Friday, December 08, 2006

Group representation

(As will become clear soon, I am totally out of my depth: Readers who find glaring mistakes, and even small ones, should note them in the comments.)

In the previous post on expanders we left off with the observation that if we want to construct an expanding Cayley graph, then we need to start from a group where all $n$ elements of the group can be obtained by a sum of $O(log n)$ terms chosen, with repetitions, from a set of $O(1)$ generators; this means that a constant root of the $n^{O(1)}$ ways of rearranging such sums must lead to different results. The group, then, has to be non-commutative in a very strong sense.

I will describe a way in which a group used to construct expanders is "very non-commutative." This will not be a sufficient property for the expansion, but it is closely related and it will come up in the paper of Gowers that we will talk about later.

With this objective in mind, let's see the notion of group representation. Let's start from the simplest group, the cyclic group of $N$ elements $Z_N$ with the operation of addition (or $Z/NZ$ as mathematicians write). One can visualize this group as $N$ points equally spaced along a circle in the plane, with element $a$ forming an angle of $2\pi a/ N$ with the $x$ axis. If we want to add an element that forms an angle $\alpha$ to an element that forms an angle $\beta$, that's the same as putting the "two angles next to each other" and we get a point that forms an angle $\alpha+\beta$. More algebraically, we identify an element $a$ with the complex number $e^{2\pi a i/n}$ and we realize group addition as multiplication of complex numbers. There is, however, another way of visualizing the cyclic group on a cycle. We can think of group element $a$ as the operation that takes a point on the cycle and moves it by an angle of $2\pi a/N$. Addition of group elements now corresponds to "composition," that is, corresponds to applying one rotation after the other.

In this view, element $a$ is now the function that maps complex number $x$ into complex number $x\cdot e^{2\pi a i/N}$.

Suppose now that our group is a product $Z_{N_1} \times \cdots \times Z_{N_k}$ of cyclic groups. This means that a group elements if a $k$-tuple of the form $(a_1,\ldots,a_d)$, and that

$(a_1,\ldots,a_d) + (b_1,\ldots,b_d) = (a_1+b_1 \bmod N_1,\ldots,a_k+b_k \bmod N_k)$

It now makes sense to view a group element $(a_1,\ldots,a_d)$ as a "high-dimensional rotation" operation that takes in input $k$ complex numbers $(x_1,\ldots,x_k)$ and outputs the $k$ complex numbers

$(x_1 \cdot e^{2\pi a_1 i /N_1},\ldots, x_k \cdot e^{2\pi a_k i /N_k})$

If we take this view, we have, again, the property that group addition becomes composition of functions. Note, also, that the function that we have associated to each group element is a very simple type of linear function: it is simply multiplication of $x$ times a diagonal matrix that has diagonal

$(e^{2\pi a_1 i /N_1},\ldots, e^{2\pi a_k i /N_k})$

Notice, also, that if $f$ is a function of the form $f(x) = A\cdot x$, where $A$ is a matrix, and $g$ is a function of the form $B\cdot x$ where $B$ is a matrix, then $f(g(x))= A\cdot B\cdot x$. That is, for linear functions, function composition is the same as matrix multiplication.

To summarize, we have started from the group $Z_{N_1} \times \cdots \times Z_{N_k}$, and we have found a way to associate a complex-valued diagonal matrix to each group element in such a way that group addition becomes matrix multiplication.

It is known that all finite Abelian groups can be written as $Z_{N_1} \times \cdots \times Z_{N_k}$, so this type of representation via diagonal matrices is possible for every finite Abelian group.

What about more general groups? If $G$ is an arbitrary finite group, it is possible to associate to each element a block-diagonal matrix in such a way that group addition becomes matrix multiplication.

(It is common to call a group operation "multiplication" when a group is not Abelian, but for consistency with the rest of the post I will keep calling it addition.)

(By the way, it is also possible to represent infinite groups by associating a linear operator to each group element, but we will only discuss finite groups here.)

If the group is Abelian, then the matrices are diagonal, that is, they are block-diagonal matrices with "block size one." So one way of quantifying "how non-Abelian" is a group is to consider how small the blocks can be in such matrix representations. That's the dimension of the representation.

Here is an example of a family of groups whose representations cannot be low-dimensional. Let $p$ be a prime (it could also be a prime power) and let us consider $2 \times 2$ matrices whose entries are integers $\bmod p$. Let us restrict ourselves to matrices whose determinant is 1 modulo $p$, and consider the operation of matrix multiplication (where, also, all operations are mod $p$). This set of matrices forms a group, because the matrices are invertible (the determinant is non-zero) and the set contains the identity matrix and is closed under product. This group is called $SL(2,p)$. The group $SL(2,p)$ contains the tiny subgroup of two elements $\{ I,-I\}$; in the representation of $SL(2,p)$ this shows up as a block of size 1. If we take the quotient of $SL(2,p)$ by $\{ I,-I \}$ then we get another group, which is called $PSL(2,p)$.

