Você está na página 1de 14

The Unapologetic Mathematician

Mathematics for the interested outsider

Gauss Law forMagnetism


Lets repeat what we did to come up with Gauss law
(https://unapologetic.wordpress.com/2012/01/11/gauss-law/), but this time on the
magnetic field.
As a first step, though, I want to finally get a good definition of current density: its a
vector field that consists of a charge density and a velocity vector , each of which is
a function of space. In our example of an infinite line current
(https://unapologetic.wordpress.com/2012/01/07/currents/), this density was
concentrated along the -axis, where the velocity was vertical. But it could exist along a
surface, or throughout space; a single particle of charge moving with velocity is a
current density concentrated at a single point.
Anyway, so the Biot-Savart law (https://unapologetic.wordpress.com/2012/01/04/the-biotsavart-law/) says that the differential contribution to the magnetic field at a point from
the current density at point is

So, as for the electric field, we want to integrate over :

Last time we spent a while noting that the fraction here is secretly a closed -form in
disguise, so its divergence is zero. This time, I say its actually a conservative vector
field (https://unapologetic.wordpress.com/2011/12/15/conservative-vector-fields/):

Indeed, this is pretty straightforward to check by rote calculation of derivatives, and Id


rather not get into it. The upshot is we can write:

where the extra term on the second line is automatically zero because the curl is in
terms of and the current density depends only on . I write it in this form because
now it looks like the other end of a product rule:

Indeed, this is clearer if we write it in terms of differential forms; since the exterior
derivative is a derivation (https://unapologetic.wordpress.com/2011/07/16/the-exteriorderivative-is-a-derivation/) we can write

for a function

and a -form . If we flip

over to a vector field

this looks like

Okay, so now we see that is the curl of some vector field, and so the divergence
of a curl is automatically zero:

Coupling this with the divergence theorem


(https://unapologetic.wordpress.com/2011/11/22/the-divergence-theorem/) like last time,
we conclude that there is no magnetic equivalent of charge, or else the outward flow
of through a closed surface would be the integral on the inside of such a charge. But
instead we find

January 12, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics |


11 Comments

Gauss Law

Rather than do any more messy integrals for special cases we will move to a more
advanced fact about the electric field
(https://unapologetic.wordpress.com/2012/01/05/the-electric-field/). We start with
Coulombs law (https://unapologetic.wordpress.com/2012/01/03/coulombs-law/):

and we replace our point charge with a charge distribution


(https://unapologetic.wordpress.com/2012/01/06/charge-distibutions/) over some
region of . This may be concentrated on some surfaces, or on curves, or at points, or
even some combination of the these; it doesnt matter. What does matter is that we can
write the amount contributed to the electric field at by the charge a point as

So to get the whole electric field, we integrate over all of space!

Now we want to take the divergence of each side with respect to . On the right we can
pull the divergence inside the integral, since the integral is over rather than . But
weve still got a hangup.
Lets consider this divergence:

Away from
this is pretty straightforward to calculate. In fact, you can do it by hand
with partial derivatives, but I know a sneakier way to see it.
If you remember our nontrivial homology classes
(https://unapologetic.wordpress.com/2011/12/20/a-family-of-nontrivial-homology-classespart-1/), this is closely related to the one we built on
the case where
. In that
case we got a -form, not a vector field, but remember that were working in our
standard
with the standard metric
(https://unapologetic.wordpress.com/2011/09/20/pseudo-riemannian-metrics/), which
lets us use the Hodge star (https://unapologetic.wordpress.com/2011/10/06/the-hodgestar-on-differential-forms/) to flip a -form into a -form, and a -form into a vector field!
The result is exactly the field were taking the divergence of; and luckily enough the
divergence of this vector field is exactly what corresponds to the exterior derivative on
the -form, which we spent so much time proving was zero in the first place!
So this divergence is automatically zero for any
, while at zero its not really welldefined. Still, in the best tradition of physicists well fail the math and calculate anyway;
what if it was well-defined, enough to take the integral inside the unit sphere at least?

