian applied to f is 0.
The right hand side here is the average value of f on S. The similar integral over S' is
evaluated in exactly the same way and is the average value of f on S'. We conclude that
the average value of v on any sphere with center at r' is the same. Obviously as the





sphere approaches radius 0 that average value becomes the value of f at r'.
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 theorem in
the form used above is often called "Green's Theorem." And the function g defined
above is called a "Green's function" for Laplaces's equation.
We can use this function g to find a vector field v that vanishes at infinity obeying
div v = , curl v = 0. (we assume that r is sufficiently well behaved, integrable, vanishes
at infinity etc...) Suppose we write v as grad f.
We get

(This equation is called Poisson's equation and is obeyed by the potential produced by a
charge distribution with charge density .)
A solution that vanishes at infinity is then given by

PROOF :
The same approach can be used to obtain a vector potential A and determine a
vector v obeying
v = 0, v = j
With v = A we obtain
( ( A)) = j
and with the double cross identity
( ( A)) = ( A) - ( ) A
and the "gauge" condition (which we are free to assume) ) A = 0, we find that each
component of A obeys Poisson's equation with source - the corresponding component of j.
We may therefore find a formula for each coordinate of A exactly like the corresponding
formula for the scalar potential V. Again v may be recovered from A by differentiating.
These results are of some use in the study of electromagnetic fields but they don't solve all





problems. Often the known charges and currents induce unknown charges and currents in
conducting surfaces, and one wants to determine the fields, in circumstances in which there are
some unknown charges and /or currents and you know conditions on the field at the conductor
surfaces instead of the charges and currents in them.
In this way laplace equation can be verified using divergence theorem.
AERODYNAMIC APPLICATION OF DIVERGENCE THEOREM:
- The Aerodynamic Continuity Equation
1. The surface integral of mass flux around a control volume without sources or sinks is
equal to the rate of mass storage.
2. If the flow at a particular point is incompressible, then the net velocity flux around the
control volume must be zero.
3. As net velocity flux at a point requires taking the limit of an integral, one instead merely
calculates the divergence.
4. If the divergence at that point is zero, then it is incompressible. If it is positive, the fluid
is expanding, and vice versa.
- Gausss Theorem can be applied to any vector field which obeys an inverse-square law
(except at the origin) such as gravity, electrostatic attraction, and even examples in
quantum physics such as probability density.
PRACTICAL APPLICATIONS OF DIVERGENCE THEOREM:


Remember that the divergence of a field f at a point p measures the tendency of the
underlying substance, or energy, to expand at the point p. If f is the velocity of a fluid, is the
fluid expanding (positive divergence), just passing through (0 divergence), or compressing
(negative divergence)? Note that f could cause water to move slower, but over a larger area,
and still give a divergence of 0. Let water rush into the neck of a funnel at high speed, and
move out the wide mouth at a slower speed. A depiction of f might present long arrows
clustered together, pushing the water quickly into the neck of the funnel, and short arrows
spread apart around the mouth of the funnel. It is possible for f to have divergence 0
throughout. At any point p, water comes in quickly from behind, and leaves at a slower speed
and in several directions. Water does not compress or expand at the point p; it is merely
passing through. In this sense, referring to divergence as "spreading out" is somewhat
misleading. The water does indeed spread out in space, but it does not actually expand. It
moves slower as it spreads out, to compensate. The same amount of material is involved from
start to finish.





- Here is another 0 divergence example. Consider the electric field, emanating from an
electron at the origin. This is an inverse square field. This can be written as some constant
(which we will ignore) times x y or z over r3, where r is the radial distance
sqrt(x2+y2+z2). The partial with respect to x is 1/r3 - 3x2/r5. Do the same for y and z and
add them up, giving 3/r3 - 3r2/r5, or 0. The electric field gets weaker as you move away
from its source, but it gets weaker in proportion to the surface area of the expanding
shell. This is the inverse square law, and it implies a divergence of 0 everywhere, except at
the origin, where 1/r is not defined. Electric charge bursts into existence at the origin, and
charge is conserved everywhere else.
- These are some real life examples related to divergence theorem.

