Additional comments related to material from the
class. If anyone wants to convert this to a blog, let me know. These additional
remarks are for your enjoyment, and will not be on homeworks or exams. These are
just meant to suggest additional topics worth considering, and I am happy to
discuss any of these further.
- Thursday, December 9:
- For one complex variable, we can consider f(z) = 1/z, which has an
isolated singularity (a pole) at z=0. What about two complex variables? If we
try f(z1,z2) = 1/z1 then it has poles in the entire plane z1=0, z2 arbitrary.
What else could we try? In real variables we have 1/r^2 in 2 or 3 dimensions,
such as 1/(x^2+y^2). If we try that here we get f(z1,z2) = 1/(z1^2+z2^2). The
problem is that this has a pole in the plane z2 = i z1. There is no isolated
singularity in several complex variables -- that's the key takeaway of the
Domain of Holomorphy arguments. In several variables, it is even more
restrictive to be holomorphic.
- See the comments from Tuesday, December 7th for more on fractal geometry
and iteration of complex functions. We didn't talk about the
3x+1 Problem, but it's a great
example of what happens when you iterate a simple function. Very complicated
behavior emerges, with some fascinating symmetries and patterns. See the
article by Jeff
Lagarias for more information.
-
Finally, a few points about optimal algorithms / efficient algorithms.
- We then ended the day by a quick tour through the subject. The power of
the Cauchy Integral Formula is striking; again and again it appears (analytic
is holomorphic, the Cauchy inequalities which lead to Liouville's theorem, the
Riemann Mapping Theorem, ...). This integral representation is entirely
different than what happens in real analysis, and is responsible for much of
what we see.
- Here is my solution to
the look-up table question posed by my brother Jeff from
MathWorks (formerly The
Mathworks). There are many related questions that are of interest -- if anyone
would like to pursue these applications further, please let me know.
- It's hard to choose the final theorem of the day,
so I'll just provide the following link and let you choose.
- Tuesday, December 7:
- Several
Complex Variables is a beautiful subject, of which we can only skim the
surface. There are many similarities between several and one complex variable,
as well as some important differences. While Cauchy's Integral Formula still
holds, an enormous difference is the notion of a
domain of
holomorphy. An open set is a domain of holomorphy if there is a
holomorphic function on U that we cannot extend across the boundary. Note that
the punctured unit disk is a domain of holomorphy in one complex variable
(simply take the function 1/z); however, the n-dimensional analogue (when n >
1) is not a domain of holomorphy! This is a bit similar to some of the issues
in applying Green's theorem in the plane versus Stokes' theorem in 3-space, as
a point singularity is not so bad for Stokes as we can deform the curve in the
higher dimension.
- It's interesting finding a function that cannot be extended. We saw one
way to do this is to take a dense subset of the boundary and look at Sum_n
(1/4^n) 1/(z-z_n). Of course, one needs to do some analysis to prove this
provides an example; the key fact (at least for circles and annuli) is that
any point in the open set is at least a certain distance from the boundary).
- See Hartog's
theorem for (yet another!) difference between several real and complex
variables.
- Key reading for Fractal Geometry and Iterations of Functions:
Julia and Fatou sets,
Mandelbrot set and a
nice
zoom here (see the
Mandelbrot explorer and the
Chaos game, which
you can play here),
the Newton fractal,
and Chaos theory. A
good program to investigate fractals is
XaoS.
- First use of fractal geometry in a movie:
Star Trek
II: The Wrath of Khan: The Genesis Effect (also known as the `good' Star
Trek movie). As long as we're doing Star Trek II scenes, great ones include
the battle with Khan (and
how things work on a starship) and the
final
battle, the Kobayashi
Maru test, though sadly cannot find quickly the 'I don't like to lose'
clip and the 'hours instead of days now we have minutes instead of hours'.
- For more on efficient algorithms, see
Chapter 1 of my book
An Invitation to Modern Number Theory (Section 1.2).
- The theorem of the day? Well, since we're doing Star Trek let's do
Fermat's Last
Theorem. The Star Trek connection? There's
Picard talking about how
he likes to try to prove unsolved math problems like Fermat's last theorem
(I don't believe he ever says he's related to the complex analyst
Picard) as
well as Dax (start around
3:25 on the clip) who talks about a previous host having the most original
approach since Wiles' successful proof.
- Thursday, December 2:
- In our application of
Laplace's method
to proving
Stirling's formula for n!, we adjusted the function Phi so that it
vanished at the critical point. There is of course no need to do so; some
authors prefer to renormalize like this, some don't. One of the homework
problems builds on this by looking at the moments of the standard normal, and
we're able to get nice formulas for (2m-1)!!. Of course, one could also derive
this from (2m)! / (2^m m!). Looking at it like this, it is clear why all the
sqrt(pi)'s vanish.
- Our sketch of the proof of the Prime Number Theorem shows how complex
analysis results come into play, and the importance of the location of the
zeros of zeta(s). A very important fact is that zeta(s) is non-zero for Re(s)
= 1. See for example
Chapter 3 of my
book An Invitation to Modern Number Theory (see Exercise 3.2.19). The
proof there is a very clever application of what has been hailed as the most
important inequality in mathematics, namely that for real x, x^2 >= 0.
- A big technique in studying sums in number theory is
partial summation
(aka, integration by parts: the discrete version). The idea is that frequently
certain sums, though not exactly what we want to study, are easier to handle.
It is thus a worthwhile trade to work with these sums first and then modify
them to return to what you want. Another instance where this occurs is in
attacking problems through the
Circle Method, such
as the Goldbach
Problem.
- Note that in two problems this week (Stirling's formula and the Prime
Number Theorem) we saw the importance of replacing a simple, discrete object
with an integral. This occurs throughout higher mathematics. There are
numerous advantages to integral representations. For example, in the analysis
for the Prime Number Theorem we computed a weighted integral of the
logarithmic derivative of zeta(s) two different ways. We want to
compute contour integrals of this quantity -- complex analysis is screaming at
us to do so; however, if we are computing it as an integral then we need to
work on the `other side' with an integral as well.
- There were questions on whether or not our arguments on pi(x^{1- epsilon})
< x^{1-epsilon} were good. It's not too hard to show that A x / log(x) < pi(x)
< B x / log x for some A, B with 0 < A < 1 < B < oo.
Chebyshev was
the first to do so, around 1850. For more details,
see here
as well as the nice
article here (on Erdos' proof).
- If you want to know more about the functional equation of the zeta
function, as well as proofs of why it isn't zero if Re(s) > 1, I'm happy to
meet and chat. You can also look at
Chapter 3 of my
book An Invitation to Modern Number Theory, where we spend a lot of time
trying to emphasize the role complex analysis plays in studying the primes.
One of the most interesting items is that it is more natural to weight the
primes than to count them equally. Further, it's easier to count all prime
powers and not just primes.
- For the Theorem of the Day: (1)
Euler's Partition Identity (I think something very close to this, if not
this itself, has surfaced in some discussions with people in the class). It's
related to generating functions and expansions. (2)
The footballer's
theorem. This is a nice article by Harold Boas on H. Bohr (N. Bohr's
brother) and some of his results on the boundary of complex analysis and
number theory. It discusses many of the concepts we've looked at recently,
including analytic continuation and the differences between uniform, absolute
and conditional convergence.
- Tuesday Nov 30:
- We've talked a few times about analytic continuation. It's always worth
asking what the source of the analytic continuation is. For the geometric
series, it's the geometric series formula. For the Gamma function, it's
integration by parts. For the Riemann zeta function, it's the
Poisson
summation formula, which relates sums of a nice function at integer
arguments to sums of its
Fourier transform at integer arguments. There are many proofs of this
result. In my book on number theory, I prove it by considering the periodic
function F(x) = Sum_{n = -oo to oo} f(x+n). This function is clearly periodic
with period 1 (if f decays nicely). Assuming f, f' and f'' have reasonable
decay, the result now follows from facts about pointwise convergence of
Fourier series. There are other proofs; in chapter 4 of our textbook we find a
nice proof based on the residue theorem -- it's worth reading this to see yet
another example of the power of the residue formula.
- Laplace's method
is a terrific way to approximate the values of integrals as a certain
parameter tends to infinity. From it we can derive
Stirling's formula
for n!. Not surprisingly, there are other ways to prove Stirling's formula.
What I like about this proof is that it highlights the need for studying the
analytic continuation of the Gamma function. We need the nice integral
representation which gives us n! for any positive n. This is not the only
application of the method -- see the appendix in the book for other examples.
- Laplace's method is the first, but not only, approach for estimating
certain integrals when a specified parameter tends to infinity. Frequently we
want to approximate complex integrals such as Int_{x = -oo to oo} Exp[i s f(x)]
g(x) dx; this leads to the
Stationary Phase Method.
- For more on the connections between complex analysis and probability
(especially when a sequence of moments uniquely determines a probability
distribution, which sadly is not always the case), see my notes from Math 341
(Probability),
available here.
- A very important part of our analysis was showing when an integral is
well-defined. This frequently involves understanding the decay of the
integrand as |x| --> oo as well as |x| --> 0. A frequent tool in such an
analysis is
L'Hopital's Rule. Key identities are x^r log x --> 0 as x --> 0 for any r
> 0, and x^r exp(-x) --> 0 as x --> oo for any r.
- It is very useful to be able to express a quantity as an integral; this
was a key input in using Laplace's method to derive Stirling's formula.
Another situation where this arises is differential equations, where we can
often write the solution as a fixed point of an integral operator. See
Picard's
Iteration Method, as well as my notes from
Math 209: Differential Equations.
- The value Gamma(1/2) is extremely important. It is related to the
normalization constant for the normal distribution (and thus, due to the
Central Limit
Theorem, occurs throughout probability).
Here is a link to some notes from a
book I'm writing; these notes include information on the Gamma function and
sketches of proofs of Stirling's formula (as well as bounds towards Stirling's
formula) -- comments welcome!
- The theorem of the day is the
Asymptotic (Half) Liar Formula.
- Tuesday Nov 23:
- The notion of
analytic continuation is one of the most important in complex analysis,
allowing us to extend the definition of a function. The Riemann
zeta function is defined by zeta(s) = Sum_{n = 1 to oo} 1/n^s for Re(s) >
1, and by unique factorization also equals Prod_{p prime} (1 - 1/p^s)^{-1}. We
can analytically continue the zeta function to a meromorphic function on the
complex plane, with only one pole, a simple pole with residue 1 at s=1.
- One of the most famous and important problems in mathematics is the
Riemann Hypothesis (this is in
fact one of the seven Clay millenial prize problems;
click here for the
rules). The full problem list is below; familiarizing yourself with these
seven problems is a great way to see what issues are driving current
mathematical research. Note that often Professor Morgan runs a senior seminar
on `The Big Questions', where these and other topics are covered.