Then the divergence theorem (https://unapologetic.wordpress.com/2011/11/22/thedivergence-theorem/) tells us that the integral of the divergence through the ball is the
same as the integral of the vector field itself through the surface of the sphere:

since the field is just the unit radial vector field on the sphere, which integrates to give
the surface area of the sphere: . Remember that the fact that this is not zero is
exactly why we said the -form cannot be exact.
So what were saying is that this divergence doesnt really work in the way we usually
think of it, but we can pretend its something that integrates to give us
whenever our
region of integration contains the point
. Well call this something
, where the
is known as the Dirac delta-function, despite not actually being a function.
Incidentally, its actually very closely related to the Kronecker delta
(https://unapologetic.wordpress.com/2008/05/27/dual-spaces/)
So anyway, that means we can calculate

This integrand is zero wherever


, so the only point that can contribute at all is
We may as well consider it a constant and pull it outside the integral:

where we have integrated away the delta function to get . Notice how this is like we
usually use the Kronecker delta to sum over one variable and only get a nonzero term
where it equals the set value of the other variable.
The result is known as Gauss law:

and, incidentally, shows why we wrote the proportionality constant the way we did
when defining Coulombs law. The meaning is that the divergence of the electric field at
a point is proportional to the amount of charge distributed at that point, and the
constant of proportionality is exactly .
If we integrate both sides over some region
form:

we can rewrite the law in integral

That is: the outward flow of the electric field through a closed surface is equal to the
integral of the charge contained within the surface. The second step here is, of course,
the divergence theorem, but this is such a popular application that people often call
this Gauss theorem. Of course, there are two very different statements here: one is
the physical identification of electrical divergence with charge distribution, and the
other is the geometric special case of Stokes theorem. Properly speaking, only the first
is named for Gauss.
January 11, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics |
8 Comments

Charged Rings andPlanes


Lets work out a couple more examples that may come in handy in the future, if only to
get the practice. Well start with a charged ring which is a charge distribution
(https://unapologetic.wordpress.com/2012/01/06/charge-distibutions/) on a circle.
Specifically, we may as well consider the circle of radius in the plane:
. If the charge density is , then the total charge is
.
First of all, symmetry tells us that we may as well just consider the points of the form
for
, and we will first specialize to the points
. Along this line, the field
generated by a little piece of charge on one side of the circle points across to the other
side of the circle and out further along the axis; the piece on the other side cancels
the horizontal contribution, but adds to the outward push. This outward direction along
the axis is all that we need to calculate in this case.
The Coulomb law (https://unapologetic.wordpress.com/2012/01/03/coulombs-law/) tells
us that the differential element of charge at
is
. The displacement
vector is
, and its length is
. And so the integral is

So theres some extra push near the origin, but when gets much bigger than , this is
effectively the Coulomb law again for a point source with charge $latex
the total
charge on the ring.
Pushing off of the center line is a bit rougher. The differential element of charge is the
same, of course, but now the displacement vector is
. Its
magnitude is
, leading to the integral

This is, not to put too fine a point on it, really ugly, and it would take us way too far
afield to go into it.
However, we can use what we know to determine the electric field generated by a
charged plane with a charge density of , measured in charge per unit area. At any
point
we can drop a perpendicular to the plane. We cut the plane into charged
rings of radius and width
, which gives us a (linear) charge density of
. The
result above says that the ring of radius has only a component, which is

So we integrate this out over all radii:

So the field points out from the plane, and it does not fall off with distance at all!
January 10, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics |
Leave a comment

Currents
Part of the reason that the Biot-Savart law
(https://unapologetic.wordpress.com/2012/01/04/the-biot-savart-law/) isnt usually stated
in the way I did is that its really about currents, which are charges in motion. The point
charge moving with velocity does give a sort of a current, but its so extremely
localized that it doesnt match with our usual notion of current. A better example is a

current flowing along a curve (without boundary) with a (constant) charge density
(https://unapologetic.wordpress.com/2012/01/06/charge-distibutions/) of . Its possible
to carry through the discussion with a variable charge density, but then things get
more complicated.
Anyway, just like the electric field the magnetic field obeys a superposition principle, so
we can add up the contributions to the magnetic field from all the tiny differential bits
of a current-carrying curve by taking an integral. The differential element of charge is
again
. The velocity is in the direction of the curve unit vector
and
has length . Thus the
term near a point is
. We will take the charge
density as charge per unit of distance and the speed
as distance per unit of time,
and combine them into the current
as the charge flowing through this point on
the curve per unit of time. For a curve
we have the integral:

