You are currently browsing the tag archive for the ‘random walks’ tag.
The wave equation is usually expressed in the form
where is a function of both time
and space
, with
being the Laplacian operator. One can generalise this equation in a number of ways, for instance by replacing the spatial domain
with some other manifold and replacing the Laplacian
with the Laplace-Beltrami operator or adding lower order terms (such as a potential, or a coupling with a magnetic field). But for sake of discussion let us work with the classical wave equation on
. We will work formally in this post, being unconcerned with issues of convergence, justifying interchange of integrals, derivatives, or limits, etc.. One then has a conserved energy
which we can rewrite using integration by parts and the inner product
on
as
A key feature of the wave equation is finite speed of propagation: if, at time (say), the initial position
and initial velocity
are both supported in a ball
, then at any later time
, the position
and velocity
are supported in the larger ball
. This can be seen for instance (formally, at least) by inspecting the exterior energy
and observing (after some integration by parts and differentiation under the integral sign) that it is non-increasing in time, non-negative, and vanishing at time .
The wave equation is second order in time, but one can turn it into a first order system by working with the pair rather than just the single field
, where
is the velocity field. The system is then
and the conserved energy is now
Finite speed of propagation then tells us that if are both supported on
, then
are supported on
for all
. One also has time reversal symmetry: if
is a solution, then
is a solution also, thus for instance one can establish an analogue of finite speed of propagation for negative times
using this symmetry.
If one has an eigenfunction
of the Laplacian, then we have the explicit solutions
of the wave equation, which formally can be used to construct all other solutions via the principle of superposition.
When one has vanishing initial velocity , the solution
is given via functional calculus by
and the propagator can be expressed as the average of half-wave operators:
One can view as a minor of the full wave propagator
which is unitary with respect to the energy form (1), and is the fundamental solution to the wave equation in the sense that
Viewing the contraction as a minor of a unitary operator is an instance of the “dilation trick“.
It turns out (as I learned from Yuval Peres) that there is a useful discrete analogue of the wave equation (and of all of the above facts), in which the time variable now lives on the integers
rather than on
, and the spatial domain can be replaced by discrete domains also (such as graphs). Formally, the system is now of the form
where is now an integer,
take values in some Hilbert space (e.g.
functions on a graph
), and
is some operator on that Hilbert space (which in applications will usually be a self-adjoint contraction). To connect this with the classical wave equation, let us first consider a rescaling of this system
where is a small parameter (representing the discretised time step),
now takes values in the integer multiples
of
, and
is the wave propagator operator
or the heat propagator
(the two operators are different, but agree to fourth order in
). One can then formally verify that the wave equation emerges from this rescaled system in the limit
. (Thus,
is not exactly the direct analogue of the Laplacian
, but can be viewed as something like
in the case of small
, or
if we are not rescaling to the small
case. The operator
is sometimes known as the diffusion operator)
Assuming is self-adjoint, solutions to the system (3) formally conserve the energy
This energy is positive semi-definite if is a contraction. We have the same time reversal symmetry as before: if
solves the system (3), then so does
. If one has an eigenfunction
to the operator , then one has an explicit solution
to (3), and (in principle at least) this generates all other solutions via the principle of superposition.
Finite speed of propagation is a lot easier in the discrete setting, though one has to offset the support of the “velocity” field by one unit. Suppose we know that
has unit speed in the sense that whenever
is supported in a ball
, then
is supported in the ball
. Then an easy induction shows that if
are supported in
respectively, then
are supported in
.
The fundamental solution to the discretised wave equation (3), in the sense of (2), is given by the formula
where and
are the Chebyshev polynomials of the first and second kind, thus
and
In particular, is now a minor of
, and can also be viewed as an average of
with its inverse
:
As before, is unitary with respect to the energy form (4), so this is another instance of the dilation trick in action. The powers
and
are discrete analogues of the heat propagators
and wave propagators
respectively.
One nice application of all this formalism, which I learned from Yuval Peres, is the Varopoulos-Carne inequality:
Theorem 1 (Varopoulos-Carne inequality) Let
be a (possibly infinite) regular graph, let
, and let
be vertices in
. Then the probability that the simple random walk at
lands at
at time
is at most
, where
is the graph distance.
This general inequality is quite sharp, as one can see using the standard Cayley graph on the integers . Very roughly speaking, it asserts that on a regular graph of reasonably controlled growth (e.g. polynomial growth), random walks of length
concentrate on the ball of radius
or so centred at the origin of the random walk.
