Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Bonus: Fourier analysis} | |
\label{ch:fourier} | |
Now that we've worked hard to define abstract inner product spaces, | |
I want to give an (optional) application: | |
how to set up Fourier analysis correctly, using this language. | |
For fun, I also prove a form of Arrow's Impossibility Theorem | |
using binary Fourier analysis. | |
In what follows, we let $\TT = \RR/\ZZ$ denote the ``circle group'', | |
thought of as the additive group of ``real numbers modulo $1$''. | |
There is a canonical map $e \colon \TT \to \CC$ sending $\TT$ to the | |
complex unit circle, given by | |
\[ e(\theta) = \exp(2\pi i \theta). \] | |
\section{Synopsis} | |
Suppose we have a domain $Z$ and are interested in functions $f \colon Z \to \CC$. | |
Naturally, the set of such functions form a complex vector space. | |
We like to equip the set of such functions | |
with an positive definite \emph{inner product}. | |
% usually something like $\left< f,g \right> = \EE_{x \in Z} f(x) \ol{g(x)}$. | |
%which makes the set of such functions into a normed vector space. | |
%(Here $\EE$ is an ``average'', which is a finite sum if $|Z| < \infty$ | |
%but otherwise is usually an integral.) | |
%In particular, $\left< f,f\right>$ is the average of $|f(x)|^2$; | |
%and thus gives a positive definite inner form. | |
The idea of Fourier analysis is to then select an \emph{orthonormal basis} | |
for this set of functions, say $(e_\xi)_{\xi}$, | |
which we call the \vocab{characters}; | |
the indexing $\xi$ are called \vocab{frequencies}. | |
In that case, since we have a basis, every function $f : Z \to \CC$ | |
becomes a sum | |
\[ f(x) = \sum_{\xi} \wh f(\xi) e_\xi \] | |
where $\wh f(\xi)$ are complex coefficients of the basis; | |
appropriately we call $\wh f$ the \vocab{Fourier coefficients}. | |
The variable $x \in Z$ is referred to as the \vocab{physical} variable. | |
This is generally good because the characters are deliberately chosen | |
to be nice ``symmetric'' functions, | |
like sine or cosine waves or other periodic functions. | |
Thus we decompose an arbitrarily complicated function into a sum of nice ones. | |
\section{A reminder on Hilbert spaces} | |
For convenience, we record a few facts about orthonormal bases. | |
\begin{proposition} | |
[Facts about orthonormal bases] | |
\label{prop:orthonormal} | |
Let $V$ be a complex Hilbert space | |
with inner form $\left< -,-\right>$ | |
and suppose $x = \sum_\xi a_\xi e_\xi$ and $y = \sum_\xi b_\xi e_\xi$ | |
where $e_\xi$ are an orthonormal basis. | |
Then | |
\begin{align*} | |
\left< x,x \right> &= \sum_\xi |a_\xi|^2 \\ | |
a_\xi &= \left< x, e_\xi \right> \\ | |
\left< x,y \right> &= \sum_\xi a_\xi \ol{b_\xi}. | |
\end{align*} | |
\end{proposition} | |
\begin{exercise} | |
Prove all of these. | |
(You don't need any of the preceding section, | |
it's only there to motivate the notation with lots of scary $\xi$'s.) | |
\end{exercise} | |
In what follows, | |
most of the examples will be of finite-dimensional inner product spaces | |
(which are thus Hilbert spaces), | |
but the example of ``square-integrable functions'' | |
will actually be an infinite dimensional example. | |
Fortunately, as I alluded to earlier, | |
this is no cause for alarm | |