-
In class we defined pi(x) to be the number of primes at most x. We discussed
Euclid's argument which shows that pi(x) tends to infinity with x, and
mentioned that with some work one can show Euclid's argument implies pi(x) >>
log log x. As a nice exercise (for fun), prove this fact. This leads to an
interesting sequence: 2,
3, 7, 43, 13, 53, 5, 6221671, 38709183810571, 139, 2801, 11, 17, 5471,
52662739, 23003, 30693651606209, 37, 1741, 1313797957, 887, 71, 7127, 109, 23,
97, 159227, 643679794963466223081509857, 103, 1079990819, 9539, 3143065813,
29, 3847, 89, 19, 577, 223, 139703, 457, 9649, 61, 4357....
This sequence is generated as follows. Let a_1 = 2, the first prime. We apply
Euclid's argument and consider 2+1; this is the prime 3 so we set a_2 = 3. We
apply Euclid's argument and now have 2*3+1 = 7, which is prime, and set a_3 =
7. We apply Euclid's argument again and have 2*3*7+1 = 43, which is prime and
set a_4 = 43. Now things get interesting: we apply Euclid's argument and
obtain 2*3*7*43 + 1 = 1807 = 13*139, and set a_5 = 13. Thus a_n is the
smallest prime not on our list genereated by Euclid's argument at the nth
stage. There are a plethora of (I believe) unknown questions about this
sequence, the biggest of course being whether or not it contains every prime.
This is a great sequence to think about, but it is a computational nightmare
to enumerate! I downloaded these terms from the Online Encyclopedia of Integer
Sequences (homepage is http://www.research.att.com/~njas/sequences/
and the page for our sequence is http://www.research.att.com/~njas/sequences/A000945 ).
You can enter the first few terms of an integer sequence, and it will list
whatever sequences it knows that start this way, provide history, generating
functions, connections to parts of mathematics, .... This is a GREAT website
to know if you want to continue in mathematics. There have been several times
I've computed the first few terms of a problem, looked up what the future
terms could be (and thus had a formula to start the induction). One last
comment: we also talked about the infinitude of primes from zeta(2) = pi^2/6.
While at first this doesn't seem to say anything about how rapidly pi(x)
grows, one can isolate a growth rate from knowing how well pi^2 can be
approximated by rationals (see http://arxiv.org/PS_cache/arxiv/pdf/0709/0709.2184v3.pdf
for details; unfortunately the growth rate is quite weak, and the only way I
know to prove the needed results on how well pi^2 is approximable by rationals
involves knowing the Prime Number Theorem!).
- Key inputs in our study of the zeta function were
unique
factorization and special values, such as the
sum of the
reciprocals of the squares equaling pi^2/6 and the
harmonic sum diverges.
We were able to use the harmonic sum to get information about the sum of the
reciprocals of the primes, `proving' there are enough so that this sum
diverges. Calculating Brun's
constant (the sum of the
reciprocals of twin primes) led Nicely to discover the Pentium
bug; a nice description of the discovery of the bug is given at http://www.trnicely.net/pentbug/pentbug.html.
- The purpose of today's lecture, rushed in places as it was, was to give
you a sense of the need for rigor. Innocent looking statements can fail;
interchanging limits and integrals can fail. If we're not careful, we can be
led to believe we have a proof of the Riemann Hypothesis! I find it
fascinating that the analytic continuation of a limit is not necessarily the
limit of the analytic continuation, and in fact there is at least one
situation where the two differ! This is a very nice result which we proved
today. I'll write more extensive notes on this for those who wish.
- Thursday Nov 18:
- We saw how much easier the
pacman contour was
than the branch cut method. For the integral of 1/(x^3+1) from 0 to oo, we
tried one last approach, namely splitting the integral into two parts, from 0
to 1 and then from 1 to oo, and then using the
geometric
series formula. Such a splitting of an integral is quite common in number
theory, allowing us to exploit the different behavior of the function in the
two regions. In fact, this is one of the key steps in the most common proof of
the
functional equation of the Riemann zeta function.
- Functional equations allow us to take a function that is initially defined
in one region and extend the definition elsewhere. In addition to zeta and
L-functions, another common (and very important example in probability,
statistics and number theory) is the
Gamma function,
which generalizes the factorial function.
- The Riemann
zeta function is defined by zeta(s) = Sum_{n = 1 to oo} 1/n^s for Re(s) >
1, and by unique factorization also equals Prod_{p prime} (1 - 1/p^s)^{-1}. An
interesting consequence of its analytic continuation is the `identity' that 1
+ 2 + 3 + 4 + ... = -1/12; amazingly,
`facts' like this arise in modern physics (string theory)!
- Another nice identity, arising from the continuation of the geometric
series, is that 1 + 2 + 4 + 8 + ... = -1. Interestingly, this is true
2-adically.
- A very important one in probability / statistics is (-1/2)! = sqrt(pi);
this comes from the
Gamma function, which interpolates the
factorial function.
Bohr and
Mollerup proved that if you assume a continuation of the factorial
function has three reasonable properties than it must be the Gamma function.
- In some sense, these functional equations are similar to the Cauchy
integral equation. There we saw that knowing a holomorphic function on the
boundary was enough to uniquely determine it everywhere.
- There are numerous
identities / series expansions for simple functions of
pi; the difficulty is knowing
how to manipulate what one has to get one of these desired relations (I
wrote a short note on how to prove Wallis' formula from the Student
t-distribution).
There
are lots of interesting algorithms to calculate digits of pi efficiently.
If you're interested in `experimental' mathematics (like this or more standard
topics), check out this site.
- We talked a bit about
cellular automata
today. A big early example is Conway's
Game of Life.
Stephen Wolfram has
also done a lot of work in the field, see especially his book
A New Kind Of
Science.
- In the spirit of when certain values determine a function, the theorem of
the day is the
Cauchy - Kovalevskaya theorem.
- Tuesday Nov 16:
- Today we returned to contour integration, one of the most useful parts of
complex analysis. We first considered trigonometric integrals; see my
lecture
notes on examples of contour integrals. Our first example was integrating
Sin[theta]^2n, and uses what I call the `bring it over' method. Here we have
some integral I, and by integrating by parts we find I = K + cI, where K is
something known. Solving, we find (1 - c)I = K, or I = K/(1-c); so long as c
is not 1 (which would force K to be 0) we can solve for I. This is one of my
favorite integration techniques, where we write the quantity we want in terms
of itself!
-
We saw that
we could simplify the recurrence relation and get the simple, closed form
expression, but we had to make some semi-clever choices of how to multiply by
1 to express our answer in terms of common symbols. It's not a horrible idea
to multiply by (2n)!!, as that `completes' the factorials in the numerator.
The trick is then pulling out the powers of 2 in the denominator. The best way
to see that this is a good thing to do is to note that once we have (2n)! in
the numerator, we want to have k! (2n-k)! in the denominator for some k, as
this will give us a binomial coefficient.
-
The
double factorial is a very important combinatorial function. It arises in
many places, perhaps most importantly as the moments of the standard normal,
where the 2m-th moment is (2m-1)!!; recall the double factorial means take
every other term until you reach 2 (if you start at an even number) or 1 (if
you start at an odd number). One interpretation of (2m)!! is that this is the
number of ways to pair off 2m objects into m pairs of 2. Here is a derivation:
there are (2m choose 2) ways to choose two elements to be our first pair,
(2m-2 choose 2) ways to choose the next pair of elements, down to (2 choose 2)
ways to choose the last pair. We then have to divide by m!, as we don't care
which pair label two elements are given. We thus have (2m) (2m-1)/2 * (2m-2)
(2m-4)/2 * ... * (2)(1)/2 / m! = (2m)! / 2^m m! = (2m)! / (2m (2m-2) * ... *
2) = (2m-1) (2m - 3) * ... * 1.
-
Here's
another way to see the above formula, this time by induction. The base case is
clear. For the inductive step, assume the number of ways to pair 2m objects in
m pairs of 2 is (2m-1)!!. If we now have 2m+2 objects, note object 1 must be
paired with something -- there are 2m+1 ways to choose its match. We are now
left with 2m objects, which by induction can be paired in (2m-1)!! ways.
Multiplying gives (2m+1)!! as desired.
-
We
converted trigonometric integrals to contour integrals, and saw the advantages
of this. This method of integration is quite useful in numerous problems,
especially in
Circle Method problems in number theory. If you want to read more about
the Circle Method, I can pass along the relevant chapter from
my book.
-
Our next
approach involved
branch cuts.
Here we show that we can actually use the multivaluedness of the complex
logarithm for good (or at least to evaluate some integrals). Not surprisingly,
for many problems we have choices as to how to proceed. For the integral from
0 to oo of 1/(x3 + 1) we have two options. One is to introduce the
logarithm function and exploit the branch, while the other is to choose a
giant pie contour. In general, it's frequently a trade-off between choosing an
easy function and an easy contour. We have to make the clever choice of
introducing the logarithm in the branch method, but the resulting algebra was
a bit easier to simplify.
-
Note how
painful it was to attack the integral via the branch cut. We had to do an
enormous amount of book-keeping to show that as epsilon --> 0 and R --> oo
that everything converged to what we wanted. There are countless special cases
to look at, we found that epsilon log(2R) had to tend to 0, .... Note the
write-up in the Bristol notes skips all these arguments, and jumps to the
answer in just a few lines. As you do more and more of these integrals, you
can `see' the cancellation / convergence. For example, the integral over the
semi-circle of radius epsilon is no problem as epsilon log(epsilon) --> 0 as
epsilon --> 0. The remaining pieces are `standard' (once you've done enough
such problems!).
-
For the
theorem of the day, I though it would be nice to have something connecting
combinatorics with complex analysis.
Here's a nice one relating certain permutations and e.
- Thursday Nov 11: TBD
- We finally made it to the proof of the
Riemann Mapping Theorem! The proof begins with a powerful
simplification. We first show that we may may our
simply connected
proper open subset of the complex plane conformally into the unit disk.
Technically this is nice, as we now are working in a compact space. The key to
this is the existence of the
complex logarithm,
and we see now why simply connected is so important -- without this we would
not be able to start the proof!
- One of course should be careful about saying that it is impossible to
prove a result without resorting to using specific facts, even though those
facts might seem quite obviously necessary to use. A terrific example is the
elementary proof of the Prime
Number Theorem (which says that as x --> oo, the number of primes at most
x is asymptotic to x/log x). It turns out that this statement is equivalent to
the fact that the Riemann zeta function zeta(s) = Sum_{n = 1 to oo} 1/n^s (or
actually its meromorphic continuation) has no zero on the lien Re(s) = 1. This
is quite clearly a complex analytic statement. It was thought that there could
be no `elementary' proof of this (elementary doesn't mean easy; it just means
without using complex analysis), but if there were one, boy would it open our
eyes! Both statements are false.
See this article by
Dorian Goldfeld for the history of the proof of the Prime Number Theorem
(and the priority dispute).
- It's not inappropriate to talk about the Prime Number Theorem on the day
we do the Riemann Mapping Theorem, as a plan of attack was sketched by Riemann
in his classic (and only!) paper in number theory!
An English translation is
available here; the
original (in German) is available here.