Proof: Let be the graph Laplacian, thus
for any , where
is the degree of the regular graph and sum is over the
vertices
that are adjacent to
. This is a contraction of unit speed, and the probability that the random walk at
lands at
at time
is
where are the Dirac deltas at
. Using (5), we can rewrite this as
where we are now using the energy form (4). We can write
where is the simple random walk of length
on the integers, that is to say
where
are independent uniform Bernoulli signs. Thus we wish to show that
By finite speed of propagation, the inner product here vanishes if . For
we can use Cauchy-Schwarz and the unitary nature of
to bound the inner product by
. Thus the left-hand side may be upper bounded by
and the claim now follows from the Chernoff inequality.
This inequality has many applications, particularly with regards to relating the entropy, mixing time, and concentration of random walks with volume growth of balls; see this text of Lyons and Peres for some examples.
For sake of comparison, here is a continuous counterpart to the Varopoulos-Carne inequality:
Theorem 2 (Continuous Varopoulos-Carne inequality) Let
, and let
be supported on compact sets
respectively. Then
where
is the Euclidean distance between
and
.
Proof: By Fourier inversion one has
for any real , and thus
By finite speed of propagation, the inner product vanishes when
; otherwise, we can use Cauchy-Schwarz and the contractive nature of
to bound this inner product by
. Thus
Bounding by
, we obtain the claim.
Observe that the argument is quite general and can be applied for instance to other Riemannian manifolds than .
In the last three notes, we discussed the Bourgain-Gamburd expansion machine and two of its three ingredients, namely quasirandomness and product theorems, leaving only the non-concentration ingredient to discuss. We can summarise the results of the last three notes, in the case of fields of prime order, as the following theorem.
Theorem 1 (Non-concentration implies expansion in
) Let
be a prime, let
, and let
be a symmetric set of elements in
of cardinality
not containing the identity. Write
, and suppose that one has the non-concentration property
for some
and some even integer
. Then
is a two-sided
-expander for some
depending only on
.
Proof: From (1) we see that is not supported in any proper subgroup
of
, which implies that
generates
. The claim now follows from the Bourgain-Gamburd expansion machine (Theorem 2 of Notes 4), the product theorem (Theorem 1 of Notes 5), and quasirandomness (Exercise 8 of Notes 3).
Remark 1 The same argument also works if we replace
by the field
of order
for some bounded
. However, there is a difficulty in the regime when
is unbounded, because the quasirandomness property becomes too weak for the Bourgain-Gamburd expansion machine to be directly applicable. On theother hand, the above type of theorem was generalised to the setting of cyclic groups
with
square-free by Varju, to arbitrary
by Bourgain and Varju, and to more general algebraic groups than
and square-free
by Salehi Golsefidy and Varju. It may be that some modification of the proof techniques in these papers may also be able to handle the field case
with unbounded
.
It thus remains to construct tools that can establish the non-concentration property (1). The situation is particularly simple in , as we have a good understanding of the subgroups of that group. Indeed, from Theorem 14 from Notes 5, we obtain the following corollary to Theorem 1:
Corollary 2 (Non-concentration implies expansion in
) Let
be a prime, and let
be a symmetric set of elements in
of cardinality
not containing the identity. Write
, and suppose that one has the non-concentration property
for some
and some even integer
, where
ranges over all Borel subgroups of
. Then, if
is sufficiently large depending on
,
is a two-sided
-expander for some
depending only on
.
It turns out (2) can be verified in many cases by exploiting the solvable nature of the Borel subgroups . We give two examples of this in these notes. The first result, due to Bourgain and Gamburd (with earlier partial results by Gamburd and by Shalom) generalises Selberg’s expander construction to the case when
generates a thin subgroup of
:
Theorem 3 (Expansion in thin subgroups) Let
be a symmetric subset of
not containing the identity, and suppose that the group
generated by
is not virtually solvable. Then as
ranges over all sufficiently large primes, the Cayley graphs
form a two-sided expander family, where
is the usual projection.
Remark 2 One corollary of Theorem 3 (or of the non-concentration estimate (3) below) is that
generates
for all sufficiently large
, if
is not virtually solvable. This is a special case of a much more general result, known as the strong approximation theorem, although this is certainly not the most direct way to prove such a theorem. Conversely, the strong approximation property is used in generalisations of this result to higher rank groups than
.
