Sunteți pe pagina 1din 7

Introduction

The purpose of these notes is to explain what I do. Hopefully it will be accessible to any
reader willing to think a bit and with a basic mathematics education that includes high
school algebra. Let me know if it does not accomplish this goal. Before we get started I
want to make a note about variables and symbols. I am not planning on using a lot of
equations or even symbols, but I do not plan to shy away from them either. In Stephen
Hawking's book "A Brief History of Time," which is aimed, I believe, at a similar class of
readers, Hawking says that he will not use any equations at all because he thinks it will
scare people off. Personally, I think that the whole point of equations is that they make
things easier. That's why mathematicians, engineers, and scientists use them! Now we often
get carried away because we use them so much and the average person does not use them
so much. But that does not mean that you can't find them useful too. Mathematical
expressions in symbols allow us to be more precise without having to write volumes. I hope
that those of you not use to using equations will not find this daunting and those of you
who are will be able to understand everything quite precisely with the aid of some
equations.

My view of differential calculus


Although I claim to do geometry, as those of you who learned calculus may know, much
geometry can be understood with the help of calculus. Thus we shall use some (okay, a lot)
of calculus to help us prove some interesting things about geometry. At least, that's the plan.
But DON'T FEAR!!!! I will hopefully give you a good idea of what you need to know
about calculus. So if you don't know a thing about it, now you can get the gist. So before
we start with the math, perhaps we should motivate. One view (maybe Newton's, for one)
of how things work is there are these forces always working and everything happens due to
these forces which are always working. For instance, if we have a piece of cork floating in
a river, there are currents acting on it at all time and that causes the cork to move. The
important point is that the forces are ALWAYS working on it. In contrast, we could think of
the second hand on a clock that clicks to the next tic mark every second. This isn't changing
continuously like the cork, but a force is acting on it at discrete instances, that is, once
every second. Calculus is not designed for this type of situation, but for the first one. So
based on the philosophy that there are forces acting on our "objects" (by this I mean
whatever we want to consider: the cork's spatial location, it's velocity, it's density, or

anything else), we need something that measures how they change. So if we have a cork in
a river, it will be moved by the currents and we want some way of understanding how the
current causes the position of the cork to change. This is something called a vector field. At
every point in the river, it tells where the cork is going to go. Well, not quite, since after it
moves a little bit, it is at another point and thus moves according to another vector (all a
vector is is an arrow that tells me which direction to move and how fast). This is the magic
of calculus. It lets us deal with what seems like a mess. Essentially you move in a straight
line in the direction the vector says, but only for a really small amount of time, and then
you are at a new point and move in a straight line according to the vector there for a small
amount of time, and so on. This is called the flow of a vector field (because it is the way the
water flows in the river). This is essentially all a differential equation is. It associates a
vector (which says where to go) to each point in your space.

What is a manifold?
In this section we are going to describe the general class of things we want to study. We are
going to start with a large class of things and then add more and more structure to it so that
it is more and more limited. In mathematics we often are interested in classifying objects up
to some form of equivalence. For instance, we may want to be able to distinguish TYPES
of US coins. In other words, we want to know how many different types of coins there are,
and if we have two, we want to know, for instance, if they are both pennies or if one is a
nickel. A mathematician might call this classifying up to the type of coin (this is the
equivalence). To contrast, we may want to classify up to type of coin and year it was
minted. In the first case, if we have a 1986 penny and a 1997 penny, they are the same up to
equivalence. But in the second case they would be different up to equivalence. Thus it is
important to know what we consider to be the same and what we consider to be different. In
the first case, our equivalence was very weak, since we only consider two things different if
they were different types of coins. Thus the only objects we have are pennies, nickels,
dimes, quarters, half dollars, and dollars (6 objects total). In the second case our
equivalence was stronger because we consider two things different if their years of minting
are different. In this case we have a lot of different objects, including 1999 pennies, 1990
pennies, 1987 dimes, 2000 nickels, 1967 quarters, 1990 dimes, etc. Now we are going to
consider objects the same based on a number of different equivalences. With manifolds, we
are mostly going to be concerned with two being "topologically" the same and being
"geometrically" the same. These two terms are actually not as specific as we might hope,
but I don't want to get into the specifics of different kinds of topological equivalence or
geometric equivalence. I will describe the relevant ones. Now the idea of two things being
topologically the same is perhaps one which you have not encountered before. Two things
are topologically the same if they have the same shape in the sense that one can be

