Lectures on Random Matrix Theory
Lectures on Random Matrix Theory
1 Overview 7
1.1 What is a random matrix? . . . . . . . . . . . . . . . . . . . . . . 7
1.2 The main limit theorems . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 The semicircle law . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 Fluctuations in the bulk: the sine process . . . . . . . . . 11
1.2.3 Fluctuations at the edge: the Airy point process . . . . . 12
1.2.4 Fredholm determinants, Painlevé equations, and integrable
systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.2.5 Universality . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3 Connections to other areas of mathematics . . . . . . . . . . . . 15
1.3.1 Number theory . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3.2 Combinatorics and enumerative geometry . . . . . . . . . 17
1.3.3 Random permutations . . . . . . . . . . . . . . . . . . . . 18
1.3.4 Spectral and inverse spectral theory of operators . . . . . 19
4 Determinantal formulas 53
4.1 Probabilities as determinants . . . . . . . . . . . . . . . . . . . . 53
4.2 The m-point correlation function . . . . . . . . . . . . . . . . . . 55
3
4 CONTENTS
D Notation 125
D.1 Indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
D.2 Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
D.3 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
D.4 Lie groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
D.5 Banach spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6 CONTENTS
Chapter 1
Overview
1 − 1 Tr(M 2 )
pGUE (M )DM = e 2 DM. (1.1.1)
Zn
7
8 CHAPTER 1. OVERVIEW
where β = 1,2 and 4 for GOE, GUE and GSE respectively. While it is true
that there are no other ensembles that respect fundamental physical invariance
(in the sense of Dyson), many fundamental results of random matrix theory
can be established for all β > 0. These results follow from the existence of
ensembles of tridiagonal matrices, whose eigenvalues have a joint distribution
that interpolates those of the β = 1,2 and 4 ensembles to all β > 0 [8].
Gaudin and Mehta [26, 25], on the Gaussian ensembles served to elucidate the
fundamental limit theorems of random matrix theory. In this section we outline
these theorems, assuming always that the ensemble is GUE. Our purpose is
to explain the form of the main questions (and their answers) in the simplest
setting. All the results hold in far greater generality as is briefly outlined at the
end of this section.
By the normalization (1.1.1), a GUE matrix has independent standard nor-
mal entries on its diagonal (mean zero, variance 1). The off-diagonal entries
have mean zero and variance 1/2. We denote the ordered eigenvalues of the
GUE matrix by λ1 ≤ λ2 ≤ . . . λn . A fundamental heuristic for GUE matrices
(that will be proven later, and may √ be easily simulated) is√ that the largest√ and
smallest eigenvalues have size O( n). In fact, λ1 ≈ −2 n and λn ≈ 2 n as
n → ∞. Since√there are n eigenvalues, the gap between these eigenvalues is
typically O(1/ n). There are thus two natural scaling limits to consider as
n → ∞:
1. Rescale M 7→ n−1/2 M so that the spectral radius is O(1). In this scaling
limit, n eigenvalues are contained within a bounded interval, and we obtain
a deterministic limit called the semicircle law .
2. Rescale M 7→ n1/2 M so that the gaps between eigenvalues are O(1). In
this scaling limit, we expect a random limiting point process. The limiting
point process is a determinantal point process called the Sine2 process.
In fact, the situation is more subtle. While the expected value of the gap
between eigenvalues for a GUE matrix is indeed O(1/n), the gaps are O(n−2/3 )
about the edge of the spectrum. There is an an entirely different scaling
√ limit
called the Airy2 process obtained by rescaling the spectrum of M ± 2 nI.
In all that follows, we consider a sequence of random matrices of size n
sampled from GUE(n). To make this explicit, the matrix is denoted Mn , and
(n) (n) (n)
its ordered eigenvalues are denoted λ1 ≤ λ2 ≤ · · · ≤ λn .
are called the semicircle density and the semicircle distribution respectively.
Theorem 2. Let Mn be a sequence of GUE matrices of size n. The rescaled
empirical spectral measures
n
1X
µn (dx) = δ (n) (dx) (1.2.2)
n j=1 n−1/2 λj
Theorem 2 may also be interpreted √ as the statement that the empirical spec-
tral distribution of the matrices Mn / n converges to the semicircle distribution.
The shortest proof of Theorem (2) uses the following integral transform.
Definition 3. Assume µ is a measure on R that satisfies the finiteness condition
Z ∞
1
√ µ(dx) < ∞. (1.2.3)
−∞ 1 + x2
The Stieltjes transform of µ is the function
Z ∞
1
µ̂(z) = µ(dx), z ∈ C\R. (1.2.4)
−∞ x − z
The general proof of Theorem 2 uses a recursive expression for the law of Tr(z −
Mn )−1 . As n → ∞, the fixed point of this recursion, Rsc solves the quadratic
equation
R2 − zR + 1 = 0. (1.2.6)
It is then easy to verify that
1 p
Rsc (z) = −z + z 2 − 4 , z ∈ C\[−2, 2]. (1.2.7)
2
We recover the semicircle law from Rsc (z) by evaluating the jump in Im(Rsc (z))
across the branch cut [−2, 2].
Remark 4. The heuristic to determine the typical spacings is the following.
(n)
Define γj ∈ [−2, 2] by the relation
(n)
Z γj
j
= psc (x)dx, j = 1, 2, . . . , n.
n −∞
(n) √ (n)
Then the approximation λj ≈ nγj should hold1 . We have
(n)
Z γj+1
1 (n) (n) (n)
= psc (x)dx ≈ (γj+1 − γj )psc (γj ). (1.2.8)
n (n)
γj
(n)
If j = j(n) is chosen so that γj → r, r ∈ (−2, 2) (i.e. in the “bulk”) we have
(n) (n) 1
λj+1 − λj ≈√ .
npsc (r)
1 This is made rigorous and quantitative by Erdős, Yau and Yin [13].
1.2. THE MAIN LIMIT THEOREMS 11
(n)
At the edge, consider (noting that γ1 > −2)
(n) (n)
γ1 γ1
1√
Z Z 3/2
1 2 (n)
= psc (x)dx ≈ 2 + xdx = γ1 + 2 ,
n −2 −2 π 3π
(n) c
γ1 +2≈
,
n2/3
√ (n) √
2 n + λ1 = O(n−1/6 ), λ(n)
n − 2 n = O(n
−1/6
). (1.2.9)
Though perhaps mysterious at first sight, the origin of this formula is rather
simple. Integral operators with some smoothness and boundedness (in particu-
Rb
lar, continuous integral operators K whose trace a |K(x, x)|dx is finite) may be
approximated on a discrete-grid of size h by a finite-dimensional discretization
Kh . The determinant (I −Kh ) is then the usual determinant of a matrix and we
may use the definition of the determinant to expand det(I − Kh ) in a finite se-
ries, which is nothing but the infinite series above in the instance when all terms
beyond m = rank(Kh ) vanish. This approach was pioneered by Fredholm in
1900 before the development of functional analysis. From a probabilistic point
of view, this formula arises from the Inclusion-Exclusion Principle, taken to the
limit. But the operator theory pioneered by Fredholm allows for that limit to
be understood.
12 CHAPTER 1. OVERVIEW
The Airy function is one of the classical special functions [1]. It admits
several alternative definitions. For instance, the oscillatory integral in (1.2.14)
may be deformed into an absolutely convergent integral in the complex plane.
This argument allows us to establish that the Airy function is entire and to
determine its asymptotic expansions as x → ±∞.
These properties may also be established using the theory of ordinary dif-
ferential equations in the complex plane [17]. It is easy to verify from (1.2.14),
after deformation, that Ai(x) satisfies the differential equation
Equation (1.2.15) admits two linearly independent solutions, only one of which
decays as x → ∞. Up to a (fixed by convention, but otherwise arbitrary)
normalization constant, the decaying solution is Ai(x).
1.2. THE MAIN LIMIT THEOREMS 13
Observe that both the sine and Airy kernel have the form
t t2 t3
σ(t) = − − 2 − 3, t ↓ 0. (1.2.20)
π π π
Theorem 11 (Tracy-Widom distribution [36]). For all real t,
Z ∞
2
F2 (t) := det 1 − KAiry 1(t,∞) = exp − (s − t)q (s) ds , (1.2.21)
t
Despite the fact that KdV is nonlinear, it may be solved explicitly through the
inverse scattering transform. We will not discuss this method in detail. But in
order to make the connection with random matrix theory, let us note that if one
seeks self-similar solutions to KdV of the form
1 x
u(x, t) = q (1.2.25)
(3t)2/3 (3t)2/3
2 Paul Painlevé was rather restless: he began in mathematics, became an early aviation
enthusiast, and then turned to politics. He rose to become the Prime Minister of France for
part of World War I, and was later the designer of the disastrous Maginot line.
1.3. CONNECTIONS TO OTHER AREAS OF MATHEMATICS 15
1.2.5 Universality
We have restricted attention to matrices from GUE to present some of the
central theorems in the subject in an efficient manner. One of the main achieve-
ments of the past decade has been the establishment of universality – informally,
this is the notion that the limiting fluctuations in the bulk and edge described
by the Sine2 and Airy2 processes, hold for both Wigner and invariant ensembles
which satisfy natural moment assumptions. The idea of universality is of clear
practical importance (we need understand only a few universal limits). It also
appears to hold the key to some of the connections between random matrix the-
ory and other areas of mathematics. The explanation of these connections may
lie in the fact that determinantal point processes, such as the Sine2 and Airy2
process, have the simplest structure of strongly interacting point processes. By
contrast, Poisson processes, while universal, describe non-interacting points.
the cornerstone of analytic number theory. The basic facts are as follows. The
function ζ(z) extends to a meromorphic function on C by analytic continuation,
which has a simple pole at s = 1 where the residue is 1. A closely related
function is the Riemann ξ-function
1 s
ξ(s) = s(s − 1)Γ ζ(s). (1.3.3)
2π s/2 2
Recall that the Γ function is a ‘continuous interpolation’ of the factorial, defined
by the integral Z ∞
Γ(s) = e−x xs−1 dx, Re(s) > 0. (1.3.4)
0
The Γ-function extends to a meromorphic function C, which has simple poles at
. . . , −2, −1, 0 where the residue is 1. These poles cancel the ‘trivial’ zeros of the
ζ function, and the essential difficulties related to the study of the ζ function
are more transparent for the ξ function. It satisfies the functional equation
ξ(s) = ξ(1 − s), s ∈ C. (1.3.5)
The celebrated Riemann Hypothesis is the conjecture that all zeros of the ξ
function lie on the critical line Re(s) = 1/2 (this line is the symmetry axis for
the functional equation above). In his fundamental paper on the distribution
of prime numbers (translated in [12] and [30]) Riemann presented a series of
asymptotic expansions that would imply rigorous bounds on the distribution of
primes if the Riemann Hypothesis is true.
The connection between random matrix theory and the Riemann Hypoth-
esis is two-fold. First, if one could construct a Hermitian operator with point
spectrum whose eigenvalues coincide with the zeros of ξ(i(s − 1/2) then the
Riemann Hypothesis would follow immediately (since all eigenvalues of a Her-
mitian operator are real). The catch, of course, is to determine such an operator.
Nevertheless, as we discuss below, random matrix theory has shed new lie on
the spectral theory of several operators, deterministic and random. Thus, the
theory provides a catalog of ‘guesses’. Second, if one assumes the Riemann hy-
pothesis, the fluctuations in the zeros of ζ(s) are described by the sine-kernel!
Under the Riemann hypothesis, the non-trivial zeros of ζ(s) may be written
γn = 21 ± itn with 0 < t1 < t2 < . . .. Let
∞
tn tn X
wn = log , and N (x) = 1wn ≤x . (1.3.6)
2π 2π
k=1
In order to study the gaps between eigenvalues, we must consider instead the
empirical measures
1 X
µ2 (dl; x) = δwj −wk (dl). (1.3.8)
x
1≤j,k≤N (x);j6=k
Under the assumption that f is band-limited, i.e. that its Fourier transform has
compact support, Montgomery established the following
Theorem 12 (Montgomery). Assume the Riemann Hypothesis. Assume f is
a Schwartz function whose Fourier transform fˆ is supported in [−1, 1]. Then
Z ∞ 2 !
sin πl
lim R2 (f ; x) = f (l)µ2 (dl), µ2 (dl) = 1 − dl. (1.3.10)
x→∞ −∞ πl
The point here is that the right hand side of (1.3.10) is precisely the 2-point
function for the sine process. More generally, Montgomery’s theorem is now
known to hold for the distribution of n-consecutive gaps. That is, the rescaled
fluctuations converge to the Sine2 process in distribution. Bourgade’s thesis is
an excellent review of the state of the art [4].
the m-the Catalan number. An analytic proof of this identity follows from a
comparison between the Stieltjes transform Rsc (z), and the generating function
√
X
m 1 − 1 − 4x
Ĉ(x) = Cm x = . (1.3.12)
x
m≥0
The quartic nonlinearity prevents us from expressing this integral in closed form.