and you can mostly close your eyes and not worry about infinity. | |
\section{Common examples} | |
\subsection{Binary Fourier analysis on $\{\pm1\}^n$} | |
Let $Z = \{\pm 1\}^n$ for some positive integer $n$, | |
so we are considering functions $f(x_1, \dots, x_n)$ accepting binary values. | |
Then the functions $Z \to \CC$ form a $2^n$-dimensional vector space $\CC^Z$, | |
and we endow it with the inner form | |
\[ \left< f,g \right> = \frac{1}{2^n} \sum_{x \in Z} f(x) \ol{g(x)}. \] | |
In particular, | |
\[ \left< f,f \right> | |
= \frac{1}{2^n} \sum_{x \in Z} \left\lvert f(x) \right\rvert^2 \] | |
is the average of the squares; | |
this establishes also that $\left< -,-\right>$ is positive definite. | |
In that case, the \vocab{multilinear polynomials} form a basis of $\CC^Z$, | |
that is the polynomials | |
\[ \chi_S(x_1, \dots, x_n) = \prod_{s \in S} x_s. \] | |
\begin{exercise} | |
Show that they're actually orthonormal under $\left< -,-\right>$. | |
This proves they form a basis, since there are $2^n$ of them. | |
\end{exercise} | |
Thus our frequency set is actually the subsets $S \subseteq \{1, \dots, n\}$. | |
Thus, we have a decomposition | |
\[ f = \sum_{S \subseteq \{1, \dots, n\}} \wh f(S) \chi_S. \] | |
\begin{example} | |
[An example of binary Fourier analysis] | |
Let $n = 2$. | |
Then binary functions $\{ \pm 1\}^2 \to \CC$ have a basis | |
given by the four polynomials | |
\[ 1, \quad x_1, \quad x_2, \quad x_1x_2. \] | |
For example, consider the function $f$ | |
which is $1$ at $(1,1)$ and $0$ elsewhere. | |
Then we can put | |
\[ f(x_1, x_2) = \frac{x_1+1}{2} \cdot \frac{x_2+1}{2} | |
= \frac14 \left( 1 + x_1 + x_2 + x_1x_2 \right). \] | |
So the Fourier coefficients are $\wh f(S) = \frac 14$ | |
for each of the four $S$'s. | |
\end{example} | |
This notion is useful in particular for | |
binary functions $f : \{\pm1\}^n \to \{\pm1\}$; | |
for these functions (and products thereof), | |
we always have $\left< f,f \right> = 1$. | |
It is worth noting that the frequency $\varnothing$ plays a special role: | |
\begin{exercise} | |
Show that | |
\[ \wh f(\varnothing) = \frac{1}{|Z|} \sum_{x \in Z} f(x). \] | |
\end{exercise} | |
\subsection{Fourier analysis on finite groups $Z$} | |
This time, suppose we have a finite abelian group $Z$, | |
and consider functions $Z \to \CC$; | |
this is a $|Z|$-dimensional vector space. | |
The inner product is the same as before: | |
\[ \left< f,g \right> = \frac{1}{|Z|} \sum_{x \in Z} f(x) \ol{g(x)}. \] | |
To proceed, we'll need to be able to multiply two elements of $Z$. | |
This is a bit of a nuisance since it actually won't really | |
matter what map I pick, so I'll move briskly; | |
feel free to skip most or all of the remaining paragraph. | |
\begin{definition} | |
We select a \emph{symmetric non-degenerate bilinear form} | |
\[ \cdot \colon Z \times Z \to \TT \] | |
satisfying the following properties: | |
\begin{itemize} | |
\ii $\xi \cdot (x_1 + x_2) = \xi \cdot x_1 + \xi \cdot x_2$ | |
and $(\xi_1 + \xi_2) \cdot x = \xi_1 \cdot x + \xi_2 \cdot x$ | |
(this is the word ``bilinear'') | |
\ii $\cdot$ is symmetric, | |
\ii For any $\xi \neq 0$, there is an $x$ with $\xi \cdot x \neq 0$ | |
(this is the word ``nondegenerate''). | |
\end{itemize} | |