Exercise 1 In the converse direction, if
is virtually solvable, show that for sufficiently large
,
fails to generate
. (Hint: use Theorem 14 from Notes 5 to prevent
from having bounded index solvable subgroups.)
Exercise 2 (Lubotzsky’s 1-2-3 problem) Let
.
- (i) Show that
generates a free subgroup of
. (Hint: use a ping-pong argument, as in Exercise 23 of Notes 2.)
- (ii) Show that if
are two distinct elements of the sector
, then there os no element
for which
. (Hint: this is another ping-pong argument.) Conclude that
has infinite index in
. (Contrast this with the situation in which the
coefficients in
are replaced by
or
, in which case
is either all of
, or a finite index subgroup, as demonstrated in Exercise 23 of Notes 2).
- (iii) Show that
for sufficiently large primes
form a two-sided expander family.
Remark 3 Theorem 3 has been generalised to arbitrary linear groups, and with
replaced by
for square-free
; see this paper of Salehi Golsefidy and Varju. In this more general setting, the condition of virtual solvability must be replaced by the condition that the connected component of the Zariski closure of
is perfect. An effective version of Theorem 3 (with completely explicit constants) was recently obtained by Kowalski.
The second example concerns Cayley graphs constructed using random elements of .
Theorem 4 (Random generators expand) Let
be a prime, and let
be two elements of
chosen uniformly at random. Then with probability
,
is a two-sided
-expander for some absolute constant
.
Remark 4 As with Theorem 3, Theorem 4 has also been extended to a number of other groups, such as the Suzuki groups (in this paper of Breuillard, Green, and Tao), and more generally to finite simple groups of Lie type of bounded rank (in forthcoming work of Breuillard, Green, Guralnick, and Tao). There are a number of other constructions of expanding Cayley graphs in such groups (and in other interesting groups, such as the alternating groups) beyond those discussed in these notes; see this recent survey of Lubotzky for further discussion. It has been conjectured by Lubotzky and Weiss that any pair
of (say)
that generates the group, is a two-sided
-expander for an absolute constant
: in the case of
, this has been established for a density one set of primes by Breuillard and Gamburd.
— 1. Expansion in thin subgroups —
We now prove Theorem 3. The first observation is that the expansion property is monotone in the group :
Exercise 3 Let
be symmetric subsets of
not containing the identity, such that
. Suppose that
is a two-sided expander family for sufficiently large primes
. Show that
is also a two-sided expander family.
As a consequence, Theorem 3 follows from the following two statments:
Theorem 5 (Tits alternative) Let
be a group. Then exactly one of the following statements holds:
- (i)
is virtually solvable.
- (ii)
contains a copy of the free group
of two generators as a subgroup.
Theorem 6 (Expansion in free groups) Let
be generators of a free subgroup of
. Then as
ranges over all sufficiently large primes, the Cayley graphs
form a two-sided expander family.
Theorem 5 is a special case of the famous Tits alternative, which among other things allows one to replace by
for any
and any field
of characteristic zero (and fields of positive characteristic are also allowed, if one adds the requirement that
be finitely generated). We will not prove the full Tits alternative here, but instead just give an ad hoc proof of the special case in Theorem 5 in the following exercise.
Exercise 4 Given any matrix
, the singular values are
and
, and we can apply the singular value decomposition to decompose
where
and
are orthonormal bases. (When
, these bases are uniquely determined up to phase rotation.) We let
be the projection of
to the projective complex plane, and similarly define
.
Let
be a subgroup of
. Call a pair
a limit point of
if there exists a sequence
with
and
.
- (i) Show that if
is infinite, then there is at least one limit point.
- (ii) Show that if
is a limit point, then so is
.
- (iii) Show that if there are two limit points
with
, then there exist
that generate a free group. (Hint: Choose
close to
and
close to
, and consider the action of
and
on
, and specifically on small neighbourhoods of
, and set up a ping-pong type situation.)
- (iv) Show that if
is hyperbolic (i.e. it has an eigenvalue greater than 1), with eigenvectors
, then the projectivisations
of
form a limit point. Similarly, if
is regular parabolic (i.e. it has an eigenvalue at 1, but is not the identity) with eigenvector
, show that
is a limit point.