Nevertheless, this integral may be expanded in a Taylor series
∞
(−z)k
Z
X k
Zn (z) = Tr(M 4 ) pGUE (M ) DM, Re(z) > 0. (1.3.14)
k!
k=0
Note that only finitely many terms in the sum are non-zero. The series above
is an instance of a genus-expansion. It illustrates the beautiful fact that matrix
integrals serve as the generating functions for Riemann surfaces with a given
combinatorial decomposition!
This problem was one of the first, if not the first, problem to be investigated by
Monte Carlo simulation on a computer — Ulam performed simulations in the
early 60’s [37] and conjectured that
1
√ E [`(Πn )] → c.
n
It was later established by Vershik and Kerov that c = 2 [38]. The detailed
numerical computations of Odlyzko and Rains [28] indicated
√
E [`(Πn )] − 2 n = O(n−1/6 ). (1.3.16)
The comparison between (1.2.9) and (1.3.16) should be striking. Indeed, the
following is often called the Baik–Deift–Johansson Theorem and it makes this
scaling rigorous.
That is, the limit is the same as the largest eigenvalue of a random Hermitian
matrix.
discover that psc (x) is the spectral density for the seminfinite tridiagonal ma-
trix that is 1 on the off-diagonal, and 0 in all other entries. This follows from
the continued fraction expansion
1
Rsc (−z) = (1.3.17)
1
z−
1
z−
z − ...
Ensembles of tridiagonal matrices are of practical important in numerical
linear algebra. For instance, a key pre-processing step while solving symmet-
ric linear systems is to transform the matrix to tridiagonal form by House-
holder’s procedure (Lecture 3). Dumitriu and Edelman pushed forward the
Gaussian measures under this procedure to obtain a family of tridiagonal en-
sembles, known as the general-β ensembles [8]. Further, Edelman and Sutton
made a formal expansion of these operators, and observed that as n → ∞, the
tridiagonal operators appeared to converge to the stochastic Airy operator [11]:
d2 2
Hβ = − 2
+ x + √ ḃ, 0<x<∞ (1.3.18)
dx β
Integration on spaces of
matrices
In this section, we review the geometry of the classical Lie groups, as well as the
spaces Symm(n), Her(n) and Quart(n) and explain how to integrate over these
groups and spaces. Given an point on a manifold M ∈ M, we use dM to
denote the differential of M , i.e., an infintesimal element on the tangent space
TM M at M . We reserve DM to refer to a (naturally induced) volume form
defined using an inner-product on the tangent space. Note that for x ∈ R,
dx = Dx. Our main goal is the following
The main strategy to prove Theorem 15 is to treat the mapping from ma-
trices with distinct eigenvalues to their eigenvalues and eigenvectors. Then we
identify the tangent spaces, and give a formula that relates the tangent space
for the the eigenvalues and the tangent space for the eigenvectors to the tangent
space for the matrix. This formula allows one to change variables in the metric
tensor and therefore in the volume form.
21
22 CHAPTER 2. INTEGRATION ON SPACES OF MATRICES
in the exercises. The essential aspect of (2.0.1) is that the Jacobian for diago-
nalization is given by |4(Λ)|β . This has far-reaching consequences for random
matrix theory and has the interesting physical interpretation of eigenvalue re-
pulsion.
In what follows, we first present a detailed description of integration on O(n)
and Symm(n). The ideas are then extended to Her(n) and Quart(n).
−1
Now, let f : X → R. We define (f is measurable if f ◦ TX is Lebesgue
measurable)
Z Z q
−1
f DX = f (TX x) det(g X ) dx1 . . . dxp , (2.1.7)
X Rp
with a similar expression for g : Y → R. One should check that this definition
is independent of the choice of basis. Now choose f (x) = g(F x). We find
Z Z q
−1
f DX = g(F TX x) det(g X ) dx1 . . . dxp . (2.1.8)
X Rp
If (ui )1≤i≤p is a basis for X then (F ui )1≤i≤p is a basis for Y and we perform
all these calculations for this basis. Consider the transformation
−1
S : Rp → R p , Sx = TY F TX . (2.1.9)
Finally, given A, Ã ∈ o(n), we define their inner product hA, Ãi = Tr(AT Ã) =
− Tr(AÃ). This inner-product is natural, because it is invariant under left-
translation. That is, for two vector OA, OÃ ∈ TO O(n) we find Tr OA)T (OÃ =
Tr(AT Ã). The associated volume form on O(n) is called Haar measure. It is
unique, up to a normalizing factor, and we write
Y
DO = 2n(n−1)/4 dAjk . (2.1.19)
1≤j<k≤n
It can be verified that this is independent of the choice of ϕO . So, now consider
such mapping at the identity, ϕI . And choose
We find
Z Z Y
f DO = 2n(n−1)/4 f (OϕI (OT A)) dAjk . (2.1.23)
Bε (O) Oϕ−1
I (Bε (I)) 1≤j<k≤n
In comparing with, (2.1.11), we use the fact that O furnishes an isometry from
TI O(n) to TO O(n) so that
Z Z Y
n(n−1)/4
f DO = 2 f (OϕI (A)) dAjk . (2.1.24)
Bε (O) ϕ−1
I (Bε (I)) 1≤j<k≤n
R
In particular, if we choose f ≡ 1, then Bε (O) DO does not depend on O ∈ O(n),
showing that this is indeed uniform measure on O(n).
TM Symm(n) ∼
= Rn ⊕ o(n), (2.2.4)
TΛ Symm(n) ∼
= Rn ⊕ o(n). (2.2.6)
Remark 17. The proof of Lemma 2 reveals that all matrices of the form
Z t
M+ Q(s) [Q(s)T Q̇(s), Λ] Q(s)T ds (2.2.8)
0
Ṁ = [A, M ]. (2.2.9)
2 Differentiability is guaranteed by classical perturbation theory [21, Theorem 5.4].
2.2. WEYL’S FORMULA ON SYMM(N ) 27
Thus, the metric tensor in these coordinates is a diagonal matrix in Symm+ (p)
that takes the value 1 on the first n coordinates, and the value 2(λj − λk )2 for
each term Ajk . By (2.1.2), the volume form is
n
Y Y
DM = 2n(n−1)/4 dλj |λj − λk | dAjk = |4(Λ)| DΛ DO. (2.2.13)
j=1 1≤j<k≤n
We now work to understand how to define such a map, and why matrices with
repeated eigenvalues do not cause further issues.
28 CHAPTER 2. INTEGRATION ON SPACES OF MATRICES
This follows from (3.3.3) and Lemma 8 below. Truly, one has to note that
the procedure of reducing a full matrix to a tridiagonal matrix that is used to
establish (3.3.3) does not affect the first row of the eigenvector matrix.
1
µ(DM ) = exp (− Tr(f (M ))) DM. (2.4.2)
Z
This is the most general form of an invariant probability distribution.
By contrast, a Wigner distribution relies on independence of the entries of
M . This means that if a Wigner distribution has a density, then it must be of
the form
n
1 Y Y
µ(DM ) = fj (Mjj ) fjk (Mjk ) DM. (2.4.3)
Z j=1
1≤j<k≤n
On the other hand, since µ(DM ) is invariant, p(M (t)) = p(M ) and
dp
= 0. (2.4.9)
dt t=0
f0 g0
= cM11 + b, = cM22 + b, (2.4.12)
f g
which integrates to
2 2
cM11 cM22
f (M11 ) = f (0)e 2 ebM11 , g(M22 ) = g(0)e 2 ebM22 . (2.4.13)
Since p(M ) integrates to 1, we must have c < 0, say c = −1/σ 2 . The scalar b is
arbitrary and contributes a shift in the mean that is a scalar multiple of I. The
combination of constants f (0)g(0)h(0) may be absorbed into the normalization
constant Z −1 . We have thus proved Theorem 18 for n = 2.
In order to prove Theorem 18 for arbitrary n we generalize the above argu-
ment as follows. Fix a pair of off-diagonal indices 1 ≤ l < m ≤ n. We consider
a rotation in Rn that rotates the xl xm plane as above, and leaves the other co-
ordinates invariant. This entails replacing the matrix R in the argument above
with the matrix Rlm ∈ so(n) with coordinates Rjk lm
= δjl δkm − δjm δkl . The
2.5. INTEGRATION ON HER(N ) AND U(N ) 31
argument above now shows that the density of p in the Mll , Mlm and Mmm
coordinates is a Gaussian distribution of the form (2.4.14):
Tr((M lm )2 ) lm
p(M lm ) = ec 2 eb Tr(M ) , (2.4.15)
At this stage, the constants c and b depend on l and m. But now note that
since the same argument applies to every pair of indices 1 ≤ l < m ≤ n, the
constants c and b must be independent of l and m.
(2.5.5)
The transformation M 7→ iM is an isomorphism between Hermitian and anti-
Hermitian matrices. In fact, the map Her(n) → U(n), M 7→ eiM is onto and
32 CHAPTER 2. INTEGRATION ON SPACES OF MATRICES
x = c0 + c1 e1 + c2 e2 + c3 e3 , ci ∈ R, i = 0, 1, 2, 3, (2.6.1)
These rules ensure that the product of any two quaternions is again a quaternion.
Each x ∈ H has a complex conjugate x̄ = c0 − c1 e1 − c2 e2 − c3 e3 , and its absolute
value |x| is determined by
It follows that the entries of M † are Mjk = M̄kj . We say that an operator is self-
adjoint if M = M † . It is anti self-adjoint if M = −M † . The space of self-adjoint
operators is denoted Quart(n). We equip this space with the Hilbert-Schmidt
norm as before.
The group USp(n) is the set of linear transformations of Hn that preserve
this inner product. We thus require that for each x, y ∈ Hn
Exercises
2.1. Show that
1 ... 1
λ1 . . . λn
4(Λ) = det . (2.6.6)
.. ..
. .
λn−1
1 ... λn−1
n .
2.2. The Pauli matrices,
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = , (2.6.7)
1 0 i 0 0 −1
(a) Show that the Pauli matrices together with the identity matrix span Her(2).
(b) Show that the matrices {iσ1 , iσ2 , iσ3 } form a basis of su(2). (This is the
subalgebra of u(2) consisting of trace-free matrices).
Sp(2n, R) = S ∈ Rn×n S T JS = J .
(2.6.9)
Verify that Sp(2n, R) is a group and compute its Lie algebra sp(2n, R).
2.4. Use the Gaussian integral
|x|2
Z
e− 2 dx1 . . . dxn .
Rn
to compute the n − 1-dimensional volume ωn−1 of the unit sphere S n−1 . Deter-
mine the asymptotic behavior of ωn−1 as n → ∞.
Hint: Do the integral two ways– once in Cartesian coordinates, and once in
polar coordinates.
2.5. Assume given a C 1 function f : (a, b) → R, and extend it to a function
f : Symm(n) → Symm(n) as in (2.4.1). Compute the Jacobian of this transfor-
mation. Apply this formula to the function f (x) = eix to compute the analogue
of Weyl’s formula on U(n) (note that each U ∈ U(n) is of the form eiM for some
M ∈ Her(n)).
2.6. Prove Lemma 4.
2.6. INTEGRATION ON QUART(N ) AND USP(N ) 35
2.7. Let A ∈ Rm×n for m < n. Show that {x | Ax = 0} ⊂ Rn has zero Lebesgue
measure.
2.8. Assume f : R → (0, ∞) satisfies the functional equation
It is easy to check that for each a ∈ R functions of the form f (x) = eax
solve (2.6.10). Show that these are the only solutions to (2.6.10) assuming
only that f is continuous. (Do not assume that f is differentiable).
Remark 19. The use of row operations in Problem (1) underlies the intro-
duction of orthogonal polynomials. Problems (2) and (3) may be combined to
show that Sp(2n, C) ∩ U(n) ∼ = USp(n). The approach in Problem (4) yields
the volume of O(n), U(n) and USp(n) when applied to GOE, GUE and GSE.
The assumptions of Problem (7) may be weakened further – measurability is
enough! You could try to develop a similar approach for the functional equation
implicit in the proof of Theorem 18. That is, can you establish a stronger form
of Theorem 18 that does not assume differentiability ?
2.9. Show that the mapping A 7→ (I −A)(A+I)−1 from o(n) to O(n) is bijective
in a neighborhood of 0 to a neighborhood of the identity. Construct an atlas of
O(n) using this mapping.
2.10. Using the Submersion Theorem [5, Proposition 3.42] (also called the Reg-
ular Value theorem) show that O(n) is a smooth manifold.
Hint: Consider φ : Rn×n → Symm(n) defined by φ(X) = X T X. Then show
that I is a regular value and therefore φ−1 (I) = O(n) is a smooth manifold.