deformed into the other one without "breaking" it. I'm being a little vague because the
actual definition is not very instructive to the beginner. But you already understand the
basics of this concept. Now consider if you had a perfect circle that you drew with a
compass and a circle that you drew freehand. Now only one is REALLY a circle, right? The
other one isn't a perfect circle in that there is no one point in the center which is EXACTLY
the same distance from every point on the circle. But, nonetheless, if you were to show it to
someone, they would recognize it as a circle, even though it's not a circle. This is in some
sense the difference between being topologically the same and geometrically the same.
Geometrically, the second is not a circle, since there is no center. But topologically they are
equivalent. Topologically both are the same, because we could change one of the "circles"
slightly to make it into the other. Now when I say change, I am only allowing certain kinds
of changes. For instance, you are not allowed to cut the circle unless you reattach the two
ends to each other. So, for instance, if your circle were a rubber band you could stretch it
but not let it break. You could even cut it, tie a knot in it, and reattach the two ends. These
would all be topologically equivalent. This idea that you can stretch things however much
you want has caused topology to be dubbed clay geometry. The typical example given is
that topologically a coffee mug is the same as a doughnut. (Think about why!) But really
we just removed some of our restrictions on two things being the same. We are just about
ready to learn what a manifold is. We first need to understand what Euclidean space is. You
already know what Euclidean space is. It's just what you consider to be space. For instance,
the infinite line (number line) is a Euclidean space. So is the infinite plane. So is the space
we live in. So what is it exactly? Well, the number line is just the set of all real numbers (so
we allow all decimal expansions, even infinite ones that give irrational numbers like
and ). This is one dimensional Euclidean space since there is only one degree of freedom.
You can only go forward or backward if you live on the line. Now consider the plane. Here
we have two degrees of freedom. We can go forward/backward or left/right. This is two
dimensional Euclidean space. It is important that forward/backward is not the same
direction as left/right, for then we would have a line, right? Now, if we want to describe a
location on the plane, we can fix a point and then say how far forward/backward to go and
how far left/right to go. Thus there are two real numbers associated to a position in the
plane. This is what we really mean by two degrees of freedom or two dimensions. Now in
space we have three directions we can go: forward/backward, left/right, and up/down. Thus
we get three dimensional Euclidean space. You can similarly imagine Euclidean space of
any dimension. All you need is some number of different degrees of freedom and a location
(called a point) is determined by that many different numbers, which tell you how far you
are in each direction from some fixed point which we will call the origin (note that the
origin is 0 units away from itself in all directions). Note that topologically the size of the
units is unimportant. Now a manifold is just something which locally looks like Euclidean
space. Consider, for instance, the surface of the Earth. For ages it was believed to be a

plane. Why? Because from what we can see it is a plane. (Topologically, hills, mountains,
and valleys don't change this, right?) But we know that it actually curves back on itself and
is topologically a sphere instead of a plane. So it is a manifold. Locally it looks like a plane
(Euclidean two dimensional space). We would call it a two dimensional manifold because it
is locally like two dimensional Euclidean space. Now here is a question: is the universe we
live in Euclidean three space or is it some other three dimensional manifold? As far as I can
tell, this is still unknown. One goal of mathematics, however, is to see what it could
possibly be and then maybe we could eliminate things on the list with experiments. (This is,
of course, under the assumption that the everywhere in the universe looks essentially like
the way we see it, which seems a reasonable assumption.)