- (v) Show that if
has no free subgroup of two generators, then all hyperbolic and regular parabolic elements of
have a common eigenvector. Conclude that all such elements lie in a solvable subgroup of
.
- (vi) Show that if an element
is neither hyperbolic nor regular parabolic, and is not a multiple of the identity, then
is conjugate to a rotation by
(in particular,
).
- (vii) Establish Theorem 5. (Hint: show that two square roots of
in
cannot multiply to another square root of
.)
Now we prove Theorem 6. Let be a free subgroup of
generated by two generators
. Let
be the probability measure generating a random walk on
, thus
is the corresponding generator on
. By Corollary 2, it thus suffices to show that
for all sufficiently large , some absolute constant
, and some even
(depending on
, of course), where
ranges over Borel subgroups.
As is a homomorphism, one has
and so it suffices to show that
To deal with the supremum here, we will use an argument of Bourgain and Gamburd, taking advantage of the fact that all Borel groups of obey a common group law, the point being that free groups such as
obey such laws only very rarely. More precisely, we use the fact that the Borel groups are solvable of derived length two; in particular we have
for all . Now,
is supported on matrices in
whose coefficients have size
(where we allow the implied constants to depend on the choice of generators
), and so
is supported on matrices in
whose coefficients also have size
. If
is less than a sufficiently small multiple of
, these coefficients are then less than
(say). As such, if
lie in the support of
and their projections
obey the word law (4) in
, then the original matrices
obey the word law (4) in
. (This lifting of identities from the characteristic
setting of
to the characteristic
setting of
is a simple example of the “Lefschetz principle”.)
To summarise, if we let be the set of all elements of
that lie in the support of
, then (4) holds for all
. This severely limits the size of
to only be of polynomial size, rather than exponential size:
Proposition 7 Let
be a subset of the support of
(thus,
consists of words in
of length
) such that the law (4) holds for all
. Then
.
The proof of this proposition is laid out in the exercise below.
Exercise 5 Let
be a free group generated by two generators
. Let
be the set of all words of length at most
in
.
- (i) Show that if
commute, then
lie in the same cyclic group, thus
for some
and
.
- (ii) Show that if
, there are at most
elements of
that commute with
.
- (iii) Show that if
, there are at most
elements
of
with
.
- (iv) Prove Proposition 7.
Now we can conclude the proof of Theorem 3:
Exercise 6 Let
be a free group generated by two generators
.
- (i) Show that
for some absolute constant
. (For much more precise information on
, see this paper of Kesten.)
- (ii) Conclude the proof of Theorem 3.
— 2. Random generators expand —
We now prove Theorem 4. Let be the free group on two formal generators
, and let
be the generator of the random walk. For any word
and any
in a group
, let
be the element of
formed by substituting
for
respectively in the word
; thus
can be viewed as a map
for any group
. Observe that if
is drawn randomly using the distribution
, and
, then
is distributed according to the law
, where
. Applying Corollary 2, it suffices to show that whenever
is a large prime and
are chosen uniformly and independently at random from
, that with probability
, one has
for some absolute constant , where
ranges over all Borel subgroups of
and
is drawn from the law
for some even natural number
.
Let denote the words in
of length at most
. We may use the law (4) to obtain good bound on the supremum in (5) assuming a certain non-degeneracy property of the word evaluations
:
Exercise 7 Let
be a natural number, and suppose that
is such that
for
. Show that
for some absolute constant
, where
is drawn from the law
. (Hint: use (4) and the hypothesis to lift the problem up to
, at which point one can use Proposition 7 and Exercise 6.)
In view of this exercise, it suffices to show that with probability , one has
for all
for some
comparable to a small multiple of
. As
has
elements, it thus suffices by the union bound to show that
for some absolute constant , and any
of length less than
for some sufficiently small absolute constant
.
Let us now fix a non-identity word of length
less than
, and consider
as a function from
to
for an arbitrary field
. We can identify
with the set
. A routine induction then shows that the expression
is then a polynomial in the eight variables
of degree
and coefficients which are integers of size
. Let us then make the additional restriction to the case
, in which case we can write
and
. Then
is now a rational function of
whose numerator is a polynomial of degree
and coefficients of size
, and the denominator is a monomial of
of degree
.