It is now a theorem of Frobenius that every representation of $PSL(2,p)$ has dimension at least $(p-1)/2$. This is really large compared to the size of $PSL(2,p)$: the group $SL(2,p)$ has $p(p-1)^2$ elements, and so $PSL(2,p)$ has $p(p-1)^2/2$ elements. The dimension of the representation is thus approximately the cube root of the number of elements of the group.

Going back to representations of Abelian groups, we see that, in that case, not only each block had size one, but also that the entire matrix had size at most logarithmic in the number of elements of the group. This shows that $SL(2,p)$ and $PSL(2,p)$ are, at least in this particular sense, "very non-Abelian," and it is an encouraging sign that they may be useful in constructing expanders and other quasi-random objects.

On his way toward a question about subsets of groups not containing triples of the form $\{ a, b, a+b\}$, Gowers shows that every dense Cayley graph constructed on $PSL(2,p)$ has constant edge expansion. (By dense, we mean that the degree is $\Omega(n)$, where $n$ is the number of vertices.) This is a result that, assuming Frobenius theorem, has a simple proof, which I hope to understand and describe at a later point.

The celebrated Ramanujan graphs of Lubotzky, Phillips and Sarnak are constant-degree Cayley graphs constructed from $PGL(2,p)$, which is the same as $PSL(2,p)$ except that we put no restriction on the determinant. Their relationship between degree and eigenvalue gap is best possible. The analysis of Lubotzky et al. is, unfortunately, completely out of reach.

Wednesday, December 06, 2006

Expanders and groups

I have started to read a new paper by Gowers titled Quasi-random groups. "Quasi-random" is the more-or-less standard term used to denote a single object that has some of the typical properties of a random object; this is distinct from the term "pseudo-random" which is more often applied to a distribution that satisfies a certain set of properties with approximately the same probability as the uniform distribution.

(The distinction is sometimes artificial: for example the support of a pseudorandom distribution can also be thought of as a quasi-random set. To add confusion, in older literature on extractors certain high-entropy sources were called "quasi-random" sources.)

Although the paper does not directly talk about expander graphs, it addresses related questions, and it gives an opportunity for a post (or more than one) on expander graph constructions.

Informally, an expander graph is a graph that, while having small bounded degree, is very well-connected. For example, a $\sqrt{n} \times \sqrt{n}$ grid can be disconnected into two equal-size connected components after removing just $\sqrt{n}$ edges. This means that a $\Omega(1)$ fraction of all pairs of vertices can be disconneted while only removing a $o(1)$ fraction of all edges; for this reason, we do not think of a 2-dimensional grid as a good expander. In a graph of degree $O(1)$ it is always possible to cut off $k$ vertices from the rest of the graph by removing $O(k)$ edges; a graph is an expander if this is best possible. That is, a graph is an expander if for every set of $k$ vertices (where $k$ is less than half of the total number of vertices) you need to remove at least $\Omega(k)$ edges to disconnect those vertices from the rest of the graph.

More quantitatively, a $d$-regular graph has edge expansion at least $\epsilon$ if for every set $S$ of vertices there are at least $\epsilon \cdot |S|$ edges going between vertices in $S$ and vertices not in $S$. (Provided $|S|$ is less than half the total number of vertices.)

We have a good construction of expanders if for fixed constants $\epsilon$ and $d$ we are able to construct $\epsilon$-expanders of degree $d$ with any desired number of vertices.

The applications of expander graphs are vast, they are useful to reduce randomness in probabilistic algorithms, they come up in certain constructions of data structures, of randomness extractors, and of error-correcting codes, and they have many applications in complexity theory. Most recently, Dinur's new proof of the PCP Theorem uses expanders. A great set of lecture notes on expanders is here, and not so great one is here (lectures 8-12 are on expanders).

Before the great zig-zag graph product revolution, almost all the good expander graph constructions were Cayley graphs of non-Abelian groups. Let's see what this means and how it helps.

If $G$ is a group and $D\subset G$ is a subset, then $G$ and $D$ define a $|D|$-regular graph with $|G|$ vertices as follows: every element of $G$ is a vertex, and vertices $u$ and $v$ are connected if $u-v \in D$. If we assume that, for every $a\in D$ we also have $-a \in D$, then we can think of the graph as being undirected.