\end{definition} | |
\begin{example} | |
[The form on $\Zc n$] | |
If $Z = \Zc n$ then $\xi \cdot x = (\xi x)/n$ satisfies the above. | |
\end{example} | |
In general, it turns out finite abelian groups | |
decompose as the sum of cyclic groups (see \Cref{sec:FTFGAG}), | |
which makes it relatively easy to find such a $\cdot$; | |
but as I said the choice won't matter, so let's move on. | |
Now for the fun part: defining the characters. | |
\begin{proposition} | |
[$e_\xi$ are orthonormal] | |
For each $\xi \in Z$ we define the character | |
\[ e_\xi(x) = e(\xi \cdot x). \] | |
The $|Z|$ characters form an orthonormal basis of the | |
space of functions $Z \to \CC$. | |
\end{proposition} | |
\begin{proof} | |
I recommend skipping this one, but it is: | |
\begin{align*} | |
\left< e_{\xi}, e_{\xi'} \right> | |
&= \frac{1}{|Z|} \sum_{x \in Z} e(\xi \cdot x) \ol{e(\xi' \cdot x)} \\ | |
&= \frac{1}{|Z|} \sum_{x \in Z} e(\xi \cdot x) e(-\xi' \cdot x) \\ | |
&= \frac{1}{|Z|} \sum_{x \in Z} e\left( (\xi-\xi') \cdot x \right). | |
\end{align*} | |
\end{proof} | |
In this way, the set of frequencies is also $Z$, | |
but the $\xi \in Z$ play very different roles from the ``physical'' $x \in Z$. | |
Here is an example which might be enlightening. | |
\begin{example} | |
[Cube roots of unity filter] | |
Suppose $Z = \Zc3$, with the inner form given by $\xi \cdot x = (\xi x)/3$. | |
Let $\omega = \exp(\frac 23 \pi i)$ be a primitive cube root of unity. | |
Note that | |
\[ e_\xi(x) = \begin{cases} | |
1 & \xi = 0 \\ | |
\omega^x & \xi = 1 \\ | |
\omega^{2x} & \xi = 2. | |
\end{cases} \] | |
Then given $f \colon Z \to \CC$ with $f(0) = a$, $f(1) = b$, $f(2) = c$, | |
we obtain | |
\[ f(x) = \frac{a+b+c}{3} \cdot 1 | |
+ \frac{a + \omega^2 b + \omega c}{3} \cdot \omega^x | |
+ \frac{a + \omega b + \omega^2 c}{3} \cdot \omega^{2x}. \] | |
In this way we derive that the transforms are | |
\begin{align*} | |
\wh f(0) &= \frac{a+b+c}{3} \\ | |
\wh f(1) &= \frac{a+\omega^2 b+ \omega c}{3} \\ | |
\wh f(2) &= \frac{a+\omega b+\omega^2c}{3}. | |
\end{align*} | |
\end{example} | |
\begin{exercise} | |
Show that in analogy to $\wh f(\varnothing)$ | |
for binary Fourier analysis, we now have | |
\[ \wh f(0) = \frac{1}{|Z|} \sum_{x \in Z} f(x). \] | |
\end{exercise} | |
Olympiad contestants may recognize the previous example | |
as a ``roots of unity filter'', which is exactly the point. | |
For concreteness, suppose one wants to compute | |
\[ \binom{1000}{0} + \binom{1000}{3} + \dots + \binom{1000}{999}. \] | |
In that case, we can consider the function | |
\[ w : \ZZ/3 \to \CC. \] | |
such that $w(0) = 1$ but $w(1) = w(2) = 0$. | |
By abuse of notation we will also think of $w$ | |
as a function $w : \ZZ \surjto \ZZ/3 \to \CC$. | |
Then the sum in question is | |
\begin{align*} | |
\sum_n \binom{1000}{n} w(n) | |
&= \sum_n \binom{1000}{n} \sum_{k=0,1,2} \wh w(k) \omega^{kn} \\ | |
&= \sum_{k=0,1,2} \wh w(k) \sum_n \binom{1000}{n} \omega^{kn} \\ | |
&= \sum_{k=0,1,2} \wh w(k) (1+\omega^k)^n. | |
\end{align*} | |
In our situation, we have $\wh w(0) = \wh w(1) = \wh w(2) = \frac13$, | |
and we have evaluated the desired sum. | |
More generally, we can take any periodic weight $w$ | |
and use Fourier analysis in order to interchange the order of summation. | |