36 CHAPTER 2. INTEGRATION ON SPACES OF MATRICES
Chapter 3
Jacobi matrices, and their closure within the space Tridiag(n) are the manifolds
37
38 CHAPTER 3. JACOBI MATRICES AND TRIDIAGONAL ENSEMBLES
The point here is that the Hermite(β) ensembles are the push-forwards of
the Gaussian ensembles when β = 1, 2 or 4. Then they interpolate Dyson’s
classification of ensembles to every β > 0. When combined with classical spec-
tral theory, they provide a distinct, and important, perspective on the limit
theorems of random matrix theory. Our immediate goal in this chapter is the
following
Theorem 21. Fix β > 0 and assume T ∼ Hermite(β). Then the marginal
distribution of its eigenvalues is
1 β 2
pHermite(β) (Λ)DΛ = e− 4 Tr(Λ) |4(Λ)|β DΛ. (3.1.3)
Zn,β
The chapter concludes with a more refined version of Theorem 21 that in-
cludes the distribution of the spectral measure of matrices T ∼ Hermite(β).
(n−1) ṽ1
ṽ1 = |w1 |e1 − w1 , v1 = , P (1) = Pv1 . (3.2.6)
|ṽ1 |
Note that Q(k) leaves the first k rows and columns of M (k) unchanged, thus
it does not destroy the tridiagonal structure of the first k rows and columns.
Thus, M (k+1) has the form (3.2.10) with the index k replaced by k + 1.
The procedure terminates when k = n − 2, and yields
The exponential weights cancel, and yield the Jacobian formula (3.2.3).
M = QT Q∗ . (3.3.1)
where SFl denotes the unit sphere in Fl , with F = C (resp. H). The volume form
DM on Her(n) (resp. Quart(n)) transforms as follows:
n n−1 n−2
β(n−k)−1
Y Y Y
DM = Cn daj bk dbk Dωl (3.3.3)
j=1 k=1 l=1
where Dωl denotes uniform measure on the sphere SFl , and Cn is a universal
constant.
For a vector w ∈ Cn with independent standard normal complex entries,
wj ∼ √12 (N1 + iN2 ), where N1 , N2 ∼ N (0, 1) are independent, |w| ∼ √12 χ2n .
For a quarternion vector w, one finds |w| ∼ 12 χ4n . So, β is introduced in this
way.
Remark 27. Note that the matrix T is always real, whereas the entries of M
and Q are in C or H.
The proof of Theorem 26 is in the same vein as that of Theorem 22. It is
only necessary to replace the Householder projections in O(n) with projections
in U(n) and USp(n). For example, given v ∈ Cn with |v| = 1, the associ-
ated Householder projection in U(n) is Pv = I − 2vv ∗ . Step 3 in the proof of
Theorem 26 also explains the role of the parameter β in the definition of the
Hermite-β ensembles. The k-th step of the Householder transformation maps a
standard Gaussian vector in Cn−k to its magnitude and direction. The law of
the magnitude is now χ2(n−k) (or χβ(n−k) with β = 2). Similarly, the direction
of the Gaussian vector is uniformly distributed on the unit sphere in Cn−k−1 .
42 CHAPTER 3. JACOBI MATRICES AND TRIDIAGONAL ENSEMBLES
The following theorem shows that the interior of the isospectral manifold is
n−1
diffeomorphic to the positive orthant S+ = {u ∈ Rn | |u| = 1, uj > 0, j =
1, 2, . . . , n} of the unit sphere. Given T , we uniquely define Q by forcing the
first non-zero entry in each column to be positive.
is an analytic diffeomorphism.
It is often more convenient to work with the Cauchy transform of the spectral
measure, µ. Define the τ -function,
n
u2j
Z
1 X
µ 7→ τ (z) = µ(dx) = , z ∈ C\{λ1 , . . . , λn }. (3.4.6)
R x−z λ −z
j=1 j
n−1
{aj }nj=1 {bj }j=1 {πj }n−1
j=0
= bj
,j +1
3-term recurrence
aj , Tj
Monic OPs
Tjj =
T ∈ Jac(n)
Spec
t ral m
ap Spectral measure
P 2
(Λ, QT e1 ) j uj δλj
Figure 3.4.1: The construction of the spectral map and its inverse. The trans-
formation to spectral variables is computed by computing eigenvalues and taking
the first component of the (normalized) eigenvectors. Then a spectral measure
(3.4.5) is created from this data and is used to define monic orthogonal polyno-
mials (3.4.16). These polynomials satisfy a three-term recurrence relation (see
Lemma 11) and the coefficients in the relation allow for the (unique) reconstruc-
n−1
tion of T , see (3.4.21). This shows the spectral map from Jac(n) to W n × S+
is invertible.
We may apply the Gram-Schmidt procedure to the monomials {xk }∞ k=0 to con-
struct a sequence of polynomials that are orthogonal in L2 (R, µ). There are two
standard normalizations that one may adopt.
Monic polynomials, denoted {πk }∞ k=0 have the property that πk (x) is of
degree k and the coefficient of xk is 1. Further,
Z
πk (x)πl (x) µ(dx) = 0, k 6= l. (3.4.16)
R
Remark 29. If µ is not a discrete measure of the form (3.4.5), but has bounded
support, the recurrence (3.4.17) defines a bounded Jacobi operator on l2 (C).
Proof. Given any µ as in (3.4.14), we obtain the sequence {πk } using the Gram-
Schmidt procedure. When µ is of the form (3.4.5) with (3.4.5), the vector space
L2 (R, µ) has dimension n and the Gram-Schmidt procedure yields an orthogonal
basis {π0 , π1 , . . . , πn−1 } for L2 (R, µ).
The three-term recurrence for the orthogonal polynomials is obtained as
follows. Since xπk (x) is a polynomial of degree k + 1 it can be expressed as a
46 CHAPTER 3. JACOBI MATRICES AND TRIDIAGONAL ENSEMBLES
Pk+1
linear combination xπk (x) = j=0 cj,k πj (x). Since the πj are monic, we must
have ck+1,k = 1. Moreover, for j = 0, . . . , k − 2
Z Z
xπk (x)πj (x) µ(dx) = πk (x)xπj (x) µ(dx) = 0,
R R
since xπj lies in the span of {π0 , . . . , πk−1 }. Thus, cj,k = 0 for j = 0, . . . , k − 2
and we find
xπk (x) = πk+1 (x) + ck,k πk (x) + ck−1,k πk−1 (x). (3.4.19)
R
RIt remains
2
to show that ck−1,k > 0. By orthogonality, R xπk (x)πk+1 (x) µ(dx) =
π
R k+1
(x) µ(dx). Thus, ck,k−1 > 0 for all k such that πk−1 (x) does not vanish in
L2 (R, µ): Assume πl does not vanish in L2 (R, µ) for l = 0, 1, 2, . . . , k −1 < n−1.
Then this recurrence defines πk which is not the zero polynomial since it is
monic.
R 2 For Λ ∈ W n , it has distinct diagonal entries, so p(x) 6= 0 implies
p (x)µ(dx) > 0 if p is a polynomial of degree less than n. This is (3.4.17)
aside from a change in notation.
Proof of Theorem 28. We have defined a forward map T 7→ µ(T ) as follows.
The matrix T defines a τ -function τ (z) = eT1 (T − zI)−1 e1 , which is expressed
as a ratio of characteristic polynomials in (3.4.7). The poles of τ (z) are the
eigenvalues of T . The norming constrants are the residues at the poles, and are
given by
dn−1 (λk )
u2k = − 0 , 1 ≤ k ≤ n. (3.4.20)
dn (λk )
The inverse map µ → T (µ) is given by Lemma 11. The orthogonal polynomials
defined by µ satisfy a three-term recurrence whose coefficients determine T .
We only need to show that the map µ 7→ T (µ) 7→ µ (T (µ)) is the identity.
Let µ ∼
= (Λ, u) be given and define T (µ) by the recurrence relations. We will
show that
n
u2k
Z
1 X
eT1 (T − zI)−1 e1 = µ(dx) = . (3.4.21)
R x−z λk − z
k=1
We first show that the eigenvalues of T coincide with {λk }. Define pj (x) =
Qj
πj (x) k=1 b−1
k , π0 (x) = p0 (x), then
This proves the theorem and this approach extends to the bi-infinite Jacobi
operators [6].
Remark 30. Observe that the recurrence relation (3.4.17) may be rewritten as
the matrix equation,
a1 − z 1 0 ... 0
b21 a2 − z 1 ... 0 π0 (z) 0
π1 (z) 0
2 ..
0 b 2 a3 − z . 0 . = . .
.
.. ..
. . .
.. .. .. ..
1 πn−1 (z) −πn (z)
2
0 0 ... bn−1 an − z
(3.4.23)
Since π0 (z) = 1, each zero of πk (z) is an eigenvalue of the matrix above. Thus,
πk (z) = det(zI − T̃k ) where T̃k denotes the upper-left k × k submatrix of T
(compare with Tk and dk (z) = det(zI − Tk )).
Thus, given µ, the entries of T are obtained from “top to bottom”. However,
given T , the τ -function is the limit of τ -functions −dk (z)/dk+1 (z) computed
‘bottom-to-top’.
48 CHAPTER 3. JACOBI MATRICES AND TRIDIAGONAL ENSEMBLES
Q
n β−1
where u
j=1 j Du refers to the joint density for n independent χβ random
variables, normalized so that the sum of their squares is one. In particular, Λ
and u are independent.
Theorem 32 follows from a computation of the Jacobian of the spectral map
n−1
S : Jac(n) → W n × S+ .
n−1
Theorem 33. The volume forms on Jac(n) and W n × S+ are related by
n n−1 n
!
Y Y Y
DT = daj bn−k−1
k dbk = Cn 4(Λ)DΛ uj Du. (3.5.2)
j=1 k=1 k=1
Proof. 1. Recall that Λ(l) denotes the diagonal matrix of eigenvalues of Tl and
Ql (l)
that dl (x) = j=1 (x − λj ). Therefore, we have the identity
l l−1 l l−1
(l) (l−1) (l) (l−1)
Y Y Y Y
λj − λk = dl−1 λj = dl λ k . (3.5.4)
j=1 k=1 j=1 k=1
2. The coefficients u2j are the residue of τn (z) at the poles λj , i.e.
dn−1 (λk )
u2k = , 1 ≤ k ≤ n. (3.5.5)
d0n (λk )
Observe also that
Y n
Y
d0n (λk ) = (λj − λk ), and d0n (λk ) = 4(Λ)2 . (3.5.6)
j6=k k=1
Therefore,
n n Qn−1 2(n−k)
Y 1 Y b
u2j = 2
|dn−1 (λk )| = k=1 k 2 . (3.5.7)
j=1
4(λ) 4(λ)
k=1
50 CHAPTER 3. JACOBI MATRICES AND TRIDIAGONAL ENSEMBLES
Hermite-1 Jac(n)
Proof of Theorem 33. 1. Our goal is to compute the Jacobian of the spectral
mapping S,
∂(T (a, b))
DT = DΛDu, (3.5.8)
∂(Λ, u)
where Du is uniform measure on {u ∈ Rn | |u| = 1, uj > 0 for all j}. Rather
than compute the change of variables directly, we will compute the push-forward
n−1
of GOE onto Jac(n) and W n × S+ separately, and obtain the Jacobian above,
see Figure 3.5.1.
2. Consider the push-forward of GOE under the map M 7→ (Λ, u), where
M = QΛQT is the diagonalization of M , with the normalization that the first
non-zero entry in each column is positive. Since Λ and the matrix of eigenvalues
Q are independent, Λ and u = QT e1 are independent. Since Q is distributed
according to Haar measure on O(n), the vector u is uniformly distributed on
n−1
S+ and the push-forward of GOE is the measure
1 2
p(Λ, u)DΛDu = cn e− 4 Tr(Λ) 4(Λ)DΛDu. (3.5.9)
3. Next consider the push-forward of GOE under the map M 7→ T , where
M = QT QT denotes the tridiagonalization of M . As we have seen in the proof
of Theorem 20, T and U are independent, and the marginal distribution of T is
given by
n n−1
1 2 Y Y
p̃(T ) DT = Cn e− 4 Tr(T ) daj bn−k−1
k dbk . (3.5.10)
j=1 k=1
n n−1
4. Since T ∈ Jac(n) and (Λ, u) ∈ W × S+ are in bijection, we have
∂(T (a, b))
p(Λ, u) = p̃(T (Λ, u)) . (3.5.11)
∂(Λ, u)
3.5. JACOBIANS FOR TRIDIAGONAL ENSEMBLES 51
We compare the expressions in (3.5.9) and (3.5.10) and use Lemma 12 to obtain
Qn−1
∂(T (a, b)) Cn k=1 bk
= . (3.5.12)
cn nj=1 uj
Q
∂(Λ, u)
Proof of Theorem 32. The law of We change variables using the spectral map-
ping and Theorem 33 to obtain the following identity for the law of the Hermite−
β ensembles
n−1
β 2 Y (β−1)(n−k)
Cn,β e− 4 Tr(T )
bk DT (3.5.13)
k=1
n
−β 2 Y
= Cn,β e 4 Tr(Λ )
4(Λ)β DΛ uβ−1
j
Du. (3.5.14)
j=1
Since the distribution factors, Λ and u are independent with the laws stated in
Theorem 32.