A first trivial observation is that if $D$ is a set of generators, that is, if we can realize every element of $G$ as a sum of elements of $D$, then the graph is connected, otherwise it is disconnected.

Another point, which is true for all graphs, not just Cayley graphs, is that we can approximate the expansion of a graph by computing the eigenvalues of the adjacency matrix and looking at the difference between the largest and the second largest in absolute value. This was a point that I used to find mystifying, until I studied it and realized that it makes perfect sense. I may talk about it in another post, but I want to mention an amusing fact. The connection between eigenvalues and expansion has an easy direction (if there is a large eigenvalue gap then there is large expansion) and a difficult direction (if there large expansion then there must be large eigenvalue gap). The easy direction was "well known," and the difficult direction was discovered around the same time in two distinct communities for opposite reasons.

Alon discovered it while working on constructions of expanders. It was known how to construct graphs with noticeable eigenvalue gap which would then (by the "simple direction") be good expanders. This is because, for the algebraic constructions that we will talk about in a minute, it is easier to reason about the eigenvalues of the adjacency matrix than about the number of edges in a generic cut. Alon, by proving the hard direction, showed that the approach via eigenvalue can be taken "with no loss of generality,"

Others (Aldous and Diaconis? Jerrum and Sinclair? I never get this reference right) discovered the same result working on the problem of bounding the time that it takes for a random walk on certain graphs to converge to the stationary distribution. To answer such questions, one needs to know the eigenvalue gap of the graphs, a quantity that is difficult to estimate directly in important cases. For those graphs, however, it is possible to reason about the size of cuts, and hence, via the difficult connection, obtain results on the eigenvalue gap (and, thus, on the convergence time, or "mixing" time of a random walk).

This "difficult direction" result has also an important algorithmic application. The counterpositive statement is that if largest and second largest eigenvalue are close in value, then the graph is not expanding, so one can cut off a set of vertices by deleting a comparatively small number of vertices. The proof is algorithmic: given the eigenvector of the second largest eigenvalue, this small cut can be found efficiently. This is the principle at work in "spectral partitioning" algorithms, often used in practice (and, occasionally, in theory).

Well, that was a long digression, but it's also an interesting story. Back to our constructions of expander graphs, we want to describe a bounded-degree graph explicitly and prove that there is a noticeable gap between largest and second largest eigenvalue. If the graph is a Cayley graph derived from a group $G$ and set of generators $D$ (and $D$ is chosen properly), it turns out that the eigenvalues of the adjacency matrix of the graph can be explicitly computed by looking at the representations of the group $G$. If $G$ is Abelian (if the group operation is commutative), then the representation theory of $G$ is very simple and it is "just" the theory of Fourier transforms of functions $f: G \rightarrow C$ (where C are the complex numbers), which is much easier than it sounds.

Unfortunately, a Cayley graph based on an Abelian group can never be a good expander. To see why, you'll have to believe my claim that in a good expander with $n$ vertices the diameter is at most $O(\log n)$. Consider now an $n$-vertex Cayley graph of degree $d$. Every vertex $x$ is reachable from vertex $0$ via a path of length $O(\log n)$. That is, $x$ is the sum of $O(\log n)$ elements of the set $D$. Because of commutativity, we can rearrange the sum so that it looks as

$(a_1+ \ldots + a_1) + (a_2 + \ldots + a_2) +\ldots (a_d + \ldots + a_d)$

where $a_1,\ldots,a_d$ are the elements of $D$. We see that the sum is specified just by saying how many time each element of $D$ must be added, and so $x$ is specified by $d$ numbers, each between $1$ and $O(\log n)$. This gives $n \leq O((\log n)^d)$ which is a contradiction.

Indeed, we see that Cayley graphs based on Abelian groups fail very badly to have the required diameter condition: if the degree is $d$, there will be vertices at distance about $n^{1/d}$ from each other, a bound that is met by a $d$-dimensional grid.

Intuitively, then, we can get an expanding bounded-degree Cayley graph only if we start from a "very non-Abelian" group. But how do we quantify the "Abelianity" of a group, and what kind of groups are "very non-Abelian"? This will be the subject of the next post.

Monday, December 04, 2006

Post-modern cryptography

Oded Goldreich has written an essay in response to two essays on "provable security" by Koblitz and Menezes. Oded says that "Although it feels ridiculous to answer [the claims of Koblitz and Menezes], we undertake to do so in this essay. In particular, we point out some of the fundamental philosophical flaws that underly the said article and some of its misconceptions regarding theoretical research in Cryptography in the last quarter of a century."

Neil Koblitz spoke here at IPAM in October on the somewhat related matter of how to interpret results in the Random Oracle model and in the Generic Group model. There is an audio file of his talk.