- It's a fascinating subject as to what we can say about the boundary under
the conformal equivalence. In some cases, such as polygons, we have formulas
available for extensions. This is the
Schwarz -
Christoffel formula. In general, it is hard to discuss what happens at the
boundary.
If the boundary is nice, however, a bit more can be said.
- The final step of the proof of the
Riemann Mapping Theorem involves a useful technique. We take an object
with a certain maximal Property I; if it doesn't have another desired Property
II, we show we can find another valid object with even larger value of
Property I, a contradiction. I see this as a variant of the following
argument. We prove all positive integers exceeding 1 have a prime
factorization. If not, let x be the smallest integer not having a prime
factorization. Clearly x is not prime, so x = ab with a, b > 1. But as x was
the smallest integer without a prime decomposition, a and b have a prime
decomposition, which gives one for x. The argument here is a bit more
involved, but similar in spirit.
- A theme running through our proof was
replacing one complicated function with a composition of simpler ones. This
way we only need to understand the simpler maps and how they piece together.
You may have seen this in other classes, such as
elementary matrices
in linear algebra (which is used for proving results ranging from
Gaussian
elimination to the
Change of Variable
theorem). It's frequently a good trade to do lots of simple problems
rather than one hard one; thus we do all these maps of our region Omega to get
it contained in the unit disk.
- This is related to the
extra credit problem on look-up tables (due December
3). The Babylonians used
base 60. If they wanted to multiply x and y, they'd need a 60 x 60 table!
Well, we can do a bit better as multiplication is commutative and see we only
need to store 60 x 59 / 2 + 60 = 60 x 61 / 2 = 1830. Now, carrying around
enough stone tablets
to record over 1800 entries is not fun! I've heard that the way they
multiplied was to note x y = ( (x+y)^2 - x^2 - y^2 ) / 2. At first this seems
foolish, as we now have three multiplications and one division
and two subtractions; however, this is a great exchange! The reason is that we
need only have a table of squares, and that's just 60 or maybe 120 entries.
This is a lot less to lug. In many modern engineering applications, it is
essential to find efficient ways to do operations.
- The theorem of the day seems clear: The
Prime Number Theorem!
- Tuesday Nov 9: TBD, including
- The
Arzela -
Ascoli Theorem did not depend on complex analysis (the first part of the
proof of Montel's
theorem heavily used Cauchy's integral formula, another wonderful example
of the utility of that result). I find it nice that
Cauchy's
integral formula makes another appearance; we had used it a lot earlier in
the semester, but not as much lately.
- Related to holomorphic functions are
harmonic functions.
These two have the wonderful property that their value at a point can be given
by an integral (here it is the
mean value property). Not surprisingly, a lot of results for harmonic
functions are similar to holomorphic functions.
- A central theme of the day was doing the
analysis on a `nice' set and then showing that sufficed. Specifically, it is
convenient to work on a compact set, as compact sets are made for uniform
arguments. Remember a key property is that given any compact set, if we have
an open cover then there is a finite subcover (see also the
Heine-Borel
theorem). There's lots of fun covering lemmas, for instance, the
Vitali covering
lemma. In today's lecture we showed we could do the analysis on a compact
sets, and then using the method of exhaustion we write our open set as an
increasing union of compact sets. You might recall seeing something similar to
this in a geometry class, namely
Archimedes
determination of pi.
- Some of you may have seen the
Arzela -
Ascoli Theorem in a course in
Functional Analysis.
There are numerous uniformity statements, such as the
Uniform
Boundedness Principle (also known as the Banach - Steinhaus theorem).
- A huge part of our analysis today needed some
set theoretic and/or topological results. We needed to be able to find a
dense subset of any
open set in the complex plane, and we also needed to generalize
Cantor's
diagonal argument to our situation. The diagonal argument is used to show
the existence of
transcendental numbers without actually constructing them.
- As long as we're mentioning Cantor, I'll add
his Continuum
Hypothesis to the list of comments, and note that it was resolved by my
mathematical grandfather,
Paul Cohen,
using his theory of
forcing.Cohen's
mathematical father was
Zygmund, who is very famous for work in trigonometric (Fourier) series.
You can look up people's
geneologies online here.
- The comments above suggest a natural choice
for the Theorem of the Day:
Cantor's uncountability theorem.
- Thursday Nov 4:
- The analysis pre-reqs for the proof of the
Riemann Mapping Theorem are significant, but provide a great way to review
many of the concepts you've seen in previous classes. One of the first items
we need is that of
uniform continuity, followed (or perhaps in parallel with) that of a
compact set. It's
frequently an interesting (but delicate!) matter to determine what type of
convergence we have. Big theorems in the subject are
Weierstrass' polynomial approximation theorem and
Fejer's theorem
(which gives us uniform approximation of continuous functions by finite
trigonometric polynomials, and Taylor expanding these gives Weierstrass'
theorem). Fejer's theorem tells us a weighted
Fourier series can
converge better than the original Fourier series.
- For the convergence of Fourier series, one of
the best results is
Carleson's theorem, which gives us almost everywhere pointwise convergence
of the Fourier series of an
L2 function to the function. If the function is not continuous it's easy
to construct a Fourier series diverging at that point; a fun fact is how much
of an overshoot there is (this is called the
Gibbs overshoot,
after J. W. Gibbs,
one of the first American scientists of note on the world stage).
- Other interesting pathological functions
include the Devil's
staircase and
Weierstrass'
continuous but nowhere differentiable function!
- A big issue in approximation theory is how
well a Fourier series converges, and whether or not it converges to the
original function. Related to the Fourier series is the Fejer series. Both
arise from convolving
our function f with a
kernel; the difference is that the
Fejer kernel has much
nicer convergence properties than the
Dirichlet kernel
(which gives the standard Fourier series). The Weierstrass and Fejer theorems
involve the concept of an
approximation to the identity (sequences of functions tending towards the
Dirac delta
functional). For the Weierstrass result, there are many proofs. One of my
favorites leads to an explicit construction, using the
Landau kernel functions (you have to scroll down on the page). This
involves finding non-negative functions that integrate to 1 on [-1, 1] and
have most of their mass near 0. This choice is K_n(x) = c_n (1 - x^2)^n. It's
a fun exercise to determine c_n. I think the easiest way involves a great
technique, the `Bring it over' method of integration. We proceed by induction.
Let I_n = Int_{-1 to 1} (1 - x^2)^n dx. Then I_{n+1} = Int_{-1 to 1} (1 -
x^2)^{n+1} dx = Int_{-1 to 1} (1 - x^2)^n (1 - x^2) dx = Int_{-1 to 1} (1 -
x^2)^n dx - Int_{-1 to 1} x (1 - x^2) x dx, where we wrote the last
integral like this to facilitate integrating by parts. Recalling the
definition of I_n gives I_{n+1} = I_n - Int_{-1 to 1} x (1 - x^2)^n x dx.
Letting u = x and dv = (1 - x^2)^n dx, we see du = 1 and v = -(1 - x^2)^{n+1}
/ (2n+2). The boundary term vanishes, and we find I_{n+1} = I_n - 1/(2n+2) Int_{-1
to 1} (1 - x^2)^{n+1} dx = I_n - I_{n+1}/(2n+2). Bringing the unknown I_{n+1}
over to the left, we find ((2n+3)/(2n+2)) I_{n+1} = I_n or, assuming I haven't
made an algebra error yet, I_{n+1} = (2n+2) I_n / (2n+3). We now have a nice
recurrence relation. I find this method somewhat miraculous -- we get the
expression we don't know on both sides, and then solve for it!
- I originally had I_{n+1} = (n+1) I_n / n,
which I knew couldn't be right. Clearly I_{n+1} < I_n, and this has I_{n+1} >
I_n. The mistake I made was that I forgot the minus sign in going from dv to
v.
- Related to approximations to the identity are
mollifiers, which play a
very important role in understanding many properties of number theoretic
functions.
- Convergence of sequences of functions is a
`central' question in mathematics -- just think about the
Central Limit
Theorem from probability!
- One of the most important questions to ask in
a given situation is how the quantities in play depend on the various
quantities. For this reason, it's often a good idea to have your parameters
subscripted. For example, in
Central Limit
Theorem we have that the (appropriately normalized) sum of independent
random variables converges to a standard normal. What's fascinating is that
such a statement seems to be independent of the fine structure of the
probability distribution. Where the structure comes into play is in the
rate of convergence. See for instance the
Berry-Esseen
Theorem.
- If anyone is interested in pursuing these
topics in greater detail, let me know and I'll make the Fourier analysis
chapter of my book available.
- The Theorem of the Day is the
Heine-Borel Theorem on the compactness of [0,1], which ties in nicely with
much of our conversations.
- Tuesday Nov 2:
- The
Schwarz lemma plays a
central role in the analysis of many problems in complex analysis. There was a
very nice exposition on its history and generalizations that just appeared in
the Mathematical Monthly; click
here for the article.
- The result is a bit surprising, namely that
we cannot have the first derivative (in absolute value) of an automorphism of
the unit disk fixing the origin exceeding 1 in absolute value. Such a
statement fails in the real case; see the
Mathematica notebook I
wrote here. There we considered the function f_a(x) = (a+1) x / (1 + a x2)
for a non-negative. If a is at most 1 then this is an automorphism of the
interval [-1,1], but it ceases to be for a exceeding 1. Note the derivative at
the origin is 1+a, which can be as large as 2. Is this the best we can do? Can
we have a larger derivative here?
- There are many great choices for additional
topics here. In addition to the article on the
Schwarz lemma,
another great choice is the
Bieberbach
conjecture (proved by de Branges).
- You should do a lot of the algebra to understand the map psi_alpha(z) =
(alpha - z) / (1 - alphabar z). Show that you can rotate and make alpha real,
and then try to understand this case (perhaps at the cost of having a rotation
outside).
- There are many applications of conformal
mappings, especially automorphisms of the upper half plane.
Mobius (or
fractional linear transformations) play a central role in the subject,
with many applications in number theory and other fields (the
cross ratio and geometry is a fun one). Important subgroups are
SL(2,R) and
SL(2,Z) (also known as the
modular group). The latter is an important ingredient in building
modular forms
(important examples of which arise from
elliptic curves).
The
j-invariant has many wonderful properties, and is even connected (through
`moonshine')
to the monster group mentioned in 10/28.
- Why are these maps so important? The modular
forms they lead to are generalizations of periodic functions, and we know how
important the periodic functions are! A central question becomes the
determination of the proper region to study these generalizations. Just as it
suffices to study a periodic function of period 1 in any interval of length 1
(such as [0,1) or [-1/2, 1/2)), we look for a basic region to study these
modular forms. This set is called the
Fundamental Domain.
- As a final (academic) note, I thought it
might be fun to end each day's comments with a link from the site `Theorem
of the Day'. I'll start today with their first choice, the
Four Color Theorem. It's not a bad idea to go to this site and familiarize
yourself with the theorems and statements.
- For a non-academic final note, the promised
well-wishing from Spaceballs:
May the Schwartz be with
you. My favorite scenes are `then
is now', `they've
gone to plaid', `industrial
strength', 'no sir, I
didn't see you...'. Okay, it's a great movie!