Differential and Integral Operators use in Divergence Theorem:

One of the most important and useful mathematical constructs is the "del operator",
usually denoted by the symbol V (which is called the "nabla"). This can be regarded as a
vector whose components in the three principle directions of a Cartesian coordinate
system are partial differentiations with respect to those three directions. Of course, the
partial differentiations by themselves have no definite magnitude until we apply them to
some function of the coordinates. Letting i, j, k denote the basis vectors in the x,y,z
directions, the del operator can be expressed as



All the main operations of vector calculus, namely, the divergence, the gradient, the curl,
and the Laplacian can be constructed from this single operator. The entities on which we
operate may be either scalar fields or vector fields. A scalar field is just a single-valued
function of the coordinates x,y,z. For example, the static pressure of air in a certain
region could be expressed as a scalar field p(x,y,z), because there is just a single value of
static pressure p at each point. On the other hand, a vector field assigns a vector v to each
point in space. An example of this would be the velocity v(x,y,z) of the air throughout a
certain region.

If we simply multiply a scalar field such as p(x,y,z) by the del operator, the result is a
vector field, and the components of the vector at each point are just the partial derivatives
of the scalar field at that point, i.e.,



This is called the gradient of p. On the other hand, if we multiply a vector field v(x,y,z)
by the del operator we first need to decide what kind of "multiplication" we want to use,
because there are two different kinds of vector multiplication, commonly called the dot





product and the cross product. For two arbitrary vectors a = a
x
i + a
y
j + a
z
k and b = b
z
i
+ b
y
j + b
z
k the dot product ab and the cross product ab are defined as



Intuitively, the dot product is a scalar equal to the product of the magnitudes of a and b
times the cosine of the angle between them, and the cross product is a vector
perpendicular to both a and b (with direction determined conventionally by the "right
hand rule") and whose magnitude is equal to the product of the magnitudes of a and b
times the sine of the angle between them. Since the dot product yields a scalar, it is often
called the "scalar product". Likewise the cross product is often called the "vector
product".

The dot product of V and a vector field v(x,y,z) = v
x
(x,y,z)i + v
y
(x,y,z)j + v
z
(x,y,z)k
gives a scalar, known as the divergence of v, for each point in space:



The cross product of V and a vector field v(x,y,z) gives a vector, known as the curl of v,
for each point in space:



Notice that the gradient of a scalar field is a vector field, the divergence of a vector field
is a scalar field, and the curl of a vector field is a vector field. Can we construct from the
del operator a natural differential operator that creates a scalar field from a scalar
field? Actually we already have the ingredients for such an operator, because if we apply
the gradient operator to a scalar field to give a vector field, and then apply the divergence
operator to this result, we get a scalar field. This is sometimes called the "div grad" of a
scalar field, and is given by



For convenience we usually denote this operator by the symbol V
2
, and it is usually
called the Laplacian operator, because Laplace studied physical applications of scalar
fields (x,y,z) (such as the potential of an inverse-square force law) that satisfy the
equation , i.e.,








This formula arises not only in the context of potential fields (such as the electro-static
potential in an electric field, and the velocity potential in a frictionless ideal fluid), it also
appears in Poisson's equation and in the fundamental wave equation



where c is the speed of propagation of the wave. Incidentally, this suggests another
useful differential operator, formed by bringing the right hand term over to the left side to
give



This operator is known as the d'Alembertian (because the wave equation was first studied
by Jean d'Alembert), and it could be regarded as the dot product of two "dal" operators



where i,j,k,l are basis vectors of a four-dimensional Cartesian coordinate system.

One of the most important theorems in vector analysis is known as the Divergence
Theorem, which is also sometimes called Gauss' Theorem. This is essentially just an
application of the fundamental theorem of calculus



This enables us to express the integral of the quantity df/dx along an interval in terms of
the values of f itself at the endpoints of that interval. Now suppose we are given a scalar
function f(x,y,z) throughout a region V enclosed by a surface S, and we want to evaluate
the integral of the quantity cf/cx over this entire region. This can be written as








where the three integrals are evaluated over suitable ranges to cover the entire region
V. For any fixed values of y = y
0
and z = z
0
the function f(x,y
0
,z
0
) is totally dependent on
x, so we can evaluate the integral along any line parallel to the x axis through the region
V for any particular y
0
,z
0
using the fundamental theorem of calculus



where a,b are the x values at which the line intersects the surface S. (For ease of
description we are assuming every line parallel to the x axis intersects the surface in only
two points, but this restriction turns out to be unnecessary.) We now just need to
integrate the above quantity over suitable ranges of y
0
and z
0
to give the integral
throughout the entire region V. Notice that the two points (a,y
0
,x
0
) and (b,y
0
,x
0
) are both,
by definition, on the surface S, at opposite ends of a line interval parallel to the x
axis. Hence, we can cover all the contributions by integrating the value of the function
f(x,y,z) over the entire surface S.