Riemannian Geometry
We have described what we are looking at topologically, but we are also interested in
geometry. Riemannian geometry is one way of looking at distances on manifolds. This
seems an easy enough concept when you first think of it, but after further though we realize
it is not so easy. Sure we know how to measure distances on a plane. The shortest distance
between two points is a straight line, right? So just draw the line and measure the distance
(first we set what unit measure is, for instance 1 meter, and then compare the distance we
want to measure to our set standard unit distance, say the meter stick). But on a sphere how
do we measure distance? Or on a torus (the surface of a doughnut)? A sphere we can think
of as living in Euclidean 3-space and then just say that the distance between 2 points on the
sphere is just the distance between those points in 3-space, as described by the Pythagorean
theorem, which says that for points

and

the distance between them is

which is the fact that the length of the hypotenuse (longest side) of a right triangle is equal
the square root of the sum of the squares of the other two sides.

In addition, we can change this slightly by deforming the sphere in 3-space so that it is still
topologically a sphere but not geometrically a perfect sphere. This is a perfectly legitimate
way to define distances. It has one disadvantage, however. It requires that the sphere be

"embedded" in 3-space. In other words, we can think of the sphere as living in 3-space, but
maybe our geometric sphere cannot live in 3-space. This concept is a little weird in the case
of the sphere, so let's look at this problem one dimension down. Consider the circle. Now,
we can think of the circle as living in Euclidean 2-space (the plane). Now I claim that we
can find topological circles that we cannot put in Euclidean 2-space! These are called knots
and they are exactly what you consider to be knots. Take a circle, then disconnect it, then
tie a knot in it, then reattach the ends. We now still have a topological circle. If you lived on
the circle, your couldn't tell the difference than if there was no knot tied in it. But we can't
crunch the knot down so it lives in the plane. Thus we need to have the circle live in 3space and we are fine. So we need to first know which sized (dimension) Euclidean space
we live in to use this method. Of course, there is no reason to believe that any manifold
lives in some appropriately sized Euclidean space, although this fact is true (but hard to
show). We are going to skirt these issues by taking another approach to distances. Suppose
we could measure the length of curves. Then we could simply define the distance between
two points to be the length of the shortest path, if one exists. Even if one does not exist, we
could express the distance as the largest lower bound for lengths of paths between the two
points (this is called the infimum). This isn't too important so let's assume that we can find
a shortest path. So we could measure distances if we could measure the lengths of paths.
This is where a Riemannian metric comes in. So if you are driving, how would you
measure the distance you traveled? Well, one way is to look at your speedometer and
remember how fast you are going at every point and then consider how long it took you to
get where you were going. Since speed is just distance per unit time (like miles per hour)
we just multiply the speed times the time and we get the distance. That's how far we
traveled. The only problem is that our speed changes and our direction changes, so we
actually need to take into account which direction we are traveling (this means using the
velocity vector instead of just the speed, which is the length of the velocity, ignoring
direction). Also, our formula only works if the velocity does not change. But our velocity
changes. What do we do? We chop our time intervals up into smaller segments which have
constant velocity. This doesn't quite work because there is no time when we are going at a
constant velocity (most likely), but as we chop up the time interval more and more, our
approximation is more and more accurate. In the limit, we end up with an integral and can
calculate the actual length by computing that integral (this involves some basic calculus).
For culture, let's look at how we would write this:

which we read as the length

of the curve (path)

is the sum (the integral sign

German 's' standing for sum) of the speed of the curve, denoted

is a

, multiplied by a

really small length of time, denoted


. The the term
just means we measure the path
from time 0 to time . Anyway, the what we get out of this discussion is that only need to
know the velocity at every point. The velocity at every point can be considered as a vector
(arrow) indicating where we are going (direction) and at what speed. It is tangent to our

path because it tells us which direction to move. And it is only defined at each point. We
just need to be able to measure these velocities and then we can compute lengths of curves.
Thus we need to be able to measure the length of vectors at every point. This is what a
Riemannian metric does. It provides us with a way of measuring vectors at every point.
Since these are local things and we understand what vectors at different points in Euclidean
space are, we can do the same things on manifolds, since locally they look like Euclidean
space. So the Riemannian metric is a function defined at every point that takes two vectors
and gives a number. Technically, a Riemannian metric must be a symmetric bilinear form
(otherwise it is called a Finsler metric). This condition isn't too important for us to
understand at this point, but it is easy to state so why don't we do it anyway. We express the
Riemannian metric as

and we express vectors at a point

. then we can measure quantities like

in our manifold as

. The fact that is symmetric

means:
and bilinearity (when coupled with symmetry) is the following condition:
But you probably don't have to understand this too much. I just get carried away
sometimes. So the important thing is that the Riemannian metric gives us a way to measure
lengths of vectors at each point in the manifold, and also gives us a way of measuring
lengths on the manifold.