- Thursday Oct 28:
- A great way to build intuition about
conformal maps is to look at lots and lots of examples. The wikipedia page on
conformal pictures
has some nice examples of what various maps do.
- We looked at sequences of simple maps today
to build more complicated ones. This is a common theme in mathematics, namely
try to reduce everything to a combination of basic `moves'. Probably the most
important example you've seen is in linear algebra, with
elementary matrices.
This has numerous applications, ranging from
Gaussian
Elimination to the
Change of
Variables formula.
- Our exponential map f(z) = exp(iz) led to
some interesting behavior. We found vertical strips were conformally
equivalent to the upper half plane, with the map almost extendable to being
from the boundary to the boundary. In fact, we could almost get a vertical
strip to be conformally equivalent to the complex plane minus the origin. One
of the homework or suggested problems will investigate this in greater detail,
and will show that a non-simply connected region cannot be conformally
equivalent to a simply connected one.
- One of the maps we looked at was from the
circle to the diamond, given by (u,v) --> (u|u|, v|v|). It's a nice
calculation to show that this map is differentiable once but not twice.
- The book
`The
Only One Club', in my opinion, has an ending that is at odds with
Russell's Paradox.
- The big object we studied at the end of class
was the
Automorphism Group of an open set. There are lots of examples listed on
the wikipedia page. What's nice is that this is one of our first connections
between analysis and algebra in this course. One of the major motivations for
studying the
Riemann Mapping Theorem is that it allows us to pass from
knowledge of the automorphism group of the unit disk to that of other regions.
We'll prove this on Tuesday.
- One of the most famous Automorphism Groups is
The Monster Group,
which is the automorphism group of the Griess algebra.
- Another famous one is the automorphism group
of the Leech Lattice.
There's a fascinating story about how
John H. Conway
discovered this in a tad over 12 hours one Sunday (he had set aside several
hours two days a week for months to work on it). The Leech Lattice has
important applications in
sphere packings,
which have applications in
coding theory.
-
John H. Conway
(one of my professors when I was at Princeton) is a big player in group
theory. He's a fascinating character, and some of his work will make a nice
way to round out the additional comments for the day.
- Tuesday Oct 26:
- We began by trying to find interesting
products subject to constraints (problems #7b and #7c from Chapter 5). A good
way to look at these is to have a split in definition for the even and odd
subscripts; it just cleans up the algebra enormously. For #7b we tried a_{2n}
= f(n) and a_{2n+1} = -f(n), and were led to a nice equation for f(n). For
#7c, we tried to find a sequence so that (1 + a_{2n}) (1 + a_{2n+1}) = 1 with
a_k --> 0. It's very important to be able to find these examples, as they
illustrate how essential conditions are in theorems. For example, #7c showed
us that we don't need the series Sum a_n to absolutely converge for the
product Prod (1 + a_n) to converge; that's a sufficient but not a necessary
condition.
- Here's a fun problem: is there a non-negative
function f(x) such that Int_{0 to oo} f(x)^k dx is finite for integer k < 2010
but infinite for k > 2010? Try to think of one before reading. At first it
seems absurd; if the integral is finite, surely f(x) --> 0 and thus f(x) >=
f(x)^k; however, it is possible that f is exploding on smaller and smaller
subsets. For example, let f be a triangular function with spikes near the
integers. We have f(n) = n (for n a positive integer) and f decays linearly to
0 at n +- 1/n^2009. Then Int_{0 to oo} f(x)^k dx = Sum_{n = 1 to oo} n^k /
n^2009; this converges for k < 2010 but diverges for k >= 2010. With a bit of
work, you can replace this f with an infinitely differentiable function! A
good function to know for problems like this is f(x) = 0 outside [a,b] and on
[a,b] is given by f(x) = exp(1/(a-x) + 1/(x-b)).
- For another fun example, let f_n(x) be zero
everywhere save [1/n - 1/2n, 1/n + 1/2n], f(1/n) = 2n and f linearly decreases
to 0 at the edges. Then Int_{0 to 1} f_n(x) dx = 1, lim f_n(x) = 0 for all x,
and lim Int f_n is not Int lim f_n. Switching orders of integration and a
limit can be tricky, and leads to various uniformity estimates entering
problems. Big theorems on when this can be done are the
Monotone
Convergence Theorem (MCT) and the
Dominated
Convergence Theorem (DCT). (For completeness, see also
Fatou's lemma.)
- Here's another fun challenge. Find a sequence
a_n of real numbers that tends to zero, converges conditionally but not
absolutely, but the sum of a_n^3 diverges?
- There are lots of examples like the above;
Uryshon's lemma is
another situation. See also the
Wikipedia entry on
cut-off functions.
- Our alternate proof of the
Open Mapping Theorem used analytic properties of our map. In particular,
if we standardize we may assume f(z) = z^n (1 + g(z)) with g holomorphic and
g(0) = 0. While this is always an open map, if n=1 more is true -- this Taylor
series is invertible!
The Wikipedia page briefly talks about the formal inverse power series.
While this proof was a lot longer than using Rouche's theorem, it led to some
important results which can be used in studying the
Riemann Mapping Theorem.
- Conformal equivalence is, not surprisingly,
an equivalence
relation. The reflexive property is trivially shown, and transitivity is
mostly easy. The difficult part there is showing the composition of
holomorphic functions is holomorphic. The easiest way to do that is via the
chain rule. It is
possible to show this directly by a brute for expansion, namely the
composition of convergent power series is a convergent power series, but the
algebra is a bit painful.
Oh yeah. We are thus in the situation where, after proving the Riemann
Mapping Theorem, we can conformally map any open simply connected proper
subset of the complex plane to the unit disk. We can do the analysis in the
unit disk, and then map back. This is amazing, as we have an infinitely
differentiable map!
- A logical question to ask is what happens on
the boundary of our open set. A good answer is provided, at least for many
cases, by the
Schwarz -
Christoffel formula.
- Finally, there are big theorems from analysis
such as the
Inverse Function Theorem and the
Implicit
Function Theorem. What is nice about these results are that we have simple
conditions to test as to when we can express some variables in terms of
another. These are big results that are used to prove the existence of
solutions to some problems.
- Thursday Oct 21:
- There are many proofs of the
Open Mapping Theorem. The book treats it as a consequence of Rouche's
theorem; it is, however, also possible to derive it just from the fact that f
is holomorphic and thus analytic. It's worthwhile doing exercises such as
this, as this allows us to see how `deep' a theorem or result is. As always,
you should be asking: does a similar statement hold for real analytic
functions, and if not why not?
- We discussed an alternative approach to
proving the Open Mapping Theorem. Whenever you see a theorem, you should
immediately think of the following: (1) Let me try testing it with some
special cases to build intuition; (2) Why do they have the conditions they do?
What consequences do we get from these conditions?
- For the first point, we studied the function
f(z) = zn. We were led to this by looking at the conditions and
seeing that we could standardize our general, non-constant holomorphic
function to f(z) = zn (1 + Σ_{k = 1 to ∞} ck
zk), and of course the simplest case is when the ck's
are all zero. We looked at the map z --> zn on a disk centered at
the origin. While it was easy to see what happens there, this is a little
dangerous, as the point z=0 is a special point for this map. We really need to
investigate it elsewhere. Fortunately by rescaling it suffices to do one other
point, say z=1.
- The next item was to think about what
theorems we have at our disposal once we assume f is holomorphic. One big one
is that a holomorphic function is analytic, so perhaps we can use the series
expansion to advantage (and this is of course related to writing the function
as we did above). The key insight is that we can replace zn (1 +
Σ_{k = 1 to ∞} ck zk) with (z
h(z))n for some nice h(z). This is essentially Newton's
Generalized Binomial
Theorem, where we have (1 + Σ_{k = 1 to ∞} ck
zk)1/n. It turns out to be relatively easy to get a
series expansion for (1+x)1/n for |x| < 1. We were left with one
last point: can we solve z h(z) = c for some fixed c and h(z) = 1 + ...? It
seems clear that the answer should be z = c plus some small correction. How do
we find that correction? This is essentially the inverse function theorem.
Note that if g(z) = z h(z) with h(z) = 1 + ... then g'(0) is not zero! The
Inverse
Function Theorem now asserts an inverse for z near 0.
- A key ingredient in our proof of the Open
Mapping Theorem was Rouche's Theorem, which follows from the
argument principle.
The argument principle is very useful in complex analysis and its applications
in number theory. One of my favorites is in numerical verifications of the
Riemann hypothesis.
We can use the argument principle to count how many zeros should be in a big
box containing the critical line Re(s) = 1/2 from say -T to T, and
`investigate' how many zeros lie on this line. If the two numbers agree, we
have verified the Riemann hypothesis in this region. Amazingly, it isn't too
`hard' to numerically find zeros of the zeta function on the critical line.
This is done by using the Intermediate Value Theorem!
We have π-s/2Γ(s/2)ζ(s) is real valued when Re(s) = 1/2, and has
the same zeros as the Riemann zeta function on the critical line Re(s) = 1/2.
We now just move up the critical line and record sign changes; every time
there is a sign change we've found a zero (this follows from the
Intermediate
Value Theorem); we can compute the zeros to whatever order we want simply
by taking smaller and smaller steps.
- Tuesday Oct 19:
- Today we discussed the
Weierstrass product and hinted at a `better' version, the Hadamard
product. The problem is we need to be very careful when dealing with infinite
sums or products. The first step is always to determine exactly what sequences
have a chance of leading to a nice result. After a little thought, we realized
that if there is an accumulation point in the zeros then our function is
identically zero. This gives our first result, namely that the sequence of
zeros is unbounded and without an accumulation point. The unboundedness is
crucial; it allows us to know that from some point onward we have |z/an|
< 1/2, and thus we can use log(1 - z/an). This was important as we
wanted to write 1 - z/an as exp(log(1 - z/an)), as we
could then combine exponential factors and convert multiplication to addition.
This is a powerful principle -- it's important to think about how to arrange
algebra.
- In the course of proving the existence of an
infinite product with prescribed zeros, we used the exponential convergence
factors Ek(z) = exp(z + z2/2 + ... + zk/k).
These factors are good for several reasons. First, of course, is that they are
never zero. Second, (1-z) Ek(z) = exp(log(1-z) + z + z2/2
+ ... + zk/k) for |z| < 1/2. We see that the sum is canceling the
first k terms in the Taylor expansion of the logarithm, and this will give us
good convergence.
- We split our infinite product into two
pieces. We fixed z and looked at all an with |z/an| <
1/2. As our sequence is unbounded, this condition must be true for all n
sufficiently large, say all n > N. We analyze the factors for n < N separately
for this z, and then use our analysis above for (1 - z/an) Ek(z/an).