Exercises
3.1. Let w ∈ Rn have iid N (0, 1) components. Show that |w| and w/|w| are
independent.
3.2. Let U ∈ O(n) be a random orthogonal matrix. For example U could
be a Householder reflection associated to a random vector w. Then assume
A ∼ GOE. Show that B := U AU T ∼ GOE and B is independent of U .
3.3. Write a numerical code to sample matrices from both GOE and the Hermite−
1 ensemble. Verify numerically that a suitably normalized density of eigenval-
ues for the GOE matrix approaches the semicircle law as n increases (n = 100
should be ample). Is the same true for the Hermite − 1 ensemble? Why or why
not?
3.10. The proofs of Dumitriu and Edelman finesse the following issue: given
T ∈ Jac(n) it requires some care to find a decomposition for the tangent space
TT Jac(n), especially the isospectral manifold, MT , that is analogous to Lemma
2. As in that lemma, we may split TT Jac(n) into orthogonal subspaces that
correspond to diagonal matrices Λ̇ and QT Q̇ ∈ o(n). However, while each
QT Q̇ ∈ o(n) generates a curve in TT Symm(n) , not all QT Q̇ give rise to curves
in TT Jac(n). Verify this. Explore this issue further by trying to find a basis for
the isospectral manifold MT (see equation (3.4.3)).
3.6 Notes
To include in improved version.
1. Tridiagonal matrices as weights in enumerations problems.
Determinantal formulas:
From Vandermonde to
Fredholm
1 1 Pn 2
P (n) (x1 , . . . , xn ) = 4(x)2 e− 2 k=1 xk . (4.1.1)
Zn
53
54 CHAPTER 4. DETERMINANTAL FORMULAS
and the integral on the right hand side of (4.1.5) appears m! times when in-
tegrating over the complementary n − m variables for each choice of indices
{j1 , . . . , jm } ∈ J1, nKm . We state the following theorem which is proved in the
following sections.
Theorem 35. The joint density and m-point functions for GUE(n) are
1
P (n) (x1 , . . . , xn ) = det (Kn (xj , xk )1≤j,k≤n ) , (4.1.6)
n!
(n)
Rm (x1 , . . . , xm ) = det (Kn (xj , xk )1≤j,k≤m ) , (4.1.7)
Remark 36. The kernel Kn may be simplified using identities for the Hermite
polynomials. The Christoffel-Darboux formula (B.2.6) allows us to write
∞
X
det (I − zKn 1S ) = Am (S)(1 − z)m , z ∈ C, (4.1.11)
m=0
The calculations above would apply to any set of monic polynomials of degree
0, 1, 2, . . . , n − 1. The Hermite polynomials and wave functions are relevant
because they satisfy the orthogonality relations
2
e−x /2
Z
hj (x)hk (x) √ dx = δjk k!, (4.2.2)
R 2π
and allow the inclusion of an exponential weight. Precisely, the Hermite wave-
functions
2
1 e−x /4
ψk (x) = √ hk (x) , (4.2.3)
k! (2π)1/4
56 CHAPTER 4. DETERMINANTAL FORMULAS
and form a basis for L2 (R). Let H denote the matrix with entries Hjk =
ψj−1 (xk ). It follows from (4.2.1) and (4.2.3) that
x2
e− 2 4(x)2 ∝ det H 2 = det H T H = det [Kn (xj , xk )]1≤j,k≤n , (4.2.5)
where sgn(σ) denotes the sign of the permutation σ. We then evaluate the
integral
Z Z 2
det(H)2 dx1 . . . dxn = det [ψj−1 (xk )]1≤j,k≤n dx1 . . . dxn
Rn Rn
X Z n
Y
= sgn(σ)sgn(τ ) ψσj −1 (xj )ψτj −1 (xj ) dx1 . . . dxn
σ,τ ∈S(n) Rn j=1
(4.2.8)
X n
Y X
= sgn(σ)sgn(τ ) δσj ,τj = 1{σ=τ } = n! .
σ,τ ∈S(n) j=1 σ,τ ∈S(n)
We combine (4.2.8) and (4.2.6) to obtain the first assertion in Theorem 35:
1
P (n) (x1 , . . . , xn ) = det [Kn (xj , xk )]1≤j,k≤n .
n!
The formulas for the correlation functions may be obtained by induction,
beginning with
We first prove formula (4.3.1) in the case m = 0. The probability that all
58 CHAPTER 4. DETERMINANTAL FORMULAS
where ρnj (x1 , . . . , xn ) is the j-th symmetric function in n variables. For example,
n
X n
X n
Y
ρn0 (x) = 1, ρn1 (x) = xj , ρn2 (x) = xj xk , ρnn (x) = xj .
j=1 j<k j=1
Using
Qm the m-point correlation function, we obtain using (4.1.3) with f (x1 , . . . , xm ) =
j=1 1S (xj ),
Z
E (Nf ) = ρnj (1S (x1 ), . . . , 1S (xn ))P (n) (x1 , . . . , xn ) dx1 . . . dxn (4.3.3)
Rn
Z
1
= det [Kn 1S (xk , xl )]1≤k,l≤j dx1 . . . dxj .
j! Rj
In the last equality, we have used (4.1.7) and multiplied the kernel on the left
and right by the diagonal matrix dS = diag(1S (x1 ), . . . , 1S (xj )), so that
(n) (n)
1S (x1 ) . . . 1S (xj )Rj (x1 , . . . , xj ) = 12S (x1 ) . . . 12S (xj )Rj (x1 , . . . , xj )
= det dS [Kn (xk , xl )]1≤k,l≤j dS = det [Kn 1S (xk , xl )]1≤k,l≤j ,
= det(I − Kn 1S ), (4.3.4)
using the infinite series (C.1.8) for the Fredholm determinant (only n terms are
non-zero, since K (n) has rank n, see Exercise 4.2).
We now turn to the case m ≥ 1. Equation (4.3.2) must now be modified to
allow exactly m eigenvalues within S and n − m eigenvalues outside S. Define
m
Y n
Y
f (x1 , . . . , xn ) = 1S (xj ) (1 − 1S (xj )).
j=1 j=m+1
4.4. SCALING LIMITS OF INDEPENDENT POINTS 59
Then from (4.1.3), when we take into account the m! permutations of the first
m elements, and the (n − m)! permutations of the last n − m elements
1
Am (S) = E (Nf )
m!(n − m)!
Z
1
= f (x1 , . . . , xn )Rn(n) (x1 , . . . , xn )dx1 · · · dxn .
m!(n − m)! Rn
We then write
m
Y m−n
X
f (x1 , . . . , xn ) = 1S (xj ) (−1)k ρn−m
k (1S (xm+1 ), . . . , 1S (xm+k ))
j=1 k=0
We use the fact that ρn−m (1S (xm+1 ), . . . , 1S (xn )) is given by a sum of n−m
k k
terms, each of which is product of k terms, and all terms integrate to the same
value. So,
Z m
Y
1S (xj )ρn−m
k (1S (xm+1 ), . . . , 1S (xm+k ))Rn(n) (x1 , . . . , xn )dx1 · · · dxn
Rn j=1
Z m+k Z
n−m Y
= 1S (xj ) Rn(n) (x1 , . . . , xm+k )dxm+k+1 · · · dxn
k Rm+k j=1 Rn−m−k
× dx1 · · · dxm+k
m+k
(n − m)!
Z
(k)
Y
= 1S (xj )Rm+k (x1 , . . . , xm+k )dx1 · · · dxm+k
k! R m+k
j=1
(n − m)!
Z
= det (Kn 1S (xj , xl )1≤j,l≤m+k ) dx1 · · · dxm+k .
k! Rm+k
√
Consider an iid vector Λ = n(λ1 , λ2 , . . . , λn )T ∈ Rn where P(λj ∈ S) =
p (x0 )dx0 . We form the empirical measure
R
S sc
n
1X
Ln (dx) = δΛk (dx), (4.4.1)
n
k=1
But, it is clear, and effectively by definition, that ELn (dx0 ) = p(x0 )dx0 =
0 √ √
1
√ psc √
n
x
n
dx0 and hence np( nx0 )dx0 = psc (x0 )dx0 .
Next, we consider a gap probability in the “bulk”. Lets ∈ (−2, 2),
I ⊂R
√ I
be an interval and consider the rescaled interval In = n s + npsc (s) . Then
by independence
Z 0 n
1 x
P ( no λj ∈ In ) = 1 − √ psc √ dx0 . (4.4.3)
n In n
From
√ this we see a (trivial) scaling limit of the density of states after rescaling
by 1/ n, gaps on the order of 1/n after this rescaling and a largest “eigenvalue”
√
that satisfies λ̂ ∼ 2 n+ξn1/6 for an appropriate random variable ξ. All of these
statements carry over to the random matrix setting, but the actual limits are
very different for local statistics.
4.5. SCALING LIMITS I: THE SEMICIRCLE LAW 61
by Theorem 35 and equation (4.1.3). The value of the kernel Kn on the diag-
onal is determined by the Christoffel-Darboux relation (4.1.9) and L’Hospital’s
lemma: √
Kn (x, x) = n ψn0 (x)ψn−1 (x) − ψn (x)ψn−1
0
(x) . (4.5.4)
The scaling limit of ELn is the semicircle law defined in (1.2.1).
Lemma 13.
1 √ √
lim √ Kn x n, x n = psc (x), x ∈ R. (4.5.5)
n→∞ n
Further, for any ε > 0, the convergence is uniform on the set {x ||x − 2| ≥ ε }.
Proof. The lemma follows from the Plancherel-Rotach asymptotics for the Her-
mite wave functions (see Cases 1 and 2 and equations (B.5.1)–(B.5.4)) in Ap-
pendix B). Define the rescaled wave functions
1 √
Ψn+p (x) = n 4 ψn+p (x n), p = −2, −1, 0. (4.5.6)
We now use the asymptotic relations (B.5.2) and (B.5.4) depending on whether
|x| < 2 or |x| > 2. Since the region |x| > 2 corresponds to exponential decay
with a rate proportional to n, we focus on the region |x| < 2. In order to simplify
notation, let
1 1 π
θ = n ϕ − sin 2ϕ − ϕ − . (4.5.8)
2 2 4
62 CHAPTER 4. DETERMINANTAL FORMULAS
(This is the argument of the cosine in (B.5.17) when p = −1.) Then (4.5.7) and
(B.5.2) yield
1 √ √
√ Kn x n, x n
n
1 1 p
cos2 θ − cos(θ + ϕ) cos(θ − ϕ) =
∼ 4 − x2 ,
π sin ϕ 2π
using x = 2 cos ϕ and the trigonometric formulae cos 2α = 2 cos2 α − 1 and
2 cos(θ+ϕ) cos(θ−ϕ) = cos 2ϕ+cos 2θ. A similar calculation with (B.5.4) shows
that the limit vanishes outside the set |x| > 2. The assertion of uniformity in the
convergence follows from the assertion of uniform convergence in the Plancherel-
Rotach asymptotics.
Using Exercise 4.9, Lemma 13 implies that ELn (dx), after rescaling, con-
verges weakly
n
1X
δxk /√n (dx) → psc (x)dx, weakly. (4.5.9)
n
k=1
1√ √ √
Proof of Theorem 38. Let K̃n (x, y) denote the rescaled kernel psc (r) K (x n, y n),
n n
x = r − npscs (r) , y = r − npsct (r) . It follows from Lemma 14, using Sections C.2.1
and C.2 that
lim det I − z K̃n 1S = det (I − zKsine 1S ) , z ∈ C, (4.6.11)
n→∞
64 CHAPTER 4. DETERMINANTAL FORMULAS
and there exists a function G(x, y) ∈ L2 ([C, ∞)2 ) for all C ∈ R such that
√ x √
1 y
1 Kn 2 n + 1 ,2 n + 1 ≤ G(x, y). (4.7.5)
n6 n6 n6
Proof. Convergence follows from (B.5.6). The function G(x, y) can be con-
structed using (B.5.45) and (B.5.46), see Exercise 4.3.
√
Letting L → ∞ √ gives convergence in probability for λn / n. Similar arguments
follow for λ1 / n. Next, define
λn λ1
E,n = √ − 2 ≤ , √ + 2 ≤ .
n n
Exercises
4.1. Prove the Christoffel-Darboux identity (B.1.7) for Hermite polynomials.
(This is a standard relation and it is easy to find a proof in many texts, but try
to do it on your own.)
4.2. Show that
Z
det[K(xp , xq )]1≤p,q≤k dx1 · · · dxk = 0, (4.9.1)
Rk
(b) Numerically compute the Hermite wave functions for large n, say n =
100, and compare the rescaled wave function with the Plancherel-Rotach
asymptotic formulas in all three regions (oscillatory, decaying and transi-
tion).