\begin{example} | |
[Binary Fourier analysis] | |
Suppose $Z = \{\pm 1\}^n$, viewed as an abelian group | |
under pointwise multiplication | |
hence isomorphic to $(\ZZ/2\ZZ)^{\oplus n}$. | |
Assume we pick the dot product defined by | |
\[ \left< \xi, x \right> | |
= \half \sum_i \frac{\xi_i-1}{2} \cdot \frac{x_i-1}{2} \] | |
where $\xi = (\xi_1, \dots, \xi_n)$ and $x = (x_1, \dots, x_n)$. | |
We claim this coincides with the first example we gave. | |
Indeed, let $S \subseteq \{1, \dots, n\}$ | |
and let $\xi \in \{\pm1\}^n$ which is $-1$ at positions in $S$, | |
and $+1$ at positions not in $S$. | |
Then the character $\chi_S$ from the previous example | |
coincides with the character $e_\xi$ in the new notation. | |
In particular, $\wh f(S) = \wh f(\xi)$. | |
Thus Fourier analysis on a finite group $Z$ subsumes | |
binary Fourier analysis. | |
\end{example} | |
\subsection{Fourier series for functions $L^2([-\pi, \pi])$} | |
This is the most famous one, and hence the one you've heard of. | |
\begin{definition} | |
The space $L^2([-\pi, \pi])$ consists of all functions | |
$f \colon [-\pi, \pi] \to \CC$ such that | |
the integral | |
$\int_{[-\pi, \pi]} \left\lvert f(x) \right\rvert^2 \; dx$ | |
exists and is finite, | |
modulo the relation that a function which is zero ``almost everywhere'' | |
is considered to equal zero.\footnote{We won't define this, yet, | |
as it won't matter to us for now. | |
But we will elaborate more on this in the parts on measure theory. | |
There is one point at which this is relevant. | |
Often we require that the function $f$ satisfies $f(-\pi) = f(\pi)$, | |
so that $f$ becomes a periodic function, | |
and we can think of it as $f \colon \TT \to \CC$. | |
This makes no essential difference | |
since we merely change the value at one point.} | |
It is made into an inner product space according to | |
\[ \left< f,g \right> | |
= \frac{1}{2\pi} \int_{[-\pi, \pi]} f(x) \ol{g(x)} \; dx. \] | |
\end{definition} | |
It turns out (we won't prove) that this is an | |
(infinite-dimensional) Hilbert space! | |
Now, the beauty of Fourier analysis is that | |
\textbf{this space has a great basis}: | |
\begin{theorem} | |
[The classical Fourier basis] | |
For each integer $n$, define | |
\[ e_n(x) = \exp(inx). \] | |
Then $e_n$ form an orthonormal basis | |
of the Hilbert space $L^2([-\pi, \pi])$. | |
\end{theorem} | |
Thus this time the frequency set $\ZZ$ is infinite, and we have | |
\[ f(x) = \sum_n \wh f(n) \exp(inx) | |
\quad\text{almost everywhere} \] | |
for coefficients $\wh f(n)$ | |
with $\sum_n \left\lvert \wh f(n) \right\rvert^2 < \infty$. | |
Since the frequency set is indexed by $\ZZ$, | |
we call this a \vocab{Fourier series} | |
to reflect the fact that the index is $n \in \ZZ$. | |
\begin{exercise} | |
Show once again | |
\[ \wh f(0) = \frac{1}{2\pi} \int_{[-\pi, \pi]} f(x) \; dx. \] | |
\end{exercise} | |
\section{Summary, and another teaser} | |
We summarize our various flavors of Fourier analysis in the following table. | |
\[ | |
\begin{array}{llll} | |
\hline | |
\text{Type} & \text{Physical var} & \text{Frequency var} | |
& \text{Basis functions} \\ \hline | |
\text{Binary} & \{\pm1\}^n | |
& \text{Subsets } S \subseteq \left\{ 1, \dots, n \right\} | |