Curvature
What's the difference between a plane and a sphere? Well, for one, a sphere is curved while
a plane is flat. Another example of a flat space is the cylinder, since we can just unroll it
and it becomes a plane. A sphere can't simply be unrolled, though. It would need to be
stretched to flatten it out. Unrolling can be done without doing any stretching. Now in
Riemannian geometry we have a way of making this general concept into a mathematical
concept. This is called curvature. We say a sphere is positively curved. In fact, it has
constant curvature which means the curvature at every point is the same and positive
(whatever that means). Similarly a plane is flat, so the curvature is zero (whatever that
means) at every point. You might ask what negative curvature means. Negative curvature is
exemplified by a saddle shape, which is the shape of a Pringles potato chip. If the curvature
is negative and the same at every point, we have what is called Hyperbolic space. This was
one of the first examples of what is called "Non-Euclidean geometry," sacrilege to
classicists! Curvature, it turns out, is a very complicated concept, especially in higher
dimensions. It is not completely understood. We do know that curvature determines the

metric and places restrictions on the topology. Thus if we want to understand our manifold,
it is often easiest to try to understand the curvature. Easier said than done, though, huh? So
let's give a couple of ideas of what curvature means in two dimensions. If we draw a
triangle in the plane, we find that the sum of the interior angles is 180 degrees (which is,
incidentally radians and can even be thought of as the definition of , some number that
happens to be approximately 3.141592653589793). No matter which triangle we draw, the
sum of the angles is 180 degrees. But what happens if we draw triangle on the sphere? Let's
say the sphere is really big (like the size of the Earth) and we draw a little triangle on it, say
on the your driveway. What will the sum of the angles be? They would essentially be 180
degrees, right? But now let's say we have a small globe and draw a triangle with a right
angle (90 degrees) at the north pole and whose third side is on the equator. (Recall that a
straight line on the sphere consists of great circles, so the equator is a straight line and the
other two straight lines are two longitudes meeting at right angles.) Thus we have a triangle
which is essentially one-fourth of the northern hemisphere or one-eighth of the sphere.
What is the sum of the angles? If you look closely, you see that there are three right angles!
Hence the sum of the angles is
degrees. It's bigger than 180
degrees! This is what positive curvature means. If you have a triangle in positive curvature,
the sum of the angles of a triangle is bigger than 180 degrees. Negative curvature, similarly,
means the sum of the angles is less than 180 degrees. You might think about what this
means on a Pringles potato chip! In the standard model of negative curvature, you can even
have triangles which have a sum of angles almost 0! Another way to think about curvature
is in terms of area of disks with the same perimeter. Let's say we have a circle. What is a
circle? It is the set of points the same distance away from a single point, say distance 1.
(This is rather arbitrary, so we just choose 1 because it makes the formulas come out better.
Hooray!) This makes sense for any surface with a distance on it. Now, in flat curvature (the
plane), the circle bounds a disk with area (this is another place where we could have
defined ). Now, suppose the disk was made out of clay and you wanted to stretch that
disk to fit on a sphere? You'd have to stretch it! Thus the enclosed area is larger than .
Now what if you wanted to put it on a Pringles potato chip? It would be too big and you
would have to make the area smaller. This is another difference between positive and
negative curvature. And the actual curvature relates to how much you need to stretch or
shrink the disk at each point to fit on the surface. Now curvature takes place at a point, so
we can have positive, negative, and flat curvature on the same surface. Picture, for instance,
the surface of a doughnut (called a torus). If you think about it, the outside surface has a
positive curvature and the part inside (the "hole") has negative curvature. In between, there
is some zero (flat) curvature. (Can you guess where? If it's flat, you should be able to sit a
plane on it so that it touches on a line or anything else with dimension more than just a
point.)

S-ar putea să vă placă și