4.7. Use the method of steepest descent to establish the Plancherel-Rotach
asymptotics in the region of exponential decay (equation (B.5.4)). This requires
more care than Q.2.
4.8. Establish the following a priori bound on the Airy kernel. For any a ∈ R,
Let S be the semi-infinite interval (a, ∞). Use the above estimate to establish
that the Fredholm determinant det(I − zKAiry 1S ) is an entire function.
4.9. Let ρn (x), n = 1, 2, . . . be probability densities on R that converge al-
most uniformly to ρ(x) with respect to Lebesgue measure on R. Assume ρ has
compact support. Show that
Z Z
lim f (x)ρn (x)dx = f (x)ρ(x)dx
n→∞ R R
4.10 Notes
To include in improved version.
69
70 CHAPTER 5. THE EQUILIBRIUM MEASURE
formula (Theorem 15), the equilibrium density (5.1.2) is precisely the joint law
of the eigenvalues for these ensembles at β = 1, 2 and 4 respectively. It is in
this sense that the ‘eigenvalues repel’.
We have scaled the energy V with n in (5.1.1) in order to obtain a simple
description of the scaling limit when n → ∞. In order to study this limit, we
view the energy function as a functional of the empirical measure, Ln , rather
than a configuration x ∈ Rn . For (r, s) ∈ R2 let
1 1 1
e(r, s) = V (r) + V (s) + log , (5.1.4)
2 2 |r − s|
that takes into account all interaction terms in I[µ], except the singular self-
interaction term from I[µ]. The logarithmic singularity in e(r, s) is integrable if
µ(ds) has an absolutely continuous density. Thus, if the particles in the log-gas
spread out sufficiently as n → ∞, we expect that µ has a smooth density, and
1
lim log Zn,V (β) = min I[µ]. (5.1.9)
n→∞ n2 µ
for every function f ∈ C(R). This topology is natural, because it yields com-
pactness by Helly’s theorem: Each sequence {µk }∞
k=1 ∈ PR has a subsequence
that converges weakly to a measure in PR .
Theorem 41. Assume V is a continuous function on [−R, R]. There exists a
unique probability measure µ∗ ∈ PR such that
Thus, µ∗ is a minimizer.
72 CHAPTER 5. THE EQUILIBRIUM MEASURE
is also weakly continuous. Finally, since each continuous function f ∈ C([−R, R]2 )
may be uniformly approximated by polynomials, the quadratic functional
Z R Z R
µ 7→ f (r, s)µ(dr) µ(ds)
−R −R
is weakly continuous.
The function e(s, t) defined in (5.1.4) is not continuous on [−R, R]2 since the
logarithmic term is unbounded on the diagonal s = t. However, for any M > 0,
the truncated function eM (r, s) = min(e(r, s), M ) is continuous. Thus, given a
weakly convergent sequence of measures {µk }∞ k=1 with limit µ ∈ PR we find
Z R Z R Z R Z R
eM (r, s)µ(dr) µ(ds) = lim eM (r, s)µk (dr) µk (ds)
−R −R k→∞ −R −R
Z R Z R
≤ lim inf e(r, s)µk (ds)µk (ds) = lim inf I[µk ].
k→∞ −R −R k→∞
We let M → ∞ on the left hand side and use the monotone convergence theorem
to obtain (5.2.3).
Proof. This proof is from [7, p.142]. We first regularize the logarithm at 0 and
use the following integral representation. For any real s and ε > 0
Z ∞
eisu − 1
log(s2 + ε2 ) = log ε2 + 2 Im e−εu du. (5.2.16)
0 iu
We now let ε ↓ 0 and use the monotone convergence theorem to obtain (5.2.14)
Theorem 44. Assume V (s) is a continuous function such that l(s) is bounded
below and l(s) → ∞ as |s| → ∞.
Proof. (a) Since V is bounded below and the addition of a constant to V does
not change the minimization problem, we may assume that l(s) ≥ 0. Then
1 1 1
e(r, s) = log + V (r) + V (s) ≥ l(r) + l(s) ≥ 0, (5.2.23)
|r − s| 2 2
On the other hand, the inequalities (5.2.26) and (5.2.27) yield the opposite
inequality
2(lM − I[µ∗ ])ν(SM ) ≤ 0. (5.2.29)
Thus, ν(SM ) = 0 for all M such that lM > I[µ∗ ].
5.3. FEKETE POINTS 77
The set Fn is called the set of n-Fekete points. The Fekete points are natu-
rally connected to the minimization problem for the functional I[µ] through the
modified functional H[Ln ], where Ln (x) is the empirical measure associated to
a point x ∈ Rn . Let δn denote the rescaled energy of Fekete points
1
δn = E(x(n) ). (5.3.2)
n(n − 1)
The main result is then the following
Theorem 45. Assume V satisfies the assumptions of Theorem 44. Let {x(n) }∞
n=1
be a sequence of points x(n) ∈ Fn and Then
(a) The rescaled energy of Fekete points increases monotonically to I[µ∗ ].
0 ≤ δn ≤ δn+1 ≤ I[µ∗ ]. (5.3.3)
Let µ(ds) be any probability measure on the line. We integrate (5.3.4) with
respect to the n-fold tensorized probability measure µ ⊗ µ · · · ⊗ µ on Rn to
obtain
E(x(n) ) (5.3.5)
Z n n
1 1
X X
≤ (V (sj ) + V (sk )) + log µ(ds1 )µ(ds2 ) · · · µ(dsn )
Rn 2 |sj − sk |
j,k=1 j6=k=1
Z Z
= n(n − 1) e(r, s)µ(ds)µ(dr) = I[µ],
R R
since for each value of the indices j and k only the integrals over µ(dsj ) and
µ(dsk ) give contributions that are not unity and there are n(n − 1) possible
unordered pairings of j and k. In particular, E(x(n) ) ≤ n(n − 1)I[µ∗ ].
78 CHAPTER 5. THE EQUILIBRIUM MEASURE
since the second term is the energy E(x̂) of the point x̂ ∈ Rn obtained from
x(n) by projecting out the coordinate xm .
Since m is arbitrary, we take the product over 1 ≤ m ≤ n + 1 to obtain
2
n(n+1)
1 (n+1) Y Y V (xj ) V (xm )
e− n E(x )
≤ e−(n−1)δn |xj − xm | e− 2 − 2
1≤m≤n+1 1≤j≤n+1,j6=m
2
− E(x(n+1) )
= e−(n−1)δn e n(n+1) . (5.3.7)
We now let M → ∞ and use the monotone convergence theorem and the fact
that µ∗ is a minimizer to obtain
I[µ∗ ] ≥ I[µ] ≥ I[µ∗ ]. (5.3.9)
Since µ∗ is unique, it follows that µ∗ = ν.
This argument proves that every subsequential limit of L(n) is µ∗ . Thus, the
entire sequence converges to µ∗ .
5.4 Exercises
The first three questions are related. The goal is to formulate and analyze the
equation for the equilibrium measure µ∗ associated to the potential V (x). In
order to simplify your calculations, assume that µ∗ has a continuous density ψ,
in all the problems below. The last two questions discuss enumeration problems
related to the Catalan numbers.
1. Basics of the Hilbert transform. Let G(z) denote the Stieltjes transform
Z ∞ Z ∞
1 1
G(z) = µ∗ (ds) = ψ(s)(ds), z ∈ C\supp(µ∗ ). (5.4.1)
−∞ s − z −∞ s − z
The Hilbert transform of ψ is the limit of the Stieltjes transform as z → x ∈ R.
The Hilbert transform also differs from the Stieltjes transform by the inclusion
of a factor of π (since this makes the Fourier transform of the operator H
particularly simple). That is, given µ∗ as above, we set
Z ∞ Z ∞
1 ψ(s) x−s
Hψ(x) = p.v. ds := lim ψ(s) ds. (5.4.2)
π −∞ x − s ε→0 −∞ (x − s)2 + ε2
(a) Show that Hψ is a bounded function when ψ(x) is continuous.
(b) Show that µ∗ may be recovered from G by evaluating the jump in the
imaginary part of G across the support of µ∗ :
1
lim (G(x + iε) − G(x − iε)) = ψ(x). (5.4.3)
ε→0 2πi
(c) Compute the Hilbert transform of the following functions to obtain a feel
for it (answers are on wikipedia):
eix , δ0 (x), 1[a,b] (x).
3ga4 + a2 − 1 = 0. (5.4.9)
5.5 Notes
To include in improved version.
1. Fixed point equation for equilibrium measure.
2. Properties of Hilbert transform.
3. Convergence of k-point distribution to tensor product of equilibrium mea-
sure.
Chapter 6
In this chapter we discuss other random matrix ensembles that differ funda-
mentally from GUE, GOE and GSE. For this discussion we concentrate on real
and complex matrices. The first ensembles we consider are the real and com-
plex Ginibre ensembles1 , GinR (m, n) on Rm×n and GinC (m, n) on Cm×n . These
are ensembles of real and complex matrices of size m × n. without symmetry
conditions. Their densities are given by
1 1 T 1 1 ∗
pGin,R (Y )DY = e− 4 Tr Y Y DY, pGin,C (X)DX = e− 2 Tr X X DX.
ZR,m,n ZC,m,n
Thus, the entries are distributed as independent (real or complex) normal ran-
dom variables. The definition DY and DX in each case follows directly from the
volume forms associated to the length elements Tr(dY T dY ) and Tr(dX ∗ dX).
When m = n we use the notation GinC (n) and GinR (n) and ZR,n and ZC,m .
The Ginibre ensembles allow us to define the Laguerre ensembles as transfor-
mations of GinC (m, n) and GinR (m, n). These are ensembles of positive (semi-)
definite matrices defined by X ∗ X where X ∼ GinC (m, n), GinR (m, n). The La-
guerre ensembles are often referred to as Wishart matrices and they get their
name from the close connection to Laguerre polynomials.
We end this chapter with a discussion of the so-called Jacobi ensembles. It
is important to note that these ensembles are not ensembles of Jacobi matrices,
rather, they get their name from their close connection to Jacobi polynomials.
Jacobi polynomials are polynomials orthogonal on the interval [−1, 1], and so
Jacobi matrices have eigenvalues that all lie in the same interval [−1, 1].
1 Often, the term Ginibre ensemble is reserved for square matrices, but we find it convenient
81
82 CHAPTER 6. OTHER RANDOM MATRIX ENSEMBLES
Y Q R
Figure 6.1.1: The full QR decomposition in the case m > n. The shaded area
columns and rows are removed to create the reduced QR decomposition.
(6.1.4)
for X 0 and Y 0 independent of X and Y , respectively. This can be performed
even if X and Y are deterministic matrices. So, in the real case, and similarly
in the complex case,
0 0 0 In
Y 7→ Y Y = QR 7→ QR := QR = Y.
0
Since it is a non-classical theorem for the Schur decomposition, we state the
following.
Theorem 48. Let X(t), X : (−a, a) → Fn×n , a > 0, be a C k matrix func-
tion. Assume X(0) has distinct eigenvalues. Then the induced factors X(t) 7→
(T (t), U (t)) or X(t) 7→ (S(t), O(t)) obtained by the Schur decomposition for
F = C or R are also C k in a neighborhood of t = 0.
84 CHAPTER 6. OTHER RANDOM MATRIX ENSEMBLES
Proof. We use induction on n for the real case. The complex case is similar. If
n = 1, then a Gaussian vector in Rn is non-zero with probability one. If n > 1,
n ≤ m − 1, assume
Solve
Y x = b = QRx = b, Rx = QT b =: b̃,
and therefore b̃ ∼ GinR (m, 1). For this equation to have a solution x, Rx = b̃,
since R ∈ Rm×n , triangular, and n < m, the last entry of b̃ must vanish. Thus
P rank Y b < m + 1 | Y = 0
P (∃λ ∈ C, v ∈ Cn , Xv = λv and v1 = 0) = 0,
P (∃λ ∈ C, v ∈ Rn , Y v = λv and v1 = 0) = 0.
Proof. We prove this for Y . The proof for X is similar. First, we write
y0 y1T
Y = ,
y2 Y 0
y0 ∼ GinR (1), y1 , y2 ∼ GinR (n − 1, 1), Y 0 ∼ GinR (n − 1, n − 1),
n2 n2
Now consider a rectangle R = [a, b] ⊂ R , and assume that
Z
1{Y ∈Rn×n | |4(Λ)|=0} DY > 0.
R
Since the set of matrices with distinct eigenvalues is dense, 4(Λ) 6= 0 for some Y .
But the only way for the zero locus of a polynomial in n variables to have positive
n-dimensional Lebesgue measure is for the polynomial to vanish identically. The
theorem follows.
86 CHAPTER 6. OTHER RANDOM MATRIX ENSEMBLES
TX Cn×n ∼
= Rn(n−1) ⊕ P TI U(n).