& \prod_{s \in S} x_s \\ | |
\text{Finite group} & Z & \xi \in Z, \text{ choice of } \cdot | |
& e(\xi \cdot x) \\ | |
\text{Fourier series} & \TT \text{ or } [-\pi, \pi] & n \in \ZZ | |
& \exp(inx) \\ | |
\text{Discrete} & \Zc n | |
& \xi \in \Zc n | |
& e(\xi x / n) \\ | |
\end{array} | |
\] | |
I snuck in a fourth row with $Z = \Zc n$, | |
but it's a special case of the second row, so no cause for alarm. | |
Alluding to the future, I want to hint at how \Cref{ch:pontryagin} starts. | |
Each one of these is really a statement | |
about how functions from $G \to \CC$ | |
can be expressed in terms of functions $\wh G \to \CC$, | |
for some ``dual'' $\wh G$. | |
In that sense, we could rewrite the above table as: | |
\[ | |
\begin{array}{llll} | |
\hline | |
\text{Name} & \text{Domain }G & \text{Dual }\wh G | |
& \text{Characters} \\ \hline | |
\text{Binary} & \{\pm1\}^n | |
& S \subseteq \left\{ 1, \dots, n \right\} | |
& \prod_{s \in S} x_s \\ | |
\text{Finite group} & Z | |
& \xi \in \wh Z \cong Z & e( i \xi \cdot x) \\ | |
\text{Fourier series} & \TT \cong [-\pi, \pi] & n \in \ZZ | |
& \exp(inx) \\ | |
\text{Discrete} & \ZZ/n\ZZ & \xi \in \ZZ/n\ZZ | |
& e(\xi x / n) \\ | |
\end{array} | |
\] | |
It will turn out that in general | |
we can say something about many different domains $G$, | |
once we know what it means to integrate a measure. | |
This is the so-called \emph{Pontryagin duality}; | |
and it is discussed as a follow-up bonus in \Cref{ch:pontryagin}. | |
\section{Parseval and friends} | |
Here is a fun section in which you get to learn a lot of big names quickly. | |
Basically, we can take each of the three results | |
from Proposition~\ref{prop:orthonormal}, | |
translate it into the context of our Fourier analysis | |
(for which we have an orthonormal basis of the Hilbert space), | |
and get a big-name result. | |
\begin{corollary} | |
[Parseval theorem] | |
Let $f \colon Z \to \CC$, where $Z$ is a finite abelian group. | |
Then \[ \sum_\xi |\wh f(\xi)|^2 = \frac{1}{|Z|} \sum_{x \in Z} |f(x)|^2. \] | |
Similarly, if $f \colon [-\pi, \pi] \to \CC$ is square-integrable then | |
its Fourier series satisfies | |
\[ \sum_n |\wh f(n)|^2 = \frac{1}{2\pi} \int_{[-\pi, \pi]} |f(x)|^2 \; dx. \] | |
\end{corollary} | |
\begin{proof} | |
Recall that $\left< f,f\right>$ is equal to the | |
square sum of the coefficients. | |
\end{proof} | |
\begin{corollary} | |
[Fourier inversion formula] | |
Let $f : Z \to \CC$, where $Z$ is a finite abelian group. | |
Then \[ \wh f(\xi) = \frac{1}{|Z|} \sum_{x \in Z} f(x) \ol{e_\xi(x)}. \] | |
Similarly, if $f : [-\pi, \pi] \to \CC$ is square-integrable then | |
its Fourier series is given by | |
\[ \wh f(n) = \frac{1}{2\pi} \int_{[-\pi, \pi]} f(x) \exp(-inx) \; dx. \] | |
\end{corollary} | |
\begin{proof} | |
Recall that in an orthonormal basis $(e_\xi)_\xi$, | |
the coefficient of $e_\xi$ in $f$ is $\left< f, e_\xi\right>$. | |
\end{proof} | |
\begin{ques} | |
What happens when $\xi = 0$ above? | |
\end{ques} | |
\begin{corollary} | |
[Plancherel theorem] | |
Let $f : Z \to \CC$, where $Z$ is a finite abelian group. | |
Then \[ \left< f,g \right> = \sum_{\xi \in Z} \wh f(\xi) \ol{\wh g(\xi)}. \] | |
Similarly, if $f : [-\pi, \pi] \to \CC$ is square-integrable then | |