We then specialise this rational function to the field . It is conceivable that when one does so, the rational function collapses to the constant polynomial
, thus
for all
with
. (For instance, this would be the case if
, by Lagrange’s theorem, if it were not for the fact that
is far too large here.) But suppose that this rational function does not collapse to the constant rational function. Applying the Schwarz-Zippel lemma (Exercise 23 from Notes 5), we then see that the set of pairs
with
and
is at most
; adding in the
and
cases, one still obtains a bound of
, which is acceptable since
and
. Thus, the only remaining case to consider is when the rational function
is identically
on
with
.
Now we perform another “Lefschetz principle” maneuvre to change the underlying field. Recall that the denominator of rational function is monomial in
, and the numerator has coefficients of size
. If
is less than
for a sufficiently small
, we conclude in particular (for
large enough) that the coefficients all have magnitude less than
. As such, the only way that this function can be identically
on
is if it is identically
on
for all
with
, and hence for
or
also by taking Zariski closures.
On the other hand, we know that for some choices of , e.g.
,
contains a copy
of the free group on two generators (see e.g. Exercise 23 of Notes 2). As such, it is not possible for any non-identity word
to be identically trivial on
. Thus this case cannot actually occur, completing the proof of (6) and hence of Theorem 4.
Remark 5 We see from the above argument that the existence of subgroups
of an algebraic group with good “independence” properties – such as that of generating a free group – can be useful in studying the expansion properties of that algebraic group, even if the field of interest in the latter is distinct from that of the former. For more complicated algebraic groups than
, in which laws such as (4) are not always available, it turns out to be useful to place further properties on the subgroup
, for instance by requiring that all non-abelian subgroups of that group be Zariski dense (a property which has been called strong density), as this turns out to be useful for preventing random walks from concentrating in proper algebraic subgroups. See this paper of Breuillard, Guralnick, Green and Tao for constructions of strongly dense free subgroups of algebraic groups and further discussion.
In the previous set of notes we saw how a representation-theoretic property of groups, namely Kazhdan’s property (T), could be used to demonstrate expansion in Cayley graphs. In this set of notes we discuss a different representation-theoretic property of groups, namely quasirandomness, which is also useful for demonstrating expansion in Cayley graphs, though in a somewhat different way to property (T). For instance, whereas property (T), being qualitative in nature, is only interesting for infinite groups such as or
, and only creates Cayley graphs after passing to a finite quotient, quasirandomness is a quantitative property which is directly applicable to finite groups, and is able to deduce expansion in a Cayley graph, provided that random walks in that graph are known to become sufficiently “flat” in a certain sense.
The definition of quasirandomness is easy enough to state:
Definition 1 (Quasirandom groups) Let
be a finite group, and let
. We say that
is
-quasirandom if all non-trivial unitary representations
of
have dimension at least
. (Recall a representation is trivial if
is the identity for all
.)
Exercise 1 Let
be a finite group, and let
. A unitary representation
is said to be irreducible if
has no
-invariant subspaces other than
and
. Show that
is
-quasirandom if and only if every non-trivial irreducible representation of
has dimension at least
.
Remark 1 The terminology “quasirandom group” was introduced explicitly (though with slightly different notational conventions) by Gowers in 2008 in his detailed study of the concept; the name arises because dense Cayley graphs in quasirandom groups are quasirandom graphs in the sense of Chung, Graham, and Wilson, as we shall see below. This property had already been used implicitly to construct expander graphs by Sarnak and Xue in 1991, and more recently by Gamburd in 2002 and by Bourgain and Gamburd in 2008. One can of course define quasirandomness for more general locally compact groups than the finite ones, but we will only need this concept in the finite case. (A paper of Kunze and Stein from 1960, for instance, exploits the quasirandomness properties of the locally compact group
to obtain mixing estimates in that group.)
Quasirandomness behaves fairly well with respect to quotients and short exact sequences:
Exercise 2 Let
be a short exact sequence of finite groups
.
- (i) If
is
-quasirandom, show that
is
-quasirandom also. (Equivalently: any quotient of a
-quasirandom finite group is again a
-quasirandom finite group.)
- (ii) Conversely, if
and
are both
-quasirandom, show that
is
-quasirandom also. (In particular, the direct or semidirect product of two
-quasirandom finite groups is again a
-quasirandom finite group.)
Informally, we will call quasirandom if it is
-quasirandom for some “large”
, though the precise meaning of “large” will depend on context. For applications to expansion in Cayley graphs, “large” will mean “
for some constant
independent of the size of
“, but other regimes of
are certainly of interest.