We again encounter a situation where we have a free parameter. This time, we
have the freedom to choose k. What should k be a function of? We could have k
= k(n,an,z); however, it would introduce many problems if k
depended on z. Amazingly, there is a choice of k(n,an) that works
for all sequences simultaneously, namely k(n,an) = n. This,
however, is quite wasteful. If our series an is such that the sum
of 1/anr converges for some r, we could actually take k
bounded (any integer k exceeding r would work). It is not surprising that
there is interest in finding the simplest choices for k(n) possible. If our
zeros an have some growth, we can bound k.
This happens if our
function is of finite order. See also some
lecture notes
of Noam Elkies from Harvard. Fortunately many important functions in
number theory, such as the
Riemann zeta
function, are of finite order;
the zeta function has a very simple Hadamard product.
- We keep using the wonderful fact that an
entire function which is never zero is the exponential of an entire function.
A fascinating question is just how many values a function can miss and still
be entire. The little
Picard theorem says just one!
- In our study of zeros and accumulation points
we talked a bit about
countable and
uncountable sets; we'll need to revisit these subjects again later when we
prove the
Riemann Mapping Theorem. If you would like to see more about these, let me
know and I'll send you Chapter 5 of
my number theory book.
An interesting fact is that if you have an uncountable sum of non-negative
terms, then it must be zero. The proof is simple: break the summands into
those in intervals (1/2n+1, 1/2n]. At least one of these
intervals has infinitely many summands (else you would have countably many
terms).
- We then moved into some topology, talking
about simply
connected regions. It was essential in defining the
complex logarithm that
we take a simply connected set, as a key ingredient was the Cauchy integral
formula, which requires our function to be holomorphic in the interior of the
curve. It's interesting to consider what logarithms we can define. This leads
to branch cuts
- As always, a great topic to study is the
Jordan Plane Curve
Theorem.
- One of the key insights in our study of
representing functions as products over their zeros was to write (1-z/an)
Ek(z/an) as exp(log(1-z/an) + (z/an)
+ ... + (z/an)k/k). Related to this, consider the
exponential of
a matrix. We set exp(A) = I + A + A2/2! + A3/3! +
.... It can be shown that this converges for any matrix. A natural question is
does exp(A) exp(B) equal exp(A+B). Amazingly, no (though it does if A and B
commute, which a little inspection shows is not unreasonable). The formula for
what exp(A) exp(B) equals is the wonderful
Baker
- Campbell - Hausdorf formula.
- Thursday Oct 14:
- It's very useful to write a function as a
product of its zeros, though obviously this cannot tell the entire story. For
example, we can always multiply by an arbitrary constant and get a new
function with the same zeros. For finite polynomials, we have formulas
relating the coefficients of the polynomial to the roots. Unfortunately, the
other direction is much harder; namely, while it is easy to get the
coefficients of the polynomial from the roots, it is hard to go the other way
and find the roots from the polynomial's coefficients in general (once the
degree is at least five). We do have beautiful formulas such as
Newton's identities.
- The
Weierstrass
Factorization Theorem allows us to write many meromorphic functions as
products over their zeros, so long as we know its value at one place where it
does not vanish. A very nice application of the infinite product for sin(πz)/π
are expressions for Sum_{n = 1 to ∞} 1 / n2k. Particularly useful
ones are k=2 (which gives π2/6) and k=4 (which gives π4/90).
These are also ζ(2) and ζ(4), where ζ(s) = Sum_{n = 1 to ∞} 1 / ns
is the Riemann
zeta function (which converges absolutely for Re(s) > 1). This is one of
the most important functions in number theory, due to the fact that it also
equals Prod_{p prime} (1 - 1/ps)-1 (this follows from
unique factorization of integers). Interestingly, the fact that the sum of
the reciprocal of the squares equals π2/6 implies that there are
infinitely many primes (if there were only finitely many primes, taking s=2 in
the product formula implies that ζ(2) is rational)! This is due to the fact
that π2 is irrational (which follows from the transcendence of π,
but can also be proved simply on its own). There are many proofs of the
irrationality and
transcendence of π.
- The above argument for the infinitude of
primes is known as a
special
value proof, and occurs frequently in the higher arithmetic (ie, number
theory). There are lots of proofs of the infinitude of primes, going all the
way back to Euclid. There's lots of clever ones; see the
wikipedia page for more information (one of my favorites is
Furstenberg's topological proof; I was fortunate enough to meet him when I
was a postdoc at Ohio State). A good source for great proofs such as this (as
well as proofs on other subjects) is the book "Proofs
from THE Book".
- We know that the
Riemann zeta
function at 2k is a rational multiple of π2k; the actual value
involves the
Bernoulli numbers. We know very little about the zeta function at odd
positive integers.
Apery succeeded in proving ζ(3) is irrational (see
also this nice article), but we don't know too much more. We have theorems
that say "at least so many of the zeta function at these odd integers are
irrational", but that's about it.
- Returning to proving the infinitude of primes
(an important question, as the primes are the building blocks of numbers we
should know how many there are!), there's still a lot of `fun' exploring that
you can do with
Euclid's proof (if there were only finitely many primes, say p1, ..., pn,
then consider p1 * ... * pn + 1; either it is prime or it is divisible by a
prime not in our list as each prime in our list leaves remainder 1 when we
divide). A fascinating question is to go through Euclid's proof and write down
what primes we get at each stage.
This leads to an interesting sequence: 2,
3, 7, 43, 13, 53, 5, 6221671, 38709183810571, 139, 2801, 11, 17, 5471,
52662739, 23003, 30693651606209, 37, 1741, 1313797957, 887, 71, 7127, 109, 23,
97, 159227, 643679794963466223081509857, 103, 1079990819, 9539, 3143065813,
29, 3847, 89, 19, 577, 223, 139703, 457, 9649, 61, 4357....
Explicitly, this sequence is generated as follows. Let a_1 = 2, the first
prime. We apply Euclid's argument and consider 2+1; this is the prime 3 so we
set a_2 = 3. We apply Euclid's argument and now have 2*3+1 = 7, which is
prime, and set a_3 = 7. We apply Euclid's argument again and have 2*3*7+1 =
43, which is prime and set a_4 = 43. Now things get interesting: we apply
Euclid's argument and obtain 2*3*7*43 + 1 = 1807 = 13*139, and set a_5 = 13.
Thus a_n is the smallest prime not on our list generated by Euclid's argument
at the nth stage. There are a plethora of (I believe) unknown questions about
this sequence, the biggest of course being whether or not it contains every
prime. This is a great sequence to think about, but it is a computational
nightmare to enumerate! I downloaded these terms from the Online Encyclopedia
of Integer Sequences (homepage is http://www.research.att.com/~njas/sequences/
and the page for our sequence is http://www.research.att.com/~njas/sequences/A000945 ).
You can enter the first few terms of an integer sequence, and it will list
whatever sequences it knows that start this way, provide history, generating
functions, connections to parts of mathematics, ....
This is a GREAT website to know if you want to continue in mathematics.
There have been several times I've computed the first few terms of a problem,
looked up what the future terms could be (and thus had a formula to start the
induction). One last comment: we also talked about the infinitude of primes
from zeta(2) = pi^2/6. While at first this doesn't seem to say anything about
how rapidly the number of primes grows, one can isolate a growth rate from
knowing how well pi^2 can be approximated by rationals (see http://arxiv.org/PS_cache/arxiv/pdf/0709/0709.2184v3.pdf
for details; unfortunately the growth rate is quite weak, and the only way I
know to prove the needed results on how well pi^2 is approximable by rationals
involves knowing the Prime Number Theorem!).
- There are many important generalizations of
Weierstrass' theorem, including
Hadamard's factorization theorem. These representations are very useful in
studying questions in number theory. One of my favorite applications is work
by several colleagues of mine (Gonek, Hughes and Keating:
A hybrid
Euler-Hadamard product formula for the Riemann zeta function). There they
combine the two different product expansions of the zeta function to great
effect.
- I'm currently working with a team of
mathematicians and engineers on applications of
Benford's law to
detecting image fraud (I'm also editing the first book on the subject -- if
you're interested in participating on that project, let me know). The reason I
became involved is the need to understand the size of certain infinite
products arising from Fourier analysis.
The quick analysis I did a few years
back is available here; if you want to improve my bounds and earn a
mention, let me know!
- Finally, we spent a bit of time today trying
to figure out what `good' should be in studying the error in the Taylor
expansion of log(1+z). Learning how to argue along these lines is one of the
most important skills you can develop. The actual nature of `good' (boy can I
see a philosophy class having fun with this!) will clearly depend on your
problem. You need to take some time and think what the key features are, how
your quantities should depend on the parameters of the problem, et cetera.
- Okay, one more tidbit. I'm a big fan of
Ohio State football (so is Cam -- he loves
Brutus).
I've done a lot of sabermetrics (applying math to baseball) projects with
students at Williams. I also can't stand it when people quantify things
poorly and then insist that the quantification is significant because, after
all, there's a number!
There's a nice article today about how bad the BCS rankings are; it's worth
reading (and very well written).
- There are numerous approaches to trying to
rank college football teams, many using graph theory. It's a very hard subject
as there are so few data points. (1)
One nice
article is here. (2)
Here's another
article. (3)
Here's
another. You get the point -- very active area of research, and the BCS
system forbids researchers from using certain key information in their
rankings.
- What's the solution? So glad you asked. One
idea is a playoff series. If you don't want to do that (too much time away
from academics), here's a simple solution. All teams who want to compete and
are considered legitimate contenders put their names into the hat and agree to
a good non-conference game (or perhaps two) with another contender. Some
teams, like Ohio State, do this already. Other teams (Penn State, I think, has
often gone the safe route) try to win their conference and advance through
that (though they did agree to play Alabama this year). Sadly, other teams
(Boise State) are willing to play out of conference but other teams won't play
them. Such a system will allow us to compare more easily the relative
strengths of each conference, and cut down on the number of teams with good
records.
Is there really a need to have a contender play Appalachia State?
- Thursday Oct 7:
-
The argument
principle shows that the integral of the logarithmic derivative of a nice
function along a closed curve is equal to the number of zeros minus the number
of poles of the function in the interior. This is a key ingredient in the
proof of three big results (discussed below). Instead of integrating f'(z)/f(z)
against 1, we could integrate it against a test function g(z). This leads to
what is known as
explicit formulas; these weighted versions of the argument principle
appear throughout number theory (see
for instance the proof sketch on Wikipedia of the Prime Number Theorem).
We will discuss that argument at length later in the course.
-
Rouche's theorem
is the first of many consequences of the argument principle; in fact, the
remaining big theorems are consequences of Rouche's theorem. Rouche's theorem
can be used to prove the
Fundamental Theorem of Algebra. From the Wikipedia article: One popular,
informal way to summarize this argument is as follows: If a person were to
walk a dog on a leash around and around a tree, and if the length of the leash
is less than the minimum radius of the walk, then the person and the dog go
around the tree an equal number of times.