\[ \left< f,g \right> = \sum_n \wh f(n) \ol{\wh g(n)}. \] | |
\end{corollary} | |
\begin{ques} | |
Prove this one in one line (like before). | |
\end{ques} | |
\section{Application: Basel problem} | |
One cute application about Fourier analysis on $L^2([-\pi, \pi])$ | |
is that you can get some otherwise hard-to-compute sums, | |
as long as you are willing to use a little calculus. | |
Here is the classical one: | |
\begin{theorem} | |
[Basel problem] | |
We have | |
\[ \sum_{n \ge 1} \frac{1}{n^2} = \frac{\pi^2}{6}. \] | |
\end{theorem} | |
The proof is to consider the identity function $f(x) = x$, | |
which is certainly square-integrable. | |
Then by Parseval, we have | |
\[ | |
\sum_{n \in \ZZ} \left\lvert \wh f(n) \right\rvert^2 | |
= \left< f,f\right> | |
= \frac{1}{2\pi} \int_{[-\pi, \pi]} \left\lvert f(x) \right\rvert^2 \; dx. | |
\] | |
A calculus computation gives | |
\[ \frac{1}{2\pi} \int_{[-\pi, \pi]} x^2 \; dx = \frac{\pi^2}{3}. \] | |
On the other hand, we will now compute all Fourier coefficients. | |
We have already that | |
\[ \wh f(0) = \frac{1}{2\pi} \int_{[-\pi, \pi]} f(x) \; dx | |
= \frac{1}{2\pi} \int_{[-\pi, \pi]} x \; dx = 0. \] | |
For $n \neq 0$, we have by definition | |
(or ``Fourier inversion formula'', if you want to use big words) | |
the formula | |
\begin{align*} | |
\wh f(n) &= \left< f, \exp(inx) \right> \\ | |
&= \frac{1}{2\pi} \int_{[-\pi, \pi]} x \cdot \ol{\exp(inx)} \; dx \\ | |
&= \frac{1}{2\pi} \int_{[-\pi, \pi]} x \exp(-inx) \; dx. | |
\end{align*} | |
The anti-derivative is equal to | |
$\frac{1}{n^2} \exp(-inx) (1+inx)$, | |
which thus with some more calculation gives that | |
\[ \wh f(n) = \frac{(-1)^n}{n} i. \] | |
So | |
\[ \sum_n \left\lvert \wh f(n) \right\rvert^2 | |
= 2 \sum_{n \ge 1} \frac{1}{n^2} \] | |
implying the result. | |
\section{Application: Arrow's Impossibility Theorem} | |
As an application of binary Fourier analysis, | |
we now prove a form of | |
\href{https://en.wikipedia.org/wiki/Arrow's_impossibility_theorem}{Arrow's theorem}. | |
Consider $n$ voters voting among $3$ candidates $A$, $B$, $C$. | |
Each voter specifies a tuple $v_i = (x_i, y_i, z_i) \in \{\pm1\}^3$ as follows: | |
\begin{itemize} | |
\ii $x_i = 1$ if person $i$ ranks $A$ ahead of $B$, and $x_i = -1$ otherwise. | |
\ii $y_i = 1$ if person $i$ ranks $B$ ahead of $C$, and $y_i = -1$ otherwise. | |
\ii $z_i = 1$ if person $i$ ranks $C$ ahead of $A$, and $z_i = -1$ otherwise. | |
\end{itemize} | |
Tacitly, we only consider $3! = 6$ possibilities for $v_i$: | |
we forbid ``paradoxical'' votes of the form $x_i = y_i = z_i$ | |
by assuming that people's votes are consistent | |
(meaning the preferences are transitive). | |
For brevity, let $x_\bullet = (x_1, \dots, x_n)$ | |
and define $y_\bullet$ and $z_\bullet$ similarly. | |
Then, we can consider a voting mechanism | |
\begin{align*} | |
f \colon \{\pm1\}^n &\to \{\pm1\} \\ | |
g \colon \{\pm1\}^n &\to \{\pm1\} \\ | |
h \colon \{\pm1\}^n &\to \{\pm1\} | |
\end{align*} | |
such that | |
\begin{itemize} | |
\ii $f(x_\bullet)$ is the global preference of $A$ vs.\ $B$, | |
\ii $g(y_\bullet)$ is the global preference of $B$ vs.\ $C$, | |
\ii and $h(z_\bullet)$ is the global preference of $C$ vs.\ $A$. | |
\end{itemize} | |
We'd like to avoid situations where the global preference | |