Of course, we recognize that f(a,y
0
,z
0
) is subtracted from the total, whereas f(b,y
0
,z
0
) is
added, so it's clear that we need to apply a weighting factor to f(x,y,z) at each point when
we integrate. The value for points on the "low" end of the intervals must be subtracted,
and the value for points on the "high" end of the intervals must be added. Furthermore, if
we want to evaluate the integration over the surface area of S by integrating over dS, we
need to scale the magnitude of the contribution of each point to give the appropriate
weight to each incremental region dS of the surface, because a portion of the surface that
is very oblique to the yz plane doesn't "contribute as many line segments" parallel to the x
axis as does an equal increment of the surface that is parallel to the yz plane.

Clearly the necessary scale factor at each point is the cosine of the angle between the
positive x axis and the normal to the surface at that point. Notice that this automatically
gives us the appropriate sign for each contribution as well, because (for a convex surface)
the normal to the surface will have a positive x component on one end of the interval and
a negative x component on the other, which means the cosines of the respective angles
will have opposite signs. (Again, for ease of description we are assuming a convex
surface, but it's not hard to show that this restriction is unnecessary, since a given line
parallel to the x axis can have multiple segments which can be treated separately.)

Consequently, we can write our original integral as



where cos(n,i) denotes the cosine of the angle between a unit vector normal to the surface
and a basis vector in the positive x direction. Naturally, since the coordinate axes are





symmetrical, we could arrive at analogous results with x replaced by either y or z, and i
replaced by j or k respectively.

Now suppose we are given a vector field F(x,y,z) with the components F
x
(x,y,z),
F
y
(x,y,z), and F
z
(x,y,z). Using the above results and simple additivity we have



The integrand on the left side is VF, i.e. the divergence of F. Also, notice that cos(n,i),
cos(n,j), and cos(n,k) are the components of the normal unit vector n, so the integrand on
the right side is simply Fn, i.e., the dot product of F and the unit normal to the
surface. Hence we can express the Divergence Theorem in its familiar form



Several interesting facts can be deduced from this theorem. For example, if we define F
as the gradient of the scalar field (x,y,z) we can substitute V for F in the above
formula to give



The integrand of the volume integral on the left is the Laplacian of , so if is harmonic
(i.e., a solution of Laplace's equation) the left side vanishes. The integrand of the right
hand integral is the normal "flux" through the surface, so we see that the integral of the
normal flux over any closed surface (in a region that everywhere satisfies Laplace's
equation) is zero. It follows that there can be no local maximum or minimum inside a
region where Laplace's equation is satisfied, because such a point would, by definition,
be completely enclosed by a surface of everywhere positive (or everywhere negative)
normal flux, making it impossible for the integral of the flux over the surface to
vanish. This, in turn, implies that if is constant over an entire closed surface (where
Laplace's equations is satisfied) then the value of is constant throughout the enclosed
volume, because otherwise the volume would have to contain a local maximum or
minimum. Combining these facts with the additivity of harmonic functions, we can
conclude that there is a unique harmonic function within an enclosed region that has
specified values on the enclosing surface, because if
1
and
2
are two functions
satisfying those boundary conditions, the harmonic field
1
-
2
is zero over the entire
boundary, and therefore it vanishes throughout the interior as well.

If we allow non-zero charge or mass density (or sources or sinks) in the enclosed region,
such that Poisson's equation V
2
= -4t is satisfied, then the left hand integral is the net





charge contained within the volume, and this equals the integral of the normal flux over
the enclosing surface. This is often called Gauss' law of electrostatics, and it constitutes
one of Maxwell's equations.