- The
Open Mapping Theorem is the next consequence. Again, the complex situation
differs enormously from the real case. The proof in the book follows from
Rouche's theorem; it is also possible to prove this by analyzing the power
series expansion of our holomorphic function. We didn't prove the Open Mapping
Theorem, but we'll give a proof later using Rouche's Theorem. Is it possible
to prove it another way? As soon as you're told you have a holomorphic
function, you should think analytic, you should think power series. Of course,
the complex numbers must come into play. The function f(x) = x^2 maps (-1,1)
to [0,1), and thus is NOT an open map. Hmm. So what's so special about the
complex numbers...?
- The
Maximum
Modulus Principle is the biggest implication of the Open Mapping Theorem.
This states that a holomorphic function attains its maximum on the boundary.
Applications of this include the
Schwarz lemma
(which is a key ingredient in proving the
Riemann Mapping
Theorem, which allows us to reduce the study of simply connected open sets
not all of the complex plane to studying the unit disk), and the
Phragmen - Lindelof Principle, which is very useful in bounding quantities
in number theory. One such application is in proving convexity bounds for the
Riemann zeta function (and more generally L-functions); see for instance
Heath-Brown's note.
It's a very good exercise to work through some similar examples for real
valued functions and see what goes wrong. Specifically, look at f(x) = 1-x^2
on (-1,1). Note f((-1,1)) = (0,1]. Note the maximum is attained in the
interior, not on the boundary.
- We talked about including all details in a
proof. A great example is when you can
construct a
regular n-gon with straight edge and compass. Gauss completely resolved
this question (well, okay, he reduced it to a determination of which primes
are Fermat primes).
Johann Hermes
has done the
65537-gon (the link is to a website where you can dowload the Mathematica
file, as opposed to traveling to his university to view the 200 pages).
- Another issue we discussed was the danger of
infinities in calculations. The
Fubini-Tonelli Theorem
discusses when we may interchange sums / integrals; a key stumbling block is
when the integral of the absolute value is infinite. One way this can happen
is to have a bounded function on an infinite domain. See the comments from
Sept 14 for more details.
- Tuesday Oct 5:
-
Today we sketched the proof of the
Residue Theorem and
briefly discussed three applications: Rouche's Theorem, the Open Mapping
Theorem, and the Maximum Modulus Principle.
- The
Residue Theorem is
an incredibly powerful tool. Even if you only care about integrals of
functions of a real variable, it is frequently useful to extend to the complex
plane. The reason is that, in general, it is not possible to write down
anti-derivatives; integration is hard! (There is an interesting
algorithm (due to Risch) to find anti-derivatives involving elementary
functions.
The
linked article has a nice example here, where changing the constant term
by 1 leads to the method failing; this is related to a change in the
Galois group.) There
are several steps to using the Residue Theorem:
- Step one: determine the function. Frequently
it is easy: given f(x) try f(z). Sometimes, though, it's a bit harder. If you
have a cos(x) term you could try cos(z) = (exp(iz) + exp(-iz))/2, or you might
try taking just exp(iz) and taking the real part.
- Step two and three are related: choose a
contour and find the poles and residues. Often the location of the poles
affects what contour you take. You DO NOT
want a pole on the contour (I've had to do this a few times in my
research, and it is not fun). Sometimes you have to split the integrand up
into different pieces, and do one part with one closed curve and another part
with another. A big factor in determining contours is how the function decays.
Remember that decay is a bit trickier in complex numbers; for instance, let's
assume |z| > 2000; then 1/|1+z2| <= 1/(R2 - 1); if we
restricted to |x| > 2000 then 1/(1+x2) <= 1/(R2 + 1).
The issue is that we have a phase.
- Step four: repeat earlier steps as needed.
For example, the HW problem (Chapter 3, #1) involved integrating exp(-z2)
over a specific contour. There was a lot of difficulty in bounding the
contribution over the arc. We solved this by splitting the arc into two
pieces; however, it was not clear initially where we should make the split. We
had a good reason for wanting the split to occur as close to π/4 as possible,
and another reason to want it further from there! The best way to figure out
where to put the break point is to introduce a free parameter. In this
problem, we make a split at 1/Ra for some a. We then analyze the
first piece and see what a we need to take to get a contribution tending to 0
as R tends to infinity. We then analyze the second arc and see what a we need
to take for that contribution to tend to 0, and then hope that there is a
choice of a that works for both simultaneously. It is very hard to just look
at this problems (without lots of experience) and know exactly where to make
the cut. Thus, it helps to have a free parameter. This gives you flexibility.
For some problems, you often want to equalize the contribution from both
pieces -- in some sense, this gives you the smallest possible error (or at
least the smallest attainable by this method). For us, that would mean
essentially solving something like R5(a-1) = exp(R2-a).
The actual `best' answer will involve a as a function of R.
- We did a very important example, finding the
normalization constant for integrating 1/(1+x2) over the real line.
This leads to the
Cauchy distribution, which is very important in probability.
- We ended by finding the poles and residues of 1/1+z2010. Key is
Euler's formula:
exp(ix) = cos(x) + i sin(x). Remember that if we want to solve z2010
= -1 = exp(iπ), we could also write -1 as exp(iπ + 2πin) for any integer n.
There will be 2010 distinct solutions (half in the upper half plane, half in
the lower half plane).
- We didn't talk too much about the applications (Rouche's Theorem the Open
Mapping Theorem and the Maximum Modulus Principle), but we will. The last two
are (yet again) instances where complex analysis is quite different than real
analysis.
- Thursday Sept 30:
-
The generalization of
Cauchy's
Integral Formula, the
Residue Theorem, is phenomenal, and allows us to evaluate many difficult
integrals with ease by converting them to algebra problems. The
wikipedia page on contour integration has a lot of great examples. There
are two main steps to applying Cauchy's Integral Formula. The first is
choosing the contours (which is influenced by the decay properties of the
function) and the second is computing the
residues.
- Computing residues: We've seen that using the
geometric series is a great way to compute residues. For example, if we want a
residue at say 3 we replace z everywhere with z-3 + 3; the first factor z-3 is
then small. Another useful approach is through differentiation. Say we have
f(z) = g(z) / (z-3)10 with g(z) holomorphic at z=3. To calculate
the residue at z=3 we need to find the (z-3)9 term of g(z). We
could do this with our trick, or we could compute 9 derivatives.
- Figuring out which paths to take to exploit
the decay is very difficult, and becomes easier with practice. It helps to
know growth rates of different functions. For this <<< notation is very
useful. We write f(x) <<< g(x) if |f(x)/g(x)| tends to 0 as x tends to
infinity. A very useful relation is log(x) <<< xr <<< exp(x) for
any r.
- There are two
Picard theorems,
Big Picard
and
Little Picard. Yet again, we see complex functions have a wildly different
behavior than real valued ones. We find that a non-constant entire function
can miss at most one complex value; thus if we have an entire function that
misses two complex numbers it must be constant! This theorem is sharp, as f(z)
= exp(z) shows.
- The
Riemann Mapping
Theorem is another amazing fact of complex analysis, yet again showing how
different the subject is from real variables (see the nice section in the
Wikipedia article on why this result is so amazing). The analytic isomorphism
between any simply connected open set not all of the complex plane and the
unit disk can sometimes be given explicitly; see the
Schwarz-Christoffel formula.
- Tuesday Sept 28:
- It's amazing how the
Fundamental Theorem of Algebra falls from
Liouville's Theorem. I strongly urge you to look at the list of
Fundamental
Theorems, and read a few that you haven't heard of to get a sense of
what's out there in mathematics. There are a multitude of applications to the
Fundamental Theorem of Calculus. One of my favorites is
Partial
Fraction Decomposition. This is extremely useful in determining integrals
of reciprocals of polynomials (and explains why the logarithm often arises).
It also makes
generating functions. These lead to lots of
wonderful
results about Fibonacci numbers (and other sequences).
- There are formulas for
quadratic equations,
cubic equations (what's amazing is that a real polynomial with real roots
can require you to work with imaginary numbers to get those roots!), and
quartic equations.
There is no general formula for the quintic and higher.
This was first
proved by Abel, and then in a
more general setting by
Galois. Lagrange had a wonderful approach to lower order polynomials (quartics
and smaller) using resolvents; unfortunately this method breaks down for
quintics (it's
good to read the article to see why).
- It is very hard in general to write down
roots of a polynomial in explicit form. This is but one of many questions of
the following nature: given a fixed tool set, what can I do? The ancient
Greeks asked this about a
straightedge and compass. Three famous problems the Greeks tried to do,
but couldn't, were
squaring the circle,
doubling the cube,
and trisecting an
arbitrary angle. All are impossible; attacks on these inspired much of
algebra. Another problem was which regular n-gons can be constructed; this was
essentially solved by Gauss, and is one of the reasons he chose to become a
mathematician. He showed you cannot do a 7-gon, 9-gon, 11-gon, 13-gon, but can
do a 17-gon (see also
the Mathworld
article on constructable polygons).
- While we cannot write down closed form
expressions for the roots in general, we have algorithms such as
Newton's method
which often do a great job of approximating them. The convergence (or
divergence!) of this method leads to fascinating behavior and questions; see
the wikipedia article on
Newton fractals, for example.
- As we've already mentioned Liouville, let's
do another of his theorems. He was the first to write down a number (these are
the Liouville numbers)
that is provably
transcendental (namely, not the solution of a finite polynomial with
integer coefficients). Cantor proved (almost 40 years later) that almost all
real numbers are transcendental, though his
diagonalization
method cannot give specific examples. Determining which numbers in certain
expressions are transcendental is an important area (it's
Hilbert's
seventh problem); see the
Gelfand -
Schneider Theorem for some results. Properties of Liouville and other
numbers are often related to arithmetic dynamics; in fact, how well log(2) /
log(10) can be approximated by rationals is a key ingredient in my proof (with
Alex Kontorovich) of
Benford behavior in the 3x+1
problem (the
paper is available here).
-
Liouville's theorem is (yet another) application of Cauchy's formula. It
completely classifies all functions that are both holomorphic and entire.
There are other situations like this in mathematics. One important one is
Harald Bohr
(brother of Niels Bohr) and Johannes Mollerup proved that the only
function which satisfies the following three properties is the
Gamma function: (1)
f(1) = 1; (2) f(x+1) = f(x) for x > 0; (3) f is logarithmically convex.
- Another one of my favorite problems is the
Hamburger
Moment Problem (and not just because of the name). This is related to when
a sequence of
moments uniquely determines a probability density (if p(x) is a
probability density, the k-th moment is the integral of xk p(x) dx).
Note how similar this is to the Bohr-Mollerup theorem. In fact, this is
related to our accumulation point results, and is one of the most important
reasons for studying teh subject. Namely, when does a sequence of moments
uniquely determine a probability distribution. There are two
probability densities that are different but have the same moments. The
problem is solved by considering non-integral moments as well, and if there is
agreement along a sequence with an accumulation point, then the densities are
equal. For full details, see my notes
from probability. To rigorously prove many results in probability, such as
the Central Limit
Theorem, requires many results from complex analysis.