The way we have set things up, the trivial group is infinitely quasirandom (i.e. it is
-quasirandom for every
). This is however a degenerate case and will not be discussed further here. In the non-trivial case, a finite group can only be quasirandom if it is large and has no large subgroups:
Exercise 3 Let
, and let
be a finite
-quasirandom group.
- (i) Show that if
is non-trivial, then
. (Hint: use the mean zero component
of the regular representation
.) In particular, non-trivial finite groups cannot be infinitely quasirandom.
- (ii) Show that any proper subgroup
of
has index
. (Hint: use the mean zero component of the quasiregular representation.)
The following exercise shows that quasirandom groups have to be quite non-abelian, and in particular perfect:
Exercise 4 (Quasirandomness, abelianness, and perfection) Let
be a finite group.
- (i) If
is abelian and non-trivial, show that
is not
-quasirandom. (Hint: use Fourier analysis or the classification of finite abelian groups.)
- (ii) Show that
is
-quasirandom if and only if it is perfect, i.e. the commutator group
is equal to
. (Equivalently,
is
-quasirandom if and only if it has no non-trivial abelian quotients.)
Later on we shall see that there is a converse to the above two exercises; any non-trivial perfect finite group with no large subgroups will be quasirandom.
Exercise 5 Let
be a finite
-quasirandom group. Show that for any subgroup
of
,
is
-quasirandom, where
is the index of
in
. (Hint: use induced representations.)
Now we give an example of a more quasirandom group.
Lemma 2 (Frobenius lemma) If
is a field of some prime order
, then
is
-quasirandom.
This should be compared with the cardinality of the special linear group, which is easily computed to be
.
Proof: We may of course take to be odd. Suppose for contradiction that we have a non-trivial representation
on a unitary group of some dimension
with
. Set
to be the group element
and suppose first that is non-trivial. Since
, we have
; thus all the eigenvalues of
are
roots of unity. On the other hand, by conjugating
by diagonal matrices in
, we see that
is conjugate to
(and hence
conjugate to
) whenever
is a quadratic residue mod
. As such, the eigenvalues of
must be permuted by the operation
for any quadratic residue mod
. Since
has at least one non-trivial eigenvalue, and there are
distinct quadratic residues, we conclude that
has at least
distinct eigenvalues. But
is a
matrix with
, a contradiction. Thus
lies in the kernel of
. By conjugation, we then see that this kernel contains all unipotent matrices. But these matrices generate
(see exercise below), and so
is trivial, a contradiction.
Exercise 6 Show that for any prime
, the unipotent matrices
for
ranging over
generate
as a group.
Exercise 7 Let
be a finite group, and let
. If
is generated by a collection
of
-quasirandom subgroups, show that
is itself
-quasirandom.
Exercise 8 Show that
is
-quasirandom for any
and any prime
. (This is not sharp; the optimal bound here is
, which follows from the results of Landazuri and Seitz.)
As a corollary of the above results and Exercise 2, we see that the projective special linear group is also
-quasirandom.
Remark 2 One can ask whether the bound
in Lemma 2 is sharp, assuming of course that
is odd. Noting that
acts linearly on the plane
, we see that it also acts projectively on the projective line
, which has
elements. Thus
acts via the quasiregular representation on the
-dimensional space
, and also on the
-dimensional subspace
; this latter representation (known as the Steinberg representation) is irreducible. This shows that the
bound cannot be improved beyond
. More generally, given any character
,
acts on the
-dimensional space
of functions
that obey the twisted dilation invariance
for all
and
; these are known as the principal series representations. When
is the trivial character, this is the quasiregular representation discussed earlier. For most other characters, this is an irreducible representation, but it turns out that when
is the quadratic representation (thus taking values in
while being non-trivial), the principal series representation splits into the direct sum of two
-dimensional representations, which comes very close to matching the bound in Lemma 2. There is a parallel series of representations to the principal series (known as the discrete series) which is more complicated to describe (roughly speaking, one has to embed
in a quadratic extension
and then use a rotated version of the above construction, to change a split torus into a non-split torus), but can generate irreducible representations of dimension
, showing that the bound in Lemma 2 is in fact exactly sharp. These constructions can be generalised to arbitrary finite groups of Lie type using Deligne-Luzstig theory, but this is beyond the scope of this course (and of my own knowledge in the subject).