The first inequality orders entries by which diagonal they lie on. The second
orders within the diagonal. Then
ζ 7→ [S− , T+ ] =: M− ζ (6.1.7)
where Λ̃ζ S− is defined through ζ S− 7→ [S− , Λ], which is diagonal, Sjk 7→ (λk −
λj )Sjk . M− and D are matrices whose exact form is irrelevant. Decomposing
all differentials into real and imaginary parts and computing the metric tensor
Tr dX ∗ dX,
Q
we find (6.1.8) by using det(Λ̃ + M− ) = j<k (λk − λj ) and computing the
associate volume form. Here one has to use that if A : Cn → Cn induces B :
R2n → R2n (by separating real and imaginary parts), then det B = | det A|2 .
Theorem 54. The Schur decomposition of GinC (n) is given by
1 − 1 Tr T ∗ T
pGin,C (X)DX = e 2 |4(Λ)|2 DT DU. (6.1.9)
ZC,n
Note that this implies that the strict upper-triangular entries of T are all iid
complex normal random variables.
X = UT OX = U 0 T 0 .
88 CHAPTER 6. OTHER RANDOM MATRIX ENSEMBLES
From the uniqueness of the QR factorization (on set of full probability where X
is invertible), T = T 0 and U = OT U 0 . But U and U 0 have the same distribution
and this distribution must therefore be invariant under left multiplication by
any element of U(m). We conclude U is distributed according to Haar measure
on U(m) [27] and to proportionality constants:
ñ
β ∗ QR β ∗ Y 2m−2j+1
e− 4 Tr X X
DX −→ e− 4 Tr T T
Tjj DT D Ũ , ñ = min{m, n},
j=1
Theorem 55. Assume that Y has exactly k real eigenvalues. Assume further
that the real part of all the eigenvalues of Y (0) = Y in the closed upper-half plane
are distinct. Finally, assume that each 2×2 block in the real Schur factorization
has γ 6= δ in (6.1.1). Then
TY Rn×n ∼
= Rn(n−1)/2 ⊕ o(n).
Proof. Assume Y (t) is a smooth curve in Rn×n such that Y (t) has k real eigen-
values for all t. As before, we have the relation
We need to show that the entries of Ṡ and Ȯ are uniquely determined by this
relation. We assume
R1 × · · · ··· ×
0 R2 × · · · ··· ×
.. ..
.. ..
. . . .
αj −γj
S= 0 · · · 0 R` × · · · × , Rj =
,
0 ··· ··· 0 δj αj
λ1 · · · ×
. .. ..
..
. .
0 ··· · · · 0 λk
6.1. THE GINIBRE ENSEMBLES 89
by noting that
[A, S] = [LG (A), LG (S)] + [DG (A), LG (S)] + [UG (A), LG (S)]
+ [LG (A), DG (S)] + [DG (A), DG (S)] + [UG (A), DG (S)]
+ [LG (A), UG (S)] + [DG (A), UG (S)] + [UG (A), UG (S)],
LG (S) = 0, and any term involving only DG and UG or only UG does not con-
tribute to LG ([A, S]). Then, it is a non-trivial but straightforward calculation
to find that LG ([DG (A), DG (S)]) = 0. This gives a linear system of equations
for LG (A). Since it will be of use in computing the metric tensor below, we
compute the determinant of this matrix in the following lemma.
Lemma 20. There exists a trivial mapping LG (A) → ξ ∈ Rn(n−1)/2−` defined
by ordering the elements of LG (A) so that when M is the matrix representation
for ξ 7→ LG ([A, S]) we have
(1) (2)
Y Y Y
det M = 4k (Λ) := (λj − λi ) 4ij 4ij
1≤i<j≤k 1≤j<k≤` 1≤i≤k,1≤j≤`
where λ1 , . . . , λk are the real eigenvalues, µj = αj + iβj , βj > 0 are the complex
eigenvalues (in the upper half plane) and
(1)
4ij = |µj − µi |2 |µj − µ̄i |2 = |µj − µi |2 |µ̄j − µi |2 ,
(2)
4ij = |µj − λi |2 .
Proof of Lemma 20. The important aspect of this is to choose the ordering.
First split
(1,1)
A 0
LG (A) = .
A(2,1) A(2,2)
in the same was as for LG (A). From the reasoning2 that went into (6.1.7), we
have that the (i, j) block of L(1,1) depends only on blocks (i0 , j 0 ) of A(1,1) for
(i0 , j 0 ) > (i, j) and entries in A(2,1) . Similarly, the (i, j) entry of L(2,2) depends
only on entries (i0 , j 0 ) of A(2,2) for (i0 , j 0 ) > (i, j) and entries in A(2,1) . Lastly,
one checks that block (i, j) of L(2,1) depends only on blocks (i0 , j 0 ) of A(2,1) for
(i0 , j 0 ) > (i, j). This gives a strong form of strict lower-triangularity for ξ 7→ L.
We now show that ξ 7→ K := LG (LG (A), DG (S)) is block-diagonal in a way
that does not overlap with this strict lower-triangularity. First, decompose K
into K (i,j) , i = 1, 2, j = 1, 2 in the same was as for LG (A) and L. We obtain
the following relations for blocks of size 2 × 2, 1 × 1 and 1 × 2, respectively:
(1,1) (1,1) (1,1)
Kij = Aij Rj − Ri Aij ,
(2,2) (2,2)
Kij = Aij (λj − λi ),
(2,1) (2,1) (2,1)
Kij = Aij Rj − λi Aij .
From this lemma, with our assumptions, we can uniquely find LG (A). But
as A is skew-symmetric, we have ` entries left undetermined. So, we consider
(OT Ẏ )2j,2j = (Ṡ + [A, S])2j,2j = (α̇j + (γj − δj )ṡ2j+1,2j ) + f2j (LG (A)),
(OT Ẏ )2j+1,2j+1 = (Ṡ + [A, S])2j+1,2j+1 = (α̇j + (δj − γj )ṡ2j+1,2j ) + f2j (LG (A)).
(6.1.10)
for some functions fj . As LG (A) is known, this gives a solvable system for α̇j
Q`
and s2j+1,2j , with determinant 2` j=1 (γj − δj ). The remaining entries of Ṡ are
given through the relation
Ṡ = OT Ẏ O − [A, S].
depends on entries (i0 , j 0 ) of the lower-triangular matrix for j 0 ≤ j with i = i0 and i0 ≥ i with
j = j 0 . With strict triangularity, fewer dependencies occur.
6.2. SVD AND THE LAGUERRE (WISHART) ENSEMBLES 91
where
`
Y k
Y Y
DS = dαj dγj dδj dλj ds, (6.1.12)
j=1 j=1 s∈UG (S)
Theorem 57. The real Schur decomposition of GinR (n) given k real eigenvalues
is
`
2` − 1 Tr S T S Y
pGin,R,k (Y )DY = (k) e 4 |4k (Λ)| |γj − δj | DS DO. (6.1.14)
ZR,n j=1
Note that this implies that the generalized upper-triangular entries of S are
all iid normal random variables.
Y = QΣOT , X = U ΣV ∗ ,
A = LLT
L−1 T −T
1 L = L1 L , where L−T = (L−1 )T .
Since the non-singular upper- and lower-triangular matrices for groups, the left-
hand (right-hand) side is lower-triangular (upper-triangular). Therefore L−1 1 L
is a diagonal matrix that is equal to its own transpose-inverse: ej L−1
1 Lej = ±1.
Positivity of the diagonal entries gives L1 = L. Now, by Gaussian elimina-
tion, without pivoting3 A = L̃U where L̃ is lower-triangular and U is upper-
triangular. Here L̃ has ones on the diagonal. We know that eTj Aej > 0
and therefore eTj L̃U ej = Ujj > 0. Then Let Ud = diagonal(U )1/2 and A =
L̃Ud Ud−1 U . It follows from the symmetry of A that L = L̃Ud gives the Cholesky
factorization. Similar considerations follow for A ∈ Cm×n .
This is a well-defined, invertible mapping, provided that the first row of V con-
tains non-vanishing entries. It will follow from Section 6.2.2 that the probability
of this is one. But we emphasize that for this decomposition X 6= U ΛV ∗ , gen-
erally. We now show that if X ∼ GinC (m, n) then U, Λ, V are independent and
we then characterize the distribution of Λ and V .
Lemma 21 (Spectral variables for Her+ (n)). If A ∈ Her+ (n) is non-singular
with distinct eigenvalues then
TA Her+ (n) ∼
= Rn ⊕ P TI U(n).
Proof. The proof is essentially the same as Lemma 6, just using that the set of
strictly positive definite matrices is open.
We define DA in the natural way as the volume form induced by the metric
tensor Tr dA2 . We then have the analogous formula to Theorem 15:
DA = |4(Λ)|2 DΛ DU.
Next, we compute the volume form associated with the change Cholesky change
of variables.
Lemma 22. Let A = LL∗ be the Cholesky decomposition for a non-singular
A ∈ Her+ (n). Let DL be the natural volume form induced by Tr(dL∗ dL). Then
n
2(n−j)+1
Y
DA = 2n Ljj DL.
j=1
Examine the structure of these matrices. Since ej eTk L∗ is the matrix that con-
tains the kth row of L∗ in its jth row, with all other row being zero we find the
following picture
0
..
.
0
L kk
∂A Lk+1,k
= .. .
∂ReLjk .
Lj−1,k
0 · · · 0 Lkk Lk+1,k · · · Lj−1,k 2ReLjk · · ·
..
.
94 CHAPTER 6. OTHER RANDOM MATRIX ENSEMBLES
Here only the jth row and jth column have non-zero entries. Here 2ReLjk is
∂A
in the (j, j) entry. A similar picture holds for ∂ImL jk
, with 2ImLjk in the (j, j)
entry. We define a mapping ReL 7→ ξ ∈ Rn(n−1)/2 and ImL 7→ η ∈ Rn(n−3)/2 by
the ordering of the non-zero elements of L:
This orders first by row, and then by columns within each row. Assume (i, `) <
(j, k), j ≥ k, i ≥ `. Then
∂Ai` ∂Ai`
= 0, = 0.
∂ReLjk ∂ImLjk
∂Ajk ∂Ajk
= Lkk , j > k, = 2Lkk , j = k,
∂ReLjk ∂ReLjk
∂Ajk
= Lkk , j > k.
∂ImLjk
This theorem allows one to understand transformations of GinC (m, n). Fol-
lowing the transformation (6.2.1), with X ∈ Cm×n with m ≥ n using T = L∗
noting that
T̃
T = .
0
Here DŨ is Haar measure on U(n) and DV represents the same distribution as
the eigenvectors of GUE(n). Also, DΣ is Lebesgue measure on Rn+ . As noted
6.2. SVD AND THE LAGUERRE (WISHART) ENSEMBLES 95
below (6.2.1), this is not the singular value decomposition for X, but we claim,
it is in a distributional sense. For X ∼ GinC (m, n), m ≥ n and consider
X = U1 ΣV, X̃ := U ΣV
where (U, V, Σ) are independent with joint distribution (6.2.4), U1 is the matrix
of left singular vectors for X, and U is independent of U1 . Then X̃ = U U1∗ X,
but then by the invariance of U , for measureable sets S1 ⊂ U(m), S2 ⊂ Cm×m ,
P(U U1∗ ∈ S1 ) = P(U ∈ S1 U1 ) = P(U ∈ S1 ),
P(U U1∗ ∈ S1 ,X ∈ S2 ) = P(U ∈ S1 U1 , X ∈ S2 )
Z Z
= DU pGin,C (X)DX = P(U ∈ S1 )P(X ∈ S2 ).
S2 S1 U1
So, U U1∗ is independent of X and therefore X̃ must have the same distribution
as X. This implies the singular value decomposition of GinC (m, n) is given by
(6.2.4).
Remark 60. If one wants to match of dimensions, then DU should be replaced
by the push-forward of uniform measure on SCm−1 × SCm−2 × · · · × SCm−n−1 onto
U(m) via Householder reflections.
We now consider the reduction of GinC (m, n) and GinR (m, n) to bidiagonal
matrices and, in the process, find a generalization of (6.2.5) to general β. This
is sometimes called Golub–Kahan bidiagonalization. The aim here is not to
preserve eigenvalues, but to preserve singular values as transformations are per-
formed. So, we can perform independent Householder reflections from the left
and the right. Recall the definition of P (Y ) from (6.1.2). Let Y ∼ GinR (m, n)
for m ≥ n. Consider the iterative method
Y0 = Y,
Ỹ1 = Q1 Y0 := P (Y0 )Y0 ,
!
1 0
Y1T = Q̃1 Ỹ1T := Ỹ1T ,
0 P [Ỹ1T ]n−1,m−1
1 0
Ỹ2 = Q2 Y1 := Y ,
0 P ([Y1 ]m−1,n−1 ) 1 (6.2.6)
..
.
Ij−1 0
Ỹj = Qj Yj−1 Y ,
0 P ([Yj−1 ]m−j+1,n−j+1 ) j−1
!