$(f(x_\bullet), g(y_\bullet), h(z_\bullet))$ is itself paradoxical. | |
Let $\EE f$ denote the average value of $f$ across all $2^n$ inputs. | |
Define $\EE g$ and $\EE h$ similarly. | |
We'll add an assumption that $\EE f = \EE g = \EE h = 0$, | |
which provides symmetry | |
(and e.g.\ excludes the possibility that $f$, $g$, $h$ | |
are constant functions which ignore voter input). | |
With that we will prove the following result: | |
\begin{theorem} | |
[Arrow Impossibility Theorem] | |
Assume that $(f,g,h)$ always avoids paradoxical outcomes, | |
and assume $\EE f = \EE g = \EE h = 0$. | |
Then $(f,g,h)$ is either a dictatorship or anti-dictatorship: | |
there exists a ``dictator'' $k$ such that | |
\[ f(x_\bullet) = \pm x_k, \qquad g(y_\bullet) = \pm y_k, | |
\qquad h(z_\bullet) = \pm z_k \] | |
where all three signs coincide. | |
\end{theorem} | |
Unlike the usual Arrow theorem, we do \emph{not} assume | |
that $f(+1, \dots, +1) = +1$ (hence possibility of anti-dictatorship). | |
\begin{proof} | |
Suppose the voters each randomly select one of the $3!=6$ | |
possible consistent votes. | |
In \Cref{prob:arrow_lemma} it is shown | |
that the exact probability of a paradoxical outcome | |
for any functions $f$, $g$, $h$ is given exactly by | |
\[ \frac14 + \frac14 \sum_{S \subseteq \{1, \dots, n\}} | |
\left( -\frac13 \right)^{\left\lvert S \right\rvert} | |
\left( \wh f(S) \wh g(S) + \wh g(S) \wh h(S) + \wh h(S) \wh f(S) \right). | |
\] | |
Assume that this probability (of a paradoxical outcome) equals $0$. | |
Then, we derive | |
\[ 1 = \sum_{S \subseteq \{1, \dots, n\}} | |
-\left( -\frac13 \right)^{\left\lvert S \right\rvert} | |
\left( \wh f(S) \wh g(S) + \wh g(S) \wh h(S) + \wh h(S) \wh f(S) \right). \] | |
But now we can just use weak inequalities. | |
We have $\wh f(\varnothing) = \EE f = 0$ and similarly for $\wh g$ and $\wh h$, | |
so we restrict attention to $|S| \ge 1$. | |
We then combine the famous inequality $|ab+bc+ca| \le a^2+b^2+c^2$ | |
(which is true across all real numbers) to deduce that | |
\begin{align*} | |
1 &= \sum_{S \subseteq \{1, \dots, n\}} | |
-\left( -\frac13 \right)^{\left\lvert S \right\rvert} | |
\left( \wh f(S) \wh g(S) + \wh g(S) \wh h(S) + \wh h(S) \wh f(S) \right) \\ | |
&\le \sum_{S \subseteq \{1, \dots, n\}} | |
\left( \frac13 \right)^{\left\lvert S \right\rvert} | |
\left( \wh f(S)^2 + \wh g(S)^2 + \wh h(S)^2 \right) \\ | |
&\le \sum_{S \subseteq \{1, \dots, n\}} \left( \frac13 \right)^1 | |
\left( \wh f(S)^2 + \wh g(S)^2 + \wh h(S)^2 \right) \\ | |
&= \frac13 (1+1+1) = 1. | |
\end{align*} | |
with the last step by Parseval. | |
So all inequalities must be sharp, and in particular $\wh f$, $\wh g$, $\wh h$ | |
are supported on one-element sets, i.e.\ they are linear in inputs. | |
As $f$, $g$, $h$ are $\pm 1$ valued, each $f$, $g$, $h$ is itself | |
either a dictator or anti-dictator function. | |
Since $(f,g,h)$ is always consistent, this implies the final result. | |
\end{proof} | |
\section{\problemhead} | |
\begin{problem} | |
[For calculus fans] | |
Prove that | |
\[ \sum_{n \ge 1} \frac{1}{n^4} = \frac{\pi^4}{90}. \] | |
\begin{hint} | |
Use Parseval again, but this time on $f(x) = x^2$. | |
\end{hint} | |
\end{problem} | |
\begin{problem} | |