As an example, lets take the infinite line of charge and set it in motion with a speed
along the -axis. The charge density makes for a current
up the line. Of course,
if the current is negative then the charge is just moving in the opposite direction, down
the line. The obvious parameterization is
, so we have
and
. Plugging in we find:

where Ive used a couple convenient substitutions to put the integral into exactly the
same form as last time. We can reuse all that work to continue:

We find again that the magnetic field now falls off as the first power of the distance
from the line current. As for the direction, it wraps around the line in accordance with
the famous right hand rule; if you place the thumb of your right hand along the -axis,
the field curls around the line in the same direction as your fingers.
As it happens, despite how popular the rule is its purely conventional, with no actual
physical significance. Its hard to explain just why that is right now, but it will become
clear later. For now, I can justify that it makes no difference in the effect of currents on
moving charges, since the Biot-Savart law involves a triple vector product which can be
rewritten:

which formula involves no cross products and no choice of right-hand or left-hand rules.
January 7, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics |
11 Comments

Charge Distibutions
The superposition principle for the electric field
(https://unapologetic.wordpress.com/2012/01/05/the-electric-field/) extends to the realm
of continuous distributions, with the sum replaced by an appropriate integral.
For example, lets say we have a curve
(https://unapologetic.wordpress.com/2011/04/08/curves/)
, and along this
curve we have a charge. It makes sense to measure the charge in units per unit of
distance, like coulombs per meter. We can even let it vary from point to point, getting a
function
describing the charge per unit length near the point with parameter . To
be a little more explicit, if
is the line element that measures a tiny bit of
distance near the point
, then
measures a little bit of charge near that
point.
We can now use the Coulomb law
(https://unapologetic.wordpress.com/2012/01/03/coulombs-law/) to see what electric
field this tiny bit of charge creates at a point with position vector :

since the displacement vector from


to is
. Now we can take this differential
electric field and integrate it over the curve, adding up all the tiny contributions to the
field at made by all the tiny bits of charge along the curve.
As an example, lets consider an infinite line of charge along the axis with a constant
charge density of ; a piece of the line of length will have charge . Admittedly, this is
not a finite-length curve like above, but the same principle applies. We set
,
so
and
.
Geometric considerations tell us that the electric field generated by the line at a point
is contained in the same plane that contains the line and the point. We can also tell that
the field points directly perpendicular to the line; if
then the vertical
component induced by the chunk at
is cancelled out by the component
induced by the chunk at
. Indeed, we can check that the first gives us

while the second gives us

and the vertical components of these two exactly cancel each other out.
All that remains is to calculate the horizontal component. Without loss of generality we
can consider the point
, and we must calculate the -component of the electric
field by taking the integral

We need an antiderivative of the integrand


fits the bill. Indeed, we check:

, and it turns out that

as asserted. Thus we continue the integration:

which is a nice, tidy value. More generally, we find

pointing directly away from the (positively) charged line, neither up nor down, and
falling off in magnitude as the first power of the distance from the line.
January 6, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics | 8
Comments

The Electric Field

What happens if instead of two particles we have three? For simplicity, lets just
consider the resultant force on one of the three particles; say it has charge and the
other particles have charges and , with displacement vectors and , respectively.
The Coulomb law (https://unapologetic.wordpress.com/2012/01/03/coulombs-law/) tells
us that the first of the other particles exerts a force

while the second exerts a force

As usual, we just add the forces together to get the resultant

Of course, as we add more particles we just add more terms to the sum. But always we
find that the force on the test particle with charge times some vector field:
.
Coulombs law tells us that a single point of charge at the origin gives rise to a vector
field whose value at the point with position vector is

That is, its sort of like the radial vector field only instead of getting larger as we move
away from the origin, it gets smaller, falling off as the square of the distance.
As weve just seen, the superposition principle for forces leads to a superposition
principle for the electric field: the field generated by two or more sources is the (vector)
sum of the fields generated by each source separately.
January 5, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics | 5
Comments

The Biot-Savart Law


What Im going to present is slightly different from what usually gets called the BiotSavart law, but I think its the most natural parallel to the Coulomb law
(https://unapologetic.wordpress.com/2012/01/03/coulombs-law/). As far as I can tell, it
doesnt get stressed all that much in modern coverage; in the first course I took on

electromagnetism way back in the summer of 1994 I didnt even see the name written
down and parsed what I heard as bee-ohs of r. So at least you know how its supposed
to be pronounced.
If we have two charged particles that are both moving, then they also feel a different
force than the electric one. We call the excess the magnetic force. In magnitude its
proportional to both the magnitudes of the charges and their speeds, and inversely
proportional to the square of the distance between them, and then it gets complicated.
Its probably going to be easier to write this down as a formula first.