Exercise 9 Let
be an odd prime. Show that for any
, the alternating group
is
-quasirandom. (Hint: show that all cycles of order
in
are conjugate to each other in
(and not just in
); in particular, a cycle is conjugate to its
power for all
. Also, as
,
is simple, and so the cycles of order
generate the entire group.)
Remark 3 By using more precise information on the representations of the alternating group (using the theory of Specht modules and Young tableaux), one can show the slightly sharper statement that
is
-quasirandom for
(but is only
-quasirandom for
due to icosahedral symmetry, and
-quasirandom for
due to lack of perfectness). Using Exercise 3 with the index
subgroup
, we see that the bound
cannot be improved. Thus,
(for large
) is not as quasirandom as the special linear groups
(for
large and
bounded), because in the latter case the quasirandomness is as strong as a power of the size of the group, whereas in the former case it is only logarithmic in size.
If one replaces the alternating group
with the slightly larger symmetric group
, then quasirandomness is destroyed (since
, having the abelian quotient
, is not perfect); indeed,
is
-quasirandom and no better.
Remark 4 Thanks to the monumental achievement of the classification of finite simple groups, we know that apart from a finite number (26, to be precise) of sporadic exceptions, all finite simple groups (up to isomorphism) are either a cyclic group
, an alternating group
, or is a finite simple group of Lie type such as
. (We will define the concept of a finite simple group of Lie type more precisely in later notes, but suffice to say for now that such groups are constructed from reductive algebraic groups, for instance
is constructed from
in characteristic
.) In the case of finite simple groups
of Lie type with bounded rank
, it is known from the work of Landazuri and Seitz that such groups are
-quasirandom for some
depending only on the rank. On the other hand, by the previous remark, the large alternating groups do not have this property, and one can show that the finite simple groups of Lie type with large rank also do not have this property. Thus, we see using the classification that if a finite simple group
is
-quasirandom for some
and
is sufficiently large depending on
, then
is a finite simple group of Lie type with rank
. It would be of interest to see if there was an alternate way to establish this fact that did not rely on the classification, as it may lead to an alternate approach to proving the classification (or perhaps a weakened version thereof).
A key reason why quasirandomness is desirable for the purposes of demonstrating expansion is that quasirandom groups happen to be rapidly mixing at large scales, as we shall see below the fold. As such, quasirandomness is an important tool for demonstrating expansion in Cayley graphs, though because expansion is a phenomenon that must hold at all scales, one needs to supplement quasirandomness with some additional input that creates mixing at small or medium scales also before one can deduce expansion. As an example of this technique of combining quasirandomness with mixing at small and medium scales, we present a proof (due to Sarnak-Xue, and simplified by Gamburd) of a weak version of the famous “3/16 theorem” of Selberg on the least non-trivial eigenvalue of the Laplacian on a modular curve, which among other things can be used to construct a family of expander Cayley graphs in (compare this with the property (T)-based methods in the previous notes, which could construct expander Cayley graphs in
for any fixed
).
Van Vu and I have just uploaded to the arXiv our preprint “A sharp inverse Littlewood-Offord theorem“, which we have submitted to Random Structures and Algorithms. This paper gives a solution to the (inverse) Littlewood-Offord problem of understanding when random walks are concentrated in the case when the concentration is of polynomial size in the length of the walk; our description is sharp up to epsilon powers of
. The theory of inverse Littlewood-Offord problems and related topics has been of importance in recent developments in the spectral theory of discrete random matrices (e.g. a “robust” variant of these theorems was crucial in our work on the circular law).
For simplicity I will restrict attention to the Bernoulli random walk. Given real numbers
, one can form the random variable
where are iid random signs (with either sign +1, -1 chosen with probability 1/2). This is a discrete random variable which typically takes
values. However, if there are various arithmetic relations between the step sizes
, then many of the
possible sums collide, and certain values may then arise with much higher probability. To measure this, define the concentration probability
by the formula
.
Intuitively, this probability measures the amount of additive structure present between the . There are two (opposing) problems in the subject:
- (Forward Littlewood-Offord problem) Given some structural assumptions on
, what bounds can one place on
?
- (Inverse Littlewood-Offord problem) Given some bounds on
, what structural assumptions can one then conclude about
?
Ideally one would like answers to both of these problems which come close to inverting each other, and this is the guiding motivation for our paper.
Recent Comments