\gim | |
\label{prob:arrow_lemma} | |
Let $f,g,h \colon \{\pm1\}^n \to \{\pm1\}$ | |
be any three functions. | |
For each $i$, we randomly select $(x_i, y_i, z_i) \in \{\pm1\}^3$ | |
subject to the constraint that not all are equal | |
(hence, choosing among $2^3-2=6$ possibilities). | |
Prove that the probability that | |
\[ f(x_1, \dots, x_n) = g(y_1, \dots, y_n) = h(z_1, \dots, z_n) \] | |
is given by the formula | |
\[ \frac14 + \frac14 \sum_{S \subseteq \{1, \dots, n\}} | |
\left( -\frac13 \right)^{\left\lvert S \right\rvert} | |
\left( \wh f(S) \wh g(S) + \wh g(S) \wh h(S) + \wh h(S) \wh f(S) \right) | |
\] | |
\begin{hint} | |
Define the Boolean function $D : \{\pm 1\}^3 \to \RR$ by | |
$D(a,b,c) = ab+bc+ca$. | |
Write out the value of $D(a,b,c)$ for each $(a,b,c)$. | |
Then, evaluate its expected value. | |
\end{hint} | |
\begin{sol} | |
Define the Boolean function $D : \{\pm 1\}^3 \to \RR$ by | |
\[ D(a,b,c) = ab + bc + ca | |
= \begin{cases} | |
3 & a,b,c \text{ all equal} \\ | |
-1 & a,b,c \text{ not all equal}. | |
\end{cases}. | |
\] | |
Thus paradoxical outcomes arise when | |
$D(f(x_\bullet), g(y_\bullet), h(z_\bullet)) = 3$. | |
Now, we compute that for randomly selected | |
$x_\bullet$, $y_\bullet$, $z_\bullet$ that | |
\begin{align*} | |
\EE D(f(x_\bullet), g(y_\bullet), h(z_\bullet)) | |
&= \EE \sum_S \sum_T | |
\left( \wh f(S) \wh g(T) + \wh g(S) \wh h(T) + \wh h(S) \wh f(T) \right) | |
\left( \chi_S(x_\bullet)\chi_T(y_\bullet) \right) \\ | |
&= \sum_S \sum_T | |
\left( \wh f(S) \wh g(T) + \wh g(S) \wh h(T) + \wh h(S) \wh f(T) \right) | |
\EE\left( \chi_S(x_\bullet)\chi_T(y_\bullet) \right). | |
\end{align*} | |
Now we observe that: | |
\begin{itemize} | |
\ii If $S \neq T$, then $\EE \chi_S(x_\bullet) \chi_T(y_\bullet) = 0$, | |
since if say $s \in S$, $s \notin T$ then $x_s$ affects | |
the parity of the product with 50\% either way, | |
and is independent of any other variables in the product. | |
\ii On the other hand, suppose $S = T$. | |
Then | |
\[ \chi_S(x_\bullet) \chi_T(y_\bullet) | |
= \prod_{s \in S} x_sy_s. \] | |
Note that $x_sy_s$ is equal to $1$ with probability $\frac13$ | |
and $-1$ with probability $\frac23$ | |
(since $(x_s, y_s, z_s)$ is uniform from $3!=6$ choices, | |
which we can enumerate). | |
From this an inductive calculation on $|S|$ gives that | |
\[ | |
\prod_{s \in S} x_sy_s | |
= | |
\begin{cases} | |
+1 & \text{ with probability } \half(1+(-1/3)^{|S|}) \\ | |
-1 & \text{ with probability } \half(1-(-1/3)^{|S|}). | |
\end{cases} | |
\] | |
Thus | |
\[ \EE \left( \prod_{s \in S} x_sy_s \right) = \left( -\frac13 \right)^{|S|}. \] | |
\end{itemize} | |
Piecing this altogether, we now have that | |
\[ | |
\EE D(f(x_\bullet), g(y_\bullet), h(z_\bullet)) | |
= | |
\left( \wh f(S) \wh g(T) + \wh g(S) \wh h(T) + \wh h(S) \wh f(T) \right) | |
\left( -\frac13 \right)^{|S|}. | |
\] | |
Then, we obtain that | |
\begin{align*} | |
&\EE \frac14 \left( 1 + D(f(x_\bullet), g(y_\bullet), h(z_\bullet)) \right) \\ | |
=& \frac14 + \frac14\sum_S | |
\left( \wh f(S) \wh g(T) + \wh g(S) \wh h(T) + \wh h(S) \wh f(T) \right) | |
\wh f(S)^2 \left( -\frac13 \right)^{|S|}. | |
\end{align*} | |
Comparing this with the definition of $D$ gives the desired result. | |
\end{sol} | |
\end{problem} | |