So many cross products! Like last time, this is the force exerted on the first particle by
the second; is the vector pointing from the second particle to the first, and
is the
unit vector pointing in the same direction. From this formula, we see that the force is
perpendicular to the direction the first particle is moving the magnetic force can only
turn a particle, not speed it up or slow it down and in the plane spanned by the
direction the second particle is moving and the displacement vector between them.
Again, weve written the constant of proportionality in a weird way. The magnetic
constant
now appears in the numerator, so its units are almost like the inverse of
those on the electric constant. But weve also got two velocities to contend with; these
lead to a factor of time squared over area, resulting in mass times distance over charge
squared. In the SI system we have another convenience unit called the henry, with
symbol , defined by

which lets us write


in units of henries per meter. Specifically, the SI units give it a
value of
. Yes, I know that this makes the
proportionality constant look that much weirder, since it just works out to
, but still.
January 4, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics | 7
Comments

Coulombs Law
I want to start in on a new topic, but it might be a bit of a surprise. I havent really
talked about any applications much at all. Still, physics is a huge area of application for
mathematics, and a lot of mathematics wouldnt have been discovered without the
physical motivation.
But why didnt I talk about classical Newtonian mechanics when discussing calculus? As
it happens, the application of calculus to Newtonian mechanics is pretty
straightforward and boring; the first-pass coverage is pretty much all there is.

Electromagnatism, however, is another story. The first-pass treatment is basically all


about vector calculus, and thats great; well go over that a bit, which may be review for
some people. But theres a much deeper story to even classical electromagnetism that
uses all this stuff Ive been saying about differential geometry lately. But for now
everything will take place in regular three-dimensional space.
Anyway, we start with Coulombs law. This is something that can be experimentally
determined, but well take it as an assertion another axiom and build from there.
When we have two charged particles, they exert a force on each other. The magnitude
of the force is proportional to the magnitude of the charge on each particle, and
inversely proportional to the square of the distance between them. The direction of the
force exerted on the first particle by the second is in the direction of the vector pointing
from the second to the first if their charges have the same sign, and in the opposite
direction if they have different signs. That is, charges of the same sign push each other
apart, while charges of the opposite sign pull each other together.
Lets write this out in a formula: if the charge on the two particles are and
measured as a positive or negative multiple of some unit and if the displacment
vector from the second particle to the first is , then we have the following formula for
the magnitude of the force exerted on the first particle by the second:

since the distance between the particles is given by the length of . To get the
direction, we use the unit vector
that points from the second particle to the first.
It turns out that we also just need to drop the absolute value signs on the charges:

Now, I havent explained why the constant of proportionality is written in the weird form
, and Im not going to quite yet. Ill just say that thats all this is: a constant that gets
the scaling right, not to mention the units. On the right, weve got (other than the
constant) units of charge squared over area, while on the left weve got force, which is
mass times distance over time squared. The electric constant , thus, must carry
units of time squared times charge squared over mass times volume.
In the common SI (metric) system we measure charge in coulombs after Coulombs
law with symbol and we have a convenience unit called the farad with symbol ,
which is given by

Using these units, we can write the electric constant with units of farads per meter.
Incidentally, it has the measured value of approximately
, but the
exact value will be largely irrelevant to us.
January 3, 2012 Posted by John Armstrong | Electromagnetism, Mathematical Physics |
16 Comments

Next Entries

About this weblog


This is mainly an expository blath, with occasional high-level excursions, humorous
observations, rants, and musings. The main-line exposition should be accessible to the
Generally Interested Lay Audience, as long as you trace the links back towards the
basics. Check the sidebar for specific topics (under Categories).
Im in the process of tweaking some aspects of the site to make it easier to refer back to
older topics, so try to make the best of it for now.

Site info
The Unapologetic Mathematician
The Andreas04 Theme. Create a free website or blog at WordPress.com.
Follow

Follow The Unapologetic Mathematician


Build a website with WordPress.com

Você também pode gostar