- Thursday Sept 23:
- We finished the contour example, namely integrating (1-cos(x))/x2
over the real line. We had to be careful as the complexification (1 - exp(iz))/z2
has a pole at z=1, and thus there would be a singularity on the real line. We
literally circumvent this by detouring, going along a semi-circle of radius
ε about the origin. Fortunately, it turns out
that most integrals in complex analysis can be reduced to integrals over
circles and semi-circles. Remember when choosing your contour, by Cauchy's
formula many will give the same answer, but through different algebra. Make
your life as easy as possible. This problem as a 1/x2 or a 1/z2
in the denominator. It's thus natural to use a large semi-circle of radius R,
as there the denominator is exactly R2 in absolute value; if we did
a rectangle or a triangle, the algebra would be a bit more unpleasant.
- Cauchy's formula is a true gem of
mathematics, and quite surprising at first. We only need to know a holomorphic
function on a circle in order to determine its value everywhere in the
interior! This is an example of a
boundary value
problem; the
wikipedia article linked here enumerates many of the more famous of these.
Note how different the situation is in real analysis, where it is unusual for
a function to be uniquely determined by its values on a boundary curve. There
are, though, situations where that happens; see for instance
harmonic functions
(which occur frequently in mathematical physics). The miracle here is that
once you specify a holomorphic function on the boundary, you have no
choice in how it is defined elsewhere!
- A big step in our analysis was interchanging
a derivative and an integral. The book proves it directly for this special
case through series expansions and radii of convergence; however, it's good to
be aware of when (in general) this can be done;
see for example here.
- Multiple time today we added zero in a clever
way; this is one of the most important skills to master. The first place you
might remember seeing this is in Calc I (in proving say the product rule),
though if you think back you've seen it even earlier in your career, namely in
completing the
square to prove the quadratic formula. One of course wants to get a feel
for when this should be done. For the proof of Cauchy's integral formula, we
have f(ζ) / (ζ-z). As ζ is close to z, it makes sense to add -f(z) + f(z) to
the numerator. This gives (f(ζ) - f(z)) / (ζ - z) + f(z) / (ζ - z). The first
term tends to f'(z), which by assumption is understood. We actually don't need
the full strength of f being holomorphic; all we really need is the quotient (f(ζ)
- f(z)) / (ζ - z) is bounded (which then will improve itself to f is
holomorphic). If f is
Lipschitz, that
suffices; Holder
continuous does not if the exponent is less than 1.
- We also saw adding zero when proving a
holomorphic function is analytic. To Taylor expand a function about a point
z0, we want to have a series Σ an (z - z0)n
where the an's are f(n)(z0)/n!. Using Cauchy's formula,
we can relate these derivatives to integrating along a circle centered at z0
the function f(ζ) / (ζ-z0). Unfortunately, we start out knowing the
integral expansion for f(z), which involves integrating f(ζ) / (ζ-z) about the
same circle. Essentially, we want to replace the z with a z0. We do
this by writing ζ-z = ζ-z0 + z0 - z, which is just
(ζ-z0) - (z - z0). We now see how well-suited this is to
our analysis. We expand with a geometric series (which has nice convergence
properties), and are now in the position where we'll be integrating against
ζ-z0. The reason we have good convergence is that we have a
geometric series with ratio |z - z0| / |ζ-z0| < 1.
- Finally, the great π-debates continue, namely
which is the best: π, 2π, 4π, or 2πi (or maybe even 4π/3)? Each has its
proponents. Is the fundamental object the radius of a circle, or its diameter,
or its perimeter, or its area, or the surface area of a sphere, or the volume
of a sphere.... For example, look and see how the constant is written in
Coulomb's law (ie,
electricity). Or, even better, in the
fine structure
constant, one of the most important numbers in mathematical physics
(because it is unitless!).
- Tuesday Sept 21:
- Our first main result is that if a function is holomorphic in a disk then
it has a primitive (in other words, an anti-derivative). The big consequence
of this is that integrals of our function over closed curves are zero, which
implies that the integral along any path depends only on the endpoints, and
not the curve taken to connect them. Situations such as this occur all the
time in physics; you might remember
conservative forces
such as gravity and electricity.
A conservative force
is the gradient of a scalar potential.
- It is impossible to do path integrals justice in just a few words in
class.
Feynman used path integral formulations to formulate quantum mechanics!
Richard Feynman was
an extremely colorful character, and the author of numerous great books,
including:
- Speaking of good books to read by famous
mathematicians / physicists, another classic is Hardy's
A
Mathematician's Apology. Hardy was one of the top analysts of his time,
making seminal contributions to many fields (including applications of complex
analysis in number theory). If you are interested in any of these books, I
have them in my office.
- There are lots of ways to do the calculation
to show a primitive exists. Instead of using the polygonal path we did one
could take the direct line segment. Why did we proceed as in the book? The
reason is that there are many times in math that polygonal approximations are
useful, and it's often very convenient to move only parallel to the coordinate
axes. You might have seen something along these lines (if you'll forgive the
pun) when proving Green's theorem. One has to be careful with polygonal
approximations, though, as if you are careless you can prove all curves
connecting two points have the same length! To see this, take the curve y = x2
from x=0 to x=1. This function is strictly increasing, and we can
approximate it by very small vertical and horizontal lines. We can make the
approximation arbitrarily close, say always withing epsilon pointwise. We
would like to then say the length of the curve is approximately the sum of the
vertical and horizontal segments; however, all the vertical segments add to
the interval [0,1], as do the horizontal segments. Thus this curve has length
2 (as would any other similar curve). What is going horribly wrong? If ds is
the infinitesimal distance differential along the curve, the correct formula
is ds = sqrt(dx2 + dy2), whereas we were using a moment
ago ds = dx + dy. It is in general a hard question to find lengths of curves (ie,
the arc length).
-
The Joran plane
curve theorem tells us when a curve divides the plane into two regions, an
inside and an outside.
- We showed that the integral of a holomorphic
function along a closed curve is zero. There is a converse, namely
Morera's theorem:
if a continuous function integrates to zero along every closed curve then it
must be constant.
- A big part of complex analysis is learning
how to choose the function that generalizes the real integrand properly, as
well as choosing the right contours. For many examples of doable contour
integrals, see the
wikipedia page here (some of these won't make complete sense until we've
done parts of Chapter 3). The second example from Section 3 is terrific, and
illustrates many of the key ideas. We can write sines and cosines in terms of
exponentials, and thus replace these real valued functions with complex valued
ones. The difficulty is that we must make sure we have rapid decay on some of
the sides of integration. Writing cos(z) as (exp(iz) + exp(-iz))/2 is not a
bad thing to try, but for z=x+iy, if y>0 then exp(-iz) is rapidly growing in
y. We thus need to be careful. One solution is to write 1 - (exp(iz) + exp(-iz))/2
as 1/2 - exp(iz)/2 + 1/2 - exp(-iz)/2, and use contours in the upper half
plane for the first and bottom half plane for the second. The other approach
is to write 1 - cos(z) as the real part of 1 - exp(iz), and then we can use
just one contour in the upper half plane.
- Thursday Sept 16:
- Stokes' Theorem (or, for us, Green's Theorem as we're in the plane) is a
gem of mathematics. We saw how it can be used to compute areas of regions by
reducing it to a line integral. The example we did in class today was the area
of the ellipse;
the
special case when it's a circle of radius 2 is done here (it's a shame
they do the special case!). Whenever you get an answer in math (in this case,
the area of the ellipse is πab), you should check
and make sure it agrees with other known, simpler cases. In this instance we
check with the case a = b = r, and recover the area of the disk. In general,
it is much easier to compute areas than lengths; the
length of the
ellipse is quite difficult to compute, and involves the
elliptic integral of the second kind, which is a special case of the
Gauss
hypergeometric function. Hypergeometric functions are a very rich family,
and arise throughout mathematical physics.
- The
Jordan Plane Curve
Theorem is one of the candidates for greatest disparity between expected
ease of proof and actual difficulty of proof. The
history and further proofs section is a fun read. One of the proos uses
the Brouwer
fixed point theorem. Fixed point theorems arise all the time, especially
in game theory in economics (see
for instance the work of Nash).
- Using polar coordinates to compute the
integral of exp(-x2/2) is a wonderful technique, but if feels a bit
artificial. There are other ways. One of my favorites is to change variables
and observe that we have a special value of the
Gamma function
(which generalizes the factorial function) and then use some identities of
this function. While we can evaluate the integral of this function over the
entire real line, there is no nice expression (involving elementary functions)
for its anti-derivative.
There is an interesting
algorithm (due to Risch) to find anti-derivatives involving elementary
functions.
The
linked article has a nice example here, where changing the constant term
by 1 leads to the method failing (this is related to a change in the
Galois group).
- We discussed a variety of applications of integration. As a number
theorist, one of my favorites is in applications to problems such as
Goldbach's problem
or Waring's problem.
One of the most powerful ways to attack these is the through the
Circle Method. If you
are interested in learning more about this, let me know and I'll print out the
chapter from my book on this. The gist of these ideas is that we have a
generating
function whose coefficients encode the answer to our problem, and we must
somehow pick off these coefficients.
- It's frequently convenient to work with triangles, as these fit nicely for
triangulation purposes (see also the article on
triangularizing
polygons). One can prove
Goursat's
theorem (the link here includes the Stokes proof). One very nice
consequence is
Morera's theorem. Another place where you may have seen triangles is in
the proof of the
Euler characteristic of a surface in topology.
- We mentioned CPCTC today,
as we needed to show each triangle was similar to the original. There are lots
of other useful acronyms (SAS,
SSS, ASA and AAS). Speaking of geometry, one of my favorite results is
Morley's
theorem. I hold the record (I believe) for the longest proof (about 40 -
50 pages, unpublished, see me for the story).
John Conway
has one of the most beautiful proofs of this fact. If you enjoy geometric
arguments like this, you might enjoy some of the
non-standard proofs of the irrationality of certain square-roots.
- There
was a nice post on the math arxiv today giving a continuous but nowhere
differentiable function.
- Finally, it's worth reiterating `checking'
formulas you can't remember. If you can remember some special cases, you can
often correct your guess. We saw how to find where the minus sign is in the
Cauchy-Riemann
equations by testing our guess with the functions f(z) = z or z2.
For another example, if you only remember some of the quadratic formula you
can try f(x) = x2 - 1 or x2 - x or x2 - 3x +
2, all of which lead to readily computable roots.
- Tuesday Sept 14:
- The main result today is interchanging a derivative and an infinite sum.
Interchanging operations is one of the most important techniques, and
sometimes one of the most delicate. The first instance you encounter is
probably Fubini's
Theorem. Note it is not always permissible to exchange integrals or sums;
the difficulty is when infinities are involved. The standard condition is if
the integral (or sum) of the absolute value is finite then you are okay. For a
problem case, consider the double sequence (m, n >= 0) given by a_{m,n} = 0
for n < m or n > m + 1, a_{m,m} = 1 and a_{m,m+1} = -1. Another important case
is
differentiating under the integral sign; I strongly urge you to look at
the examples here of the integrals this allows you to evaluate.