Ij 0
YjT = Q̃j Ỹj−1
T
:= T
Ỹj−1 .
0 P [ỸjT ]n−j,m−j
where Dω̃l and Dωp denote uniform measure on SRl and SRp , respectively. Simi-
larly, by applying this algorithm to X ∼ GinC (m, n) we find
n n−1 n−2 n−1
2(m−j)+1 2(n−k)−1
Y Y Y Y
DX ∝ cj dcj dk ddk Dω̃k Dωp ,
j=1 k=1 l=1 p=1
where Dω̃l and Dωp denote uniform measure on SCl and SCp , respectively.
from (6.1.8)
n
1 1
Pn 2 Y
P̂ (n) (z1 , . . . , zn )Dz = |4(Z)|2 e− 2 j=1 |zj | dRe zj dIm zj .
Zn j=1
(n)
where R̂m is the m-point correlation function defined by (4.1.4) with P̂ (n)
instead of P (n) and dRe zj dIm zj instead of dxj . To show that this is the correct
choice for K̂n and to determine cj we need to show that {Φj }n−1 j=0 are orthogonal
and choose cj > 0 to normalize the functions. Consider for j < k
Z Z
1 2
Φj (z)Φk (z) dRez dImz = cj c̄k z̄ k−j |z|2j e− 2 |z| dRez dImz
C C
Z ∞ Z 2π
1 2
= cj c̄k (cos θ + i sin θ) dθ rk+j+1 e− 2 |r| dRez dImz = 0.
k−j
0 0
If j = k we find
Z Z
1 2
|Φj (z)|2 dRez dImz = |cj |2 |z|2j e− 2 |z| dRez dImz,
C C
√
and using r = 2s
Z Z ∞ Z ∞
2j − 12 |z|2 2j+1 − 21 r 2
|z| e dRez dImz = 2π r e j+1
dr = 2 π sj e−s ds
C 0 0
j+1
=2 πΓ(j + 1) = 2j+1 j!π
so
1 1
cj = √ , cj √ √ = cj+1 .
2j/2+1/2 πj! 2 j+1
So, we find a simple two-term recurrence formula
z 1
Φj+1 (z) = √ √ Φj (z), Φ0 (z) = √ .
2 j+1 2π
The corresponding Christoffel-Darboux-type formula is
R∞
where Γ(n, z) = z tn−1 e−t dt is the incomplete Gamma function. To see this
let fn (z) = ez Γ(n, z), and we find
fn(j) (0) = (n − 1)!, j = 0, 1, 2, . . . , n − 1,
fn(j) (0) = 0, j ≥ n,
so that
n−1
X (n − 1)! j
fn (z) = z .
j=0
j!
Then
Z ∞ Z ∞ Z ∞
tn−1 e−t dt = nn tn−1 e−nt dt = nn t−1 e−ng(t) dt,
n|z|2 |z|2 |z|2
where g(t) = t − log t. The stationary phase point here is t = 1, g 0 (1) = 0 and
g 00 (1) = 1. So, if |z| ≤ 1 − , the stationary phase point is in the interval of
integration and
Z ∞ √ √
nn t−1 e−ng(t) dt = e−n nn−1/2 2π(1 + O(n−1 )) = e−n nn−1 2πn(1 + O(n−1 ))
|z|2
= (n − 1)!(1 + O(n−1 ))
uniformly as n → ∞ by Stirling’s approximation. Then for |z| ≥ 1 + , by
integrating by parts
Z ∞
1 2n−2 −n|z|2 n − 1 ∞ n−2 −nt
Z
n−1 −nt
In (z) := t e dt = |z| e + t e dt
|z|2 n n |z|2
1 2 n−1
≤ |z|2n−2 e−n|z| + In (z).
n n
Therefore
2
In (z) ≤ |z|2n−2 e−n|z| .
From these estimates, the following lemma follows.
6.2. SVD AND THE LAGUERRE (WISHART) ENSEMBLES 99
uniformly.
This shows that (see Exercise 4.9)
1
ELn (Dz) → 1{|z|≤1} Dz
π
weakly. This is the averaged circular law.
Marchenko–Pastur law
Again, consider X ∼ GinC (m, n) or Y ∼ GinR (m, n), m ≥ n, and consider the
sample covariance matrices X ∗ X/m and Y T Y /m, and let x1 , . . . , xn be their
unordered eigenvalues. Define the empirical spectral measure
n
1X
Ln (dx) = δx (dx).
n j=1 j
Assume further that n/m → d ∈ (0, 1]. The Marchenko–Pastur law states that
r
1 |(λ+ − x)(x − λ− )|1[λ− ,λ+ ] √
ELn (dx) → pMP (x; d)dx := 2
dx, λ± = (1 ± d)2 ,
2πd x
weakly as n → ∞. If d = 0, then the limiting law is δ0 (dx), by the Law of Large
Numbers.
100 CHAPTER 6. OTHER RANDOM MATRIX ENSEMBLES
Chapter 7
101
102 CHAPTER 7. SAMPLING RANDOM MATRICES
Chapter 8
Additional topics
103
104 CHAPTER 8. ADDITIONAL TOPICS
Appendix A
There are several different conventions for the definition of the Airy function.
The standardization adopted here follows [1]. The Airy function, Ai(x) is defined
as the oscillatory integral
∞ Z b
t3
Z 3
1 1 t
Ai(x) = cos + xt dt = lim cos + xt dt. (A.1.1)
π 0 3 π b→∞ 0 3
This is an improper integral, that is, the integral converges conditionally, not
absolutely. In order to obtain an absolutely convergent integral, it is necessary
to work in the complex plane. Let C denote a contour in the complex plane
that starts and ends at the point at infinity, and is asymptotically tangent to
the rays e−iπ/3 and e+iπ/3 respectively. Then first setting t = −iz and then
deforming the contour, we have
Z ∞ 3 Z
z3
1 1
i z3 −xz i −xz
Ai(x) = e dz = e 3
dz. (A.1.2)
2πi −∞ 2πi C
3
i z3 −xz 3 3
e ≤ e|x|r e−r cos(3θ)/3
∼ e−r /3 r|x|
e (A.1.3)
105
106 APPENDIX A. THE AIRY FUNCTION
y 00 = xy, x ∈ C. (A.2.2)
A.3 Asymptotics
The functions Ai(x) and Bi(x) have the following asymptotic properties.
Asymptotics as x → ∞.
1
2 3 e−ζ x4
ζ = x2 , Ai(x) ∼ 1√ , Bi(x) ∼ √ eζ . (A.3.1)
3 π
2x 4 π
Asymptotics as x → −∞.
2 3 1 π 1 π
ζ= (−x) 2 , Ai(x) ∼ 1√ sin ζ + , Bi(x) ∼ 1√ cos ζ + .
3 4 4
x4 π x4 π
(A.3.2)
Appendix B
Hermite polynomials
1
Z
x2 √
√ hk (x)hl (x)e− 2 dx = 2πk!δkl . (B.1.3)
2π R
k−1
X 1 (hk (x)hk−1 (y) − hk−1 (x)hk (y))
hj (x)hj (y) = , x 6= y. (B.1.7)
j=0
j! (k − 1)!(x − y)
107
108 APPENDIX B. HERMITE POLYNOMIALS
Z
ψk (x)ψl (x) dx = δkl . (B.2.2)
R
√ √
xψk (x) = k + 1ψk+1 (x) + kψk−1 (x). (B.2.3)
x √
ψk0 (x) = − ψk (x) + kψk−1 (x). (B.2.4)
2
1 x2
ψk00 (x) + k+ − ψk (x) = 0. (B.2.5)
2 4
n−1
X √ (ψn (x)ψn−1 (y) − ψn−1 (x)ψn (x))
ψk (x)ψk (y) = n . (B.2.6)
x−y
k=0
√
(−1)n n
x 1
lim n
h2n √ = √ cos x. (B.3.1)
n→∞ 2 n! 2n π
r
(−1)n
x 2
lim h2n+1 √ = sin x. (B.3.2)
n→∞ 2n n! 2n π
1/4 n x 1
lim (2n) (−1) ψ2n √ = √ cos x. (B.3.5)
n→∞ 2n π
x 1
lim (2n)1/4 (−1)n ψ2n+1 √ = √ sin x. (B.3.6)
n→∞ 2n π
The asymptotic formulas (B.3.1) and (B.3.2) are proved by applying Laplace’s
method to the integral formula (B.1.2). We only explain how to prove (B.3.1)
since equation (B.3.2) is similar. Since (i)2n = (−1)n , we take the real part
of (B.1.2) to find
2 x2 ∞ 2n − ξ2
r Z
x xξ
(−1)2n h2n √ = e 4n ξ e 2 cos √ dξ
2n π 0 2n
1 Z
2n+1 nn+ 2 ∞ −n(t2 −2 log t)
= √ e cos xt dt, (B.3.7)
π 0
√
by rescaling ξ = n t. We now apply Laplace’s method to the integral above.
The function g(t) = t2 − 2 log t has a single minimum on the interval (0, ∞) at
t = 1. At this point
where f and Φ are entire functions. Assume Φ(t∗ ) = 0, Φ0 (t∗ ) = 0, Φ00 (t∗ ) 6= 0,
ImΦ(t) = 0 for t ∈ Γ. Further assume Γ is the path of steepest ascent for Φ,
i.e. the path of steepest descent for −Φ(t). Having ImΦ(t) = 0 is enough to
B.4. STEEPEST DESCENT FOR INTEGRALS 111
ensure that Γ is either the path of steepest ascent (locally) or steepest descent:
Let t = x(s) + iy(s) be a smooth local parameterization of Γ, then by the
Cauchy–Riemann equations
d dx dy dx dy
0= Im Φ(t) = Im Φx (t) + Im Φy (t) = −Re Φx (t) + Re Φx (t) .
ds ds ds ds ds
This shows that ∇Re Φ is orthogonal to the tangent vector (−y 0 (s), x0 (s)), im-
plying that Γ is in the direction of greatest increase/decrease for Re Φ.
Performing a Taylor expansion, we have
Φ00 (t∗ )
Φ(t) = (t − t∗ )2 (1 + O(|t − t∗ |). (B.4.2)
2
The point is that Φ is locally quadratic at t∗ and we use this to inform the
change of variables. But if we naı̈vely looked to solve
Φ(t∗ + v) = s2 ,
for v as a function of s, v(0) = 0, we would fail. The implicit function theorem
fails because we have two solution branches! Instead we consider
Φ(t∗ + sv) Φ00 (t∗ ) 2
2
−1=0= v − 1 + O(|sv 3 |). (B.4.3)
s 2
00 ∗
We can choose v = ±R−1/2 e−iφ/2 where Φ 2(t ) = Reiφ . For either choice, we can
apply the implicit function theorem (the derivative with respect to v, evaluated
at (s, v) = (0, ±R−1/2 e−iφ/2 ) does not vanish). We use v = ±R−1/2 e−iφ/2 to
obtain v(s), and our local parameterization of Γ: t(s) = t∗ + sv(s). We use
this a change of variables, within a neighborhood B(t∗ , ) on which the implicit
function theorem applies (here we assume the orientation of Γ is the same as
the induced orientation on t((−δ1 , δ2 )))
Z Z δ2
2 ds
f (t)e−nΦ(t) dt = f (t∗ + sv(s))e−ns , δ1 , δ2 > 0.
∗
Γ\B(t ,) −δ1 v(s) + sv 0 (s)
(B.4.4)
For n ≥ 1, we have
Z Z
2 2
|f (t)|e−n(Φ(t)−δ ) dt ≤ |f (t)|e−(Φ(t)−δ ) dt := M. (B.4.6)
Γδ Γδ
2
And therefore (B.4.5) is exponentially small in n, less than M e−nδ . Now,
consider
Z δ
2
f (t∗ + sv(s))e−ns (v(s) + sv 0 (s)) ds (B.4.7)
−δ
112 APPENDIX B. HERMITE POLYNOMIALS
and we can directly apply Laplace’s method. Taylor expand the function
f (t∗ + sv(s))(v(s) + sv 0 (s))
at s = 0, and term by term integration gives an expansion in powers of n−1/2
with the leading order term being
Z
f (t)e−nΦ(t) dt
Γ
Z δ
2
= f (t∗ + sv(s))e−ns (v(s) + sv 0 (s))ds + O(n−α )
−δ
Zδ
2
= f (t∗ )v(0)(1 + O(s))e−ns ds + O(n−α ), for all α > 0. (B.4.8)
−δ
√
Performing a change of variables s = y/ 2n we have
Z δ Z √2nδ r
−ns2
−y 2 /2 π
e ds = √ e dy = + O(n−α ), for all α > 0,
−δ − 2nδ n
Z δ Z √2nδ
−ns2 1 2 C
|s|e ds = √ √ |y|e−y /2 dy = + O(n−α ), for all α > 0.