Another important consequence of the Divergence Theorem can be seen by noting that
the scalar quantity Vn at any given point on the surface equals the partial derivative
c/cn where n is the displacement parameter in the direction normal to the
surface. Furthermore, if the surface in question is a sphere - which we can assume
without loss of generality is centered at the origin - then the normal displacement
parameter n is equal to the radial parameter r. So for such a spherical surface we have,



Recall that, in terms of spherical coordinates r, u, and |, where u is latitude (zero at the
North Pole) and | is longitude (zero on the positive x axis) the basic line element in space
is,



so an incremental change du at constant r and | corresponds to a change ds = r du,
whereas an incremental change d| at constant r and u gives ds = r sin(u) d|. Hence the
surface element in terms of u and | is dS = r
2
sin(u) du d| , so the preceding integral can
be written as,



Applying Leibniz's Rule for the derivative of an integral, and multiplying by 4t/4t, this
becomes,



The quantity in the square brackets is just the mean value of (r,u,|) on the surface of the
sphere, and this equation shows that the derivative of the mean value with respect to the
radius r vanishes, so the mean value of on the surface of a sphere centered at any fixed
point is independent of the radius. Considering the limit as r approaches zero, it's clear
that the mean value of a harmonic function on the surface of a sphere (of any radius) is
equal to the value of at the center of the sphere.






SOME INTERESTING FACTS OF DIVERGENCE THEOREM:
Let's say I have a rigid container filled with some gas. If the gas starts to expand but the
container does not expand, what has to happen? Since we assume that the container does not
expand (it is rigid) but that the gas is expanding, then gas has to somehow leak out of the
container. (Or I suppose the container could burst, but that counts as both gas leaking out of the
container and the container expanding.)
If I go to a gas station and pump air into one of my car's tires, what has to happen to the air
inside the tire? (Assume the tire is rigid and does not expand as I put air inside it.) The air inside
of the tire compresses.
These two examples illustrate the divergence theorem (also called Gauss's theorem). 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 theorem says that the total expansion of
the fluid inside some three-dimensional region W equals the total flux of the fluid out of the
boundary of W. In math terms, this means the triple integral of divF over the region W is equal
to the flux integral (or surface integral) of F over the surface W that is the boundary of W (with
outward pointing normal):
WdivFdV=WFdS.
I hope that this makes sense intuitively from the above two examples. In the first example, the
gas expanding meant divF>0 everywhere in W, the inside of the container. Therefore, the net
flux out of W, WFdS, must also be greater than zero, i.e., the gas must leak out through the
container walls W. In the second example, by pumping air into the tire W, I insisted that the net
flux out of the tire, WFdS, must be negative (since there was a net flux \textbf{into} the tire,
and we are assuming an outward pointing normal). By the divergence theorem, the total
expansion inside W, WdivFdV, must be negative, meaning the air was compressing.
Notice that the divergence theorem equates a surface integral with a triple integral over the
volume inside the surface. In this way, it is analogous to Green's theorem, which equates a line
integral with a double integral over the region inside the curve. Remember that Green's theorem
applies only for closed curves. For the same reason, the divergence theorem applies to the
surface integral
SFdS
only if the surface S is a closed surface. Just like a closed curve, a closed surface has no boundary. A
closed surface has to enclose some region, like the surface that represents a container or a tire. In other
words, the surface has to be a boundary of some W (i.e., S=W), as described above. You cannot use
the divergence theorem to calculate a surface integral over S if S is an open surface, like part of a cone





or a paraboloid. If you want to use the divergence theorem to calculate the ice cream flowing out of a
cone, you have to include a top to your cone to make your surface a closed surface.

APPLICATIONS OF ELECTRODYNAMICS IN DIVERGENCE THEOREM:
The way to define a three-dimensional delta function is just to take the product of three one-
dimensional functions:

The integral of this function over any volume containing the origin is again 1, and the integral of
any function of is a simple extension of the one-dimensional case:

In electrostatics, there is one situation where the delta function is needed to explain an apparent
inconsistency involving the divergence theorem. If we have a point charge at the origin, the
electric field of that charge is

According to the divergence theorem, the surface integral of the field is equal to the volume
integral of the divergence of that field:

where the integral on the left is over some closed surface, and that on the right is over the volume
enclosed by the surface. In electrostatics, the integral on the right evaluates to the total charge
contained in the volume divided by

Now for the catch. If we calculate (in spherical coordinates) for the point charge, we get,
since only the radial component of the field is non-zero:











At this stage, we might be tempted to say that the derivative is zero (since the derivative of any
constant is zero), but the problem is that at we also have a zero in the denominator, so we
have the indeterminate fraction of zero-over-zero. Thus although it is true that
everywhere except the origin, we know from the divergence theorem that
so we must have

and

These two conditions can be satisfied if

Você também pode gostar