Feynman was a master at
this technique. Today we discussed differentiating under the summation sign,
which is possible if the series converges absolutely in a disk. The key
concepts were the
radius of convergence, which involves the notion of the
limit superior (or the
limsup). Lurking beneath all these proofs is a comparison test with a
geometric series.
- There are many nice ways to prove the geometric series formula. My
favorite involves a game of hoops with two players. A always gets a basket
with probability p, B with probability q (assume p, q < 1), and first to get a
basket wins. Let r = (1-p)(1-q) and x the probability A wins. Then x = p + rp
+ r2 p + r3 p + ...; these add the probabilities of A
winning on the first, second, third, ... attempt (as to get to the second
attempt both A and B must miss). I claim that x = p + rx, as after both miss
the probability A wins from this point onward is just x again! This gives x =
p / (1-r), or 1 + r + r2 + r3 + ... = 1/(1-r), the
geometric series formula! Using the memorylessness nature s a key ingredient
in many problems in economics. For more, see the
article on
martingales.
-
Today was a fast introduction to path
integrals, line integrals, and Green's
Theorem (which is a special
case of the Generalized Stokes'
Theorem). If you are continuing in certain parts of math, physics or
engineering you will meet these again and again (for example, see Maxwell
equations for electricity and magnetism). In fact, one can view all of
classical mechanics as path
integrals where the trajectory of the particle (its c(t)) minimizes the action;
there is also a path
integral approach to quantum mechanics.
- One of the gems of complex analysis is Cauchy's
Integral Theorem, A complex differentiable function satisfies what is
called the Cauchy-Riemann
equations, and these are essentially the combination of partial
derivatives one sees in Green's theorem. In other words, the mathematics
used for Green's theorem is crucial in understanding functions of a complex
variable.
- For me, I consider it one of the most beautiful gems in mathematics that
we can in some sense move the derivative of the function we're integrating
to act on the region of integration! This allows us to exchange a double
integral for a single integral for Green's theorem (or a triple integral for
a double integral in the divergence theorem). As we've seen constantly
throughout the year, often one computation is easier than another, and thus
many difficult area or volume integrals are reduced to simpler, lower
dimensional integrals. This subject is done properly in
differential forms.
- The fact that Int_{t = a to b} grad(f)(c(t)) . c'(t) dt = f(c(b)) -
f(c(a)) means that this integral does not depend on the path. If a vector
field F = (F1, F2, F3) equals grad(f) for some f, we say F is a conservative
force field and f is the potential.
The fact that these integrals do not depend on the path has, as you would
expect, profound applications.
- This is a good point to stop and think about the number of spatial
dimensions in the universe. Imagine a universe with two point masses under
gravity, and assume gravity is proportional to 1/r^{n-1} with r the distance
between the masses and n the number of spatial dimensions. If there are
three or more dimensions, then the work done in moving a particle from
infinity to a fixed, non-zero distance from the other mass is finite, while
if there are two dimensions the work is infinite! One should of course ask
why the correct generalization to other dimensions is 1/r^{n-1} and not
1/r^2 always. There is a nice geometric justification in terms of flux and
surface area; the surface area of a sphere grows like r^2 and thus the only
way to have the total flux of force out of it be constant is to assume the
force drops like 1/r^2; click
here for a bit on the justification of inverse-square laws.
- Speaking of dimensions, one of my favorite problems from undergraduate
days was the Random
Walk. In 1-dimension, imagine a person so completely drunk that he/she
has a 50% chance at any moment of stepping to the left or the right; what is
the probability the drunkard eventually returns home? It turns out that this
happens with probability 1. In 2-dimension, we have a 25% chance of moving
north, south, east or west, and again the probability of returning is 1. In
3 dimensions, however, the drunkard only returns home with probability about
34%. As my professor Peter
Jones said, a
three-dimensional universe is the smallest one that could be created that
will be interesting for drunks, as they really get to explore! These random
walk models are very important, and have been applied to economics (the
random walk hypothesis), as well as playing a role in statistical
mechanics in physics.
- We sketched the proof of Green's theorem by saying it suffices to prove
it for a rectangle. It might be better to do a triangle as these are a bit
easier for
triangularizing a region (as the name implies!). Of course, If you can
do any rectangle you can do any triangle by using appropriately many
rectangles.
- In real analysis, one develops the theory of
Riemann sums to
rigorously investigate integrals. To do the subject properly, one should
consider an arbitrary
partition of
the set. The most commonly used ones are into equal pieces, and into 2^n
equal pieces. If we divide into n equal pieces, the disadvantage is that at
the next step the new subintervals can overlap two previous subintervals; if
we do 2n divisions and then divide again, we have 2n+1 divisions
and now each new subinterval is half of a previous subinterval. Usually we
want to do one of these two partitions, but not always. In solving
differential
equations numerically, say via the
Euler method or the
Runge-Kutta method,
one may not want to take equally spaced points. The reason is that if the
function is not changing much, it is not a good idea to waste a lot of
computational time there; rather, given a finite number of places to evaluate,
you want to spend your resources where the function's behavior is wildest.
This is a vast topic;
see for instance the link here. I did some searching on the web, and
found this thesis, which has a nice introduction.
- In sketching the proof of Green's theorem, we used rectangles and talked
about pushing the line integrals to the boundary. Some care is needed here, as
otherwise we can prove all curves from (0,0) to (1,1) have the same length!
How? Take the polygonal approximation to the given curve; for convenience we
assume the curve is
convex; in other words, it looks like y = x2 or y = x3.
Then if we look at the sum of the horizontal polygonal part we get the
interval [0,1], which is what we get when we look at the sum of the vertical
parts! Thus all curves have the same length, namely 2, which is clearly
absurd. What went wrong? We're approximating the infinitesimal distance ds
with dx + dy and not sqrt(dx^2 + dy^2). This is all related to how much harder
it is in calculus to compute lengths of curves than areas;
see the section on arc
length in wikipedia.
- Thursday Sept 9: (Yes, I know it was
really Wednesday the 8th, but Wednesday = Thursday today!). In today's class
we discussed some of the basics of complex analysis, from the definition and
properties of complex numbers to the definition of the derivative to some
differences between real valued and complex valued functions. The comments
below are meant to delve more deeply into some of these topics. These are
meant for your enjoyment and personal edification; they are entirely optional
and are independent of the course work and your grade. That said, reading
these will help review some of the concepts we've seen.
- Complex numbers:
complex numbers are of the form z = x + iy, with the complex conjugate z =
x - iy with i2 = -1. While it is impossible to
totally order the
complex numbers (unlike the reals, which can), they still have many nice
properties. They are associative and commutative. It is possible to generalize
the complex numbers further. The first is the four dimensional space of the
quaternions, which are
commutative but not associative. The next generalization is the
octonians, which are eight
dimensional (over the reals) and now even associativity is lost! What's next?
The sedenions!
- We showed the definition of the derivative can be rewritten and
interpreted as the first order
Taylor series does an
excellent job approximating the function. Interestingly, a function can have
continuous partial derivatives without being differentiable. The big theorem
is that if the partial derivatives are continuous then the function is
differentiable. It is possible for a function to have a limit along some paths
but not others, or to have different limits along different paths. For a
function to be differentiable, the quotient must tend to the same value no
matter what path you take. In one variable there is essentially just two
choices (approach from the left or the right); it's much more interesting in
several variables, as
Cam and Kayla's artwork show.
- Here is a nice example: it has the same limit along any straight line or
parabola, but a different limit along some cubic paths! Consider f(x,y) = (x8
+ y8)/(x2 + y8) - (x10 + y10)/(x4
+ y10) and take the limit as (x,y) approaches (0,0).
- While the definition of complex differentiable looks innocuously similar
to that for one variable, remember that the limit must hold for any path. This
implies a variety of strange facts. A good way of viewing this is that
functions of a complex variable z can be written as functions of x and y.
Using z = x + iy and zbar = x - iy, we have x = (z+zbar)/2 and y = (z-zbar)/2i.
Essentially the complex differentiable functions are those which depend on z
and not zbar. Thus only special combinations of x and y are allowed. Later
we'll see that complex differentiability is equivalent to our function
satisfying the
Cauchy-Riemann differential equations. One way of interpreting this is
that we have a function whose corresponding function of two real variables x
and y has zero curl.
This means we have restricted ourselves to a very nice subset of real valued
functions, in particular ones where
Green's Theorem (or
Stokes' Theorem) is
particularly easy to apply.
- In comparing results in real analysis to complex analysis, we met some
very interesting functions. One is
sin(1/x); though this
function isn't differentiable,
x2 sin(1/x)
is (but not infinitely often!).
- Another great example is the function f(x) = exp(-1/x2) for x
not zero and 0 otherwise. To take the derivatives requires
L'Hopital's rule.
This function is extremely important in probability, as it shows that a Taylor
series does not uniquely determine a function. In probability this arises as
the moments of
a probability
distribution do not always uniquely determine the probability
distribution. This is covered in great depth in some
notes I wrote up for Math 341 (probability).
- We mentioned that there is a way to define a measure on the space of
continuous function from the reals to the reals, and in this metric almost all
functions are differentiable nowhere!
Weierstrass has
a wonderful example (see also
the article by McCarthy
as well as the
article by
Johnsen).
- We used
dimensional analysis to prove the
Pythagorean Theorem.
There are many proofs; one particularly nice one is due to
James Garfield, a
Williams alum (and president of the US). Victor Hill, an emeritus professor of
mathematics here, has a very
enjoyable article on Garfield and his proof.
-
Recall the exponential
function exp is defined by e^z
= exp(z) = sum_{n = 0 to oo} z^n/n!. This series converges for all z. The
notation suggests that e^z e^w = e^(z+w); this is true, but it needs to be
proved. (What we have is an equality of three infinite sums; the proof uses
the binomial theorem.)
Using the Taylor series expansions for cosine
and sine, we find e^(iθ) = cos θ + i sin θ. From this we find |e^(iθ)| =
1; in fact, we can use these ideas to prove all trigonometric identities! For
example:
- Inputs: e^(iθ) = cos θ + i sin θ and e^(iθ)
e^(iφ) = e^(i (θ+φ))
- Identity: from e^(iθ) e^(iφ) = e^(i (θ+φ))
we get, upon substituting in the first identity, that (cos θ + i sin θ) (cos
φ + i sin φ) = cos(θ+φ) + i sin(θ+φ). Expanding the left hand side gives (cos
θ cos φ - sin θ sin φ) + i (sin θ cos φ + cos θ sin φ) = cos(θ+φ) + i
sin(θ+φ). Equating the real parts and the imaginary parts gives the
identities
- One can prove other identities along these
lines....
- Finally, a common theme in mathematics is the
need to simplify tedious algebra. Frequently we have claims that can be proven
by long and involved computations, but these often leave us without a real
understanding of why the claim is true. If you want, let me know and I'll show
you my 40-50 page proof of Morley's
theorem; Conway
has a beautiful proof which you can read here (it's after the irrationality of
sqrt(2)). If you like non-standard proofs of the irrationality of sqrt(2),
see the article I wrote with a SMALL
student (to appear in Mathematics Magazine).