−δ 2n − 2nδ n
So, we have
Z r
−nΦ(t) 2π
f (t)e dt = f (t∗ )|Φ00 (t∗ )|−1/2 e−iφ/2 + O(n−1 ) as n → ∞. (B.4.9)
Γ n
√
n1/12−p/2 ψn (x n) (B.5.6)
s
n! −1/3 1 0 −2/3
= Ai(s) + n − p Ai (s) + O(n ) .
(n + p)! 2
(B.5.7)
t− t+ π
4 − ϕ
2 Γ
0 (0, ∞)
Figure B.5.1:
(b) |x| > 2. The two critical points lie on the imaginary axis, and may be
written in the form
√ !
x ± x2 − 4
t± = i = i sgn(x)e±ϕ , (B.5.12)
2
Thus, we have
Z ∞ Z
p −ng(t) −ng(t+ )
In,p (x) = t e dt = e tp e−n(g(t)−g(t+ )) dt
0 Γ
Z ∞
dt n 00 2
= e−ng(t+ ) tp+ e− 2 |g (t+ )|s ds + O(n−1 ). (B.5.14)
ds t+ −∞
B.5. PLANCHEREL–ROTACH ASYMPTOTICS 115
t+
t−
0 (0, ∞)
Figure B.5.2:
In the second line, we have used the fact that Im(g(t) − g(t+ )) = 0 on Γ, and we
have further approximated the integral over Γ by an integral over the tangent
to Γ at t+ . More precisely, the approximation here is
which implies
dt ϕ
= ei( 4 − 2 ) .
π
(B.5.15)
ds t+
Finally, since x is real, we have In,p (x) = In,p (−x). We combine (B.5.9) with
(B.5.16) to obtain
√
r
n+p 2 n cos 2ϕ 1 1 π
hn+p (x n) ∼ n 2 e2 cos n ϕ − sin 2ϕ + p + ϕ− ,
sin ϕ 2 2 4
(B.5.17)
where x and ϕ are related via (B.5.1). We now use (B.2.1) and Stirling’s ap-
proximation (B.3.4) to obtain (B.5.2).
The asymptotics in case (b) are obtained as follows. We make the change of
variables (B.5.3), and deform the domain of integration for In,p to the contour
consisting of two straight lines shown in Figure B.5.2. So see that this is enough,
we have that g 00 (t+ ) > 0 while g 00 (t− ) < 0. So the path of steepest ascent
116 APPENDIX B. HERMITE POLYNOMIALS
(descent for −g) through t− is the imaginary axis. Then, the steepest ascent
path through t+ makes a right angle with the imaginary axis. And one can check
that the real part of g(t) is strictly increasing along the contour t+ + i[0, ∞).
The dominant contribution comes from t− . The remaining calculations are left
to the reader. The final asymptotic relation is
n
√ n+p e− 2
e(p+ 2 )ϕ− 2 (sinh(2ϕ)−2ϕ) (1 + o(1)),
1 n
hn+p (x n) = n 2 √ (B.5.18)
sinh ϕ
which combines with (B.2.1) and Stirling’s approximation (B.3.4) to yield (B.5.4).
We now turn to case (c). We begin with the integral representation(B.5.8)
and substitute
r √ s
t=i+ 1 , x=2 n+ 1 , (B.5.19)
n3 n6
moving the integral over R to an integral over the line i + R, to obtain
1 ∞
√ √
Z
n6
hn (x n) = (−i n)n √ enh(r) dr, (B.5.20)
2π −∞
where
2
r 1 r s
h(r) = log i + 1 − i+ 1 −i 2+ 2
n3 2 n3 n3
2
1 s 1 i s − 43 4
= + log i + 2 + isr + r3 + 4 + O(n r ), (B.5.21)
2 n 3 n 3 2n 3
using the Taylor expansion of the logarithm. The terms that depend on s may
be pulled out of the integral and we are left with
n 1
√ n 2 + 6 n sn 13 ∞ isr+ i r3
Z
hn (x n) ≈ √ e2e e 3 dr (B.5.22)
2π −∞
√ n 1 n
1
hn (x) = 2πn 2 + 6 e 2 esn 3 (Ai(s) + O(n−1/3 ))
To make this rigorous, and to obtain the next term in the expansion, We take
the integral
√
r Z
n 1 2
hn+p (x) = (−i n)n+p tp e−n( 2 (t−ix) −log t)
dt (B.5.23)
2π R
√
r Z
n 1 2
hn (x) = (−i n)n+p (i + r)p e−n( 2 (r+i(1−x)) −log(i+r))
dr. (B.5.24)
2π R
Then this can be deformed to a contour Γ = e−iπ/6 (−∞, 0] ∪ eiπ/6 [0, ∞).
B.5. PLANCHEREL–ROTACH ASYMPTOTICS 117
so that
r4
Z Z
n 2 π r3
e− 2 (1−x) (i + r)p e−nH(r) dr = ein 2 ip + pip−1 r − nip ein(x−2)r+in 3 dr
Γ Γ∩B(0,δ) 4
+O(n−1 ).
Continuing, we obtain
s
√ −1/12+p/2 n! s2 n1/3
ψn+p (x n) = n e 4 (B.5.34)
(n + p)!
1 (4)
× Ai(s) + n −1/3
−pAi0 (s) − Ai (s) + O(n −2/3
) (B.5.35)
4
s
n!
= n−1/12+p/2 (B.5.36)
(n + p)!
1 2
× Ai(s) + n−1/3 −pAi0 (s) + (s Ai(s) − Ai(4) (s)) + O(n−2/3 )
4
(B.5.37)
s
−1/12+p/2 n!
=n (B.5.38)
(n + p)!
1
× Ai(s) + n−1/3 − p Ai0 (s) + O(n−2/3 ) , (B.5.39)
2
where we used Ai(4) (s) = s2 Ai(s) + 2Ai0 (s) in the last line.
Then, we deform
Z Z
2 2
−n −nH(r) −n
e 2 (1−x) e dr = e 2 (1−x) e−nH(r) dr (B.5.40)
Γ∩B(0,δ) C
n1/3 r3
ein(x−2)r+in
R
Then, define fn (s) = 2π C
3 dr and we have
s
√ n! s2 n−1/3 s2 n1/3
√
n1/12−p/2 ψn+p (x n) − e 4 fn (s) ≤ M n−2/3 e 4 −nδ/ 2(x−2) .
(n + p)!
(B.5.44)
√ s2 n1/3
√ 3 1/3
Choosing δ = 2, we find e 4 −nδ/ 2(x−2) ≤ e−s 4 n . A similar estimate
follows for ψn+p and we obtain that there exist a constant M > 0 such that for
0 ≤ s ≤ n2/3
√ s2 n−1/3 3 1/3
n1/12 ψn (x n) − e 4 fn (s) ≤ M n−2/3 e−s 4 n ,
(B.5.45)
√ s2 n−1/3 3 1/3
n1/12 ψn−1 (x n) − e 4 fn (s) ≤ M n−2/3 e−s 4 n .
Fredholm determinants
C.1 Definitions
Our purpose in this section is to explain the notion of a Fredholm determi-
nant and resolvent in a simple and concrete setting. The ideas presented here
originated in Fredholm’s attempt to find a solution formula akin to Cramer’s
rule for linear integral equations. The notion of a determinant for an infinite-
dimensional linear operator is, of course, of independent interest and has at-
tracted the interest of many mathematicians. Simon’s book provides an excel-
lent overview of current knowledge [31].
Assume a given continuous kernel K : [0, 1] × [0, 1] → R and a continuous
function h : [0, 1] → R. Fix a spectral parameter z ∈ C and consider the linear
integral equation
Z 1
ϕ(x) − z K(x, y)ϕ(y) dy = h(x), x ∈ [0, 1]. (C.1.1)
0
The integral equation (C.1.1) may be written in the more compact form
(I − zK)ϕ = h, (C.1.2)
where I − zK denotes the bounded linear operator on L2 ([a, b]) defined by
Z b
ϕ 7→ (I − zK)ϕ, (I − zK)ϕ(x) = ϕ(x) − z K(x, y)ϕ(y) dy x ∈ [a, b].
a
(C.1.3)
Integral equations such as (C.1.1) may naturally be viewed as continuum
limits of linear equations. More precisely, we fix a positive integer n, consider
a uniform grid xj = j/n, with uniform weights wj = 1/n, define the vector
(n) (n)
hj = h(xj ), matrix Kj,k = wj K(xj , xk ), 1 ≤ j, k ≤ n and discretize (C.1.1)
by the linear equation
n
(n) (n) (n) (n)
X
ϕj −z Kj,k ϕk = hj , 1 ≤ j ≤ n. (C.1.4)
k=1
121
122 APPENDIX C. FREDHOLM DETERMINANTS
(n) det(Mjk )
Rj,k = (−1)j+k , (C.1.5)
det(In − zK (n) )
where Mjk denotes the matrix obtained from In − zK (n) by removing the j-
th row and k-th column. Further, if zj , j = 1, . . . n, denote the zeros of the
polynomial det(In − zK (n) ), the eigenvalues of K (n) are given by 1/zj . Both
these notions may be extended to (C.1.1) via the Fredholm determinant. The
basic observation that allows passage to the limit is the identity
(−z)k
Z
det (K(xp , xq )1≤p,q≤k ) dx1 . . . dxk . (C.1.7)
k! [0,1]k
We have assumed that K is bounded on [0, 1] × [0, 1], say max |K| ≤ M < ∞.
By the inequality above,
(−z)k
Z
det [K(xp , xq )]1≤p,q≤k dx1 · · · dxk
k! [0,1]k
k k/2 1 1
≤ (|z|M )k ∼ √ (|z|M e)k k+1 ,
k! 2π k 2
where we have used Stirling’s approximation in the last step.
Remark 64. If [0, 1] is replaced by a general Borel set S, we assume
for m ≥ 1. Recall that the zeros of a non-zero entire function form a discrete,
countable set. The entire function det(I − λ−1 K) is an infinite-dimensional gen-
eralization of the characteristic polynomial of the matrix K (n) in the following
sense:
Theorem 65 (Eigenvalues of K). Assume that K is a continuous kernel. The
complex number λ is an eigenvalue of K if and only if D(λ−1 ) = 0.
For more on Fredholm determinants, see [23, Ch.24].
C.2 Convergence
Suppose a kernel Kn (x, y) → K∞ (x, y), (x, y) ∈ S 2 , pointwise. One needs the
additional convergence criteria to conclude
The following are from [31]. Let Kn and K∞ be the operators on L2 (S) with
kernels Kn and K∞ , respectively. Then the trace norm of an operator K is
given by
√
kKkTr = Tr K∗ K, (C.2.2)
√
where K∗ is the adjoint of K. The general definition of K∗ K for general
operators is unimportant for us and an operator with finite trace norm is said
124 APPENDIX C. FREDHOLM DETERMINANTS
for each x, y ∈ S and there exists G(x, y) ∈ L2 (S 2 ) such that |Kn (x, y)| ≤
G(x, y). This allows one to use the dominated convergence theorem.
Notation
D.1 Indices
The integers m and n are reserved for the number of rows and columns of a
matrix. For square matrices, we use√n. The letters j and k are used to denote
indices. The letter i is reserved for −1.
D.2 Fields
R real numbers.
√
C complex numbers with imaginary unit i = −1.
n
X
|x|2 = |xj |2 . (D.2.1)
j=1
125
126 APPENDIX D. NOTATION
D.3 Matrices
The fundamental spaces of matrices are denoted as follows:
Jac(n)
[8] I. Dumitriu and A. Edelman, Matrix models for beta ensembles, Journal
of Mathematical Physics, 43 (2002), pp. 5830–5847.
129
130 BIBLIOGRAPHY
[12] H. M. Edwards, Riemann’s zeta function, vol. 58, Academic Press, 1974.
[14] J. Harer and D. Zagier, The euler characteristic of the moduli space of
curves, Inventiones Mathematicae, 85 (1986), pp. 457–485.
[21] T. Kato, Perturbation theory for linear operators, Springer, New York,
NY, 1995.
[24] B. M. McCoy and T. T. Wu, The two-dimensional Ising model, vol. 22,
Harvard University Press, 1973.
[27] L. Nachbin, The Haar integral, R.E. Krieger Pub. Co, 1976.
[31] B. Simon, Trace ideals and their applications, vol. 120 of Mathematical
Surveys and Monographs, American Mathematical Society, Providence, RI,
second ed., 2005.
[32] R. P. Stanley, Enumerative combinatorics, vol. 2, Cambridge University
Press, 2011.
[33] D. W. Stroock, Probability theory: an analytic view, Cambridge univer-
sity press, 2010.
[34] G. ’t Hooft, A planar diagram theory for strong interactions, Nuclear
Physics B, 72 (1974), pp. 461–473.
[35] Tao, T, Topics in random matrix theory, AMS, Providence, RI, 2011.
[36] C. A. Tracy and H. Widom, Level–Spacing Distributions and the Airy
Kernel, Communications in Mathematical Physics, 159 (1994), pp. 151–
174.