Random Matrices and Random Partitions Normal Convergence, Volume 1 PDF
Random Matrices and Random Partitions Normal Convergence, Volume 1 PDF
com
Random Matrices
and Random Partitions
Normal Convergence
www.ebook3000.com
www.ebook3000.com
Random Matrices
and Random Partitions
Normal Convergence
Zhonggen Su
Zhejiang University, China
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TA I P E I • CHENNAI
www.ebook3000.com
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
www.ebook3000.com
www.ebook3000.com
v
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
www.ebook3000.com
March 3, 2015 13:52 9197-Random Matrices and Random Partitions ws-book9x6 page vii
Preface
vii
www.ebook3000.com
March 3, 2015 13:52 9197-Random Matrices and Random Partitions ws-book9x6 page viii
www.ebook3000.com
March 3, 2015 13:52 9197-Random Matrices and Random Partitions ws-book9x6 page ix
Preface ix
attention since it will also appear in the study of other similar models. In
addition to circular and Gaussian unitary ensembles, we shall consider their
extensions like circular β matrices and Hermite β matrices where β > 0 is a
model parameter. These models were introduced and studied at length by
Dyson in the early 1960s to investigate energy level behaviors in complex
dynamic systems. A remarkable contribution at this direction is that there
is a five (resp. three) diagonal sparse matrix model representing circular β
ensemble (resp. Hermite β ensemble).
In Chapters 4 and 5 we shall deal with random uniform partitions and
random Plancherel partitions. The study of integer partitions dates back
to Euler as early as the 1750s, who laid the foundation of partition theory
by determining the number of all distinct partitions of a natural number.
We will naturally produce a probability space by assigning a probability
to each partition of a natural number. Uniform measure and Plancherel
measure are two best-studied objects. Young diagram and Young tableau
are effective geometric representation in analyzing algebraic, combinatorial
and probabilistic properties of a partition. Particularly interesting, there
exists a nonrandom limit shape (curve) for suitably scaled Young diagrams
under both uniform and Plancherel measure. This is a kind of weak law
of large numbers from the probabilistic viewpoint. To proceed, we shall
further investigate the second-order fluctuation of a random Young diagram
around its limit shape. We need to treat separately three different cases:
at the edge, in the bulk and integrated. It is remarkable that Gumbel
law, normal law and Tracy-Widom law can be simultaneously found in
the study of random integer partitions. A basic strategy of analysis is
to construct a larger probability space (grand ensemble) and to use the
conditioning argument. Through enlarging probability space, we luckily
produce a family of independent geometric random variables and a family
of determinantal point processes respectively. Then a lot of well-known
techniques and results are applicable.
Random matrices and random partitions are at the interface of many
science branches and they are fast-growing research fields. It is a formidable
and confusing task for a new learner to access the research literature,
to acquaint with terminologies, to understand theorems and techniques.
Throughout the book, I try to state and prove each theorem using lan-
guage and ways of reasoning from standard probability theory. I hope
it will be found suitable for graduate students in mathematics or related
sciences who master probability theory at graduate level and those with
interest in these fields. The choice of results and references is to a large
www.ebook3000.com
March 3, 2015 13:52 9197-Random Matrices and Random Partitions ws-book9x6 page x
Zhonggen Su
Hangzhou
December 2014
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page xi
Contents
Preface vii
1. Normal Convergence 1
1.1 Classical central limit theorems . . . . . . . . . . . . . . . 1
1.2 The Stein method . . . . . . . . . . . . . . . . . . . . . . 19
1.3 The Stieltjes transform method . . . . . . . . . . . . . . 25
1.4 Convergence of stochastic processes . . . . . . . . . . . . . 29
xi
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page xii
Bibliography 261
Index 267
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 1
Chapter 1
Normal Convergence
1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 2
Later on, Laplace further extended the work of De Moivre to the case
p 6= 1/2 to obtain
Z b
Sn − np 1 2
P a≤ p ≤b ≈ √ e−x /2 dx. (1.3)
np(1 − p) 2π a
Formulas (1.2) and (1.3) are now known as De Moivre-Laplace central limit
theorem (CLT). p
Note ESn = np, V ar(Sn ) = np(1 − p). So (Sn − np)/ np(1 − p) is
a normalized random
√ variable with mean zero and variance one. Denote
−x2 /2
φ(x) = e / 2π, x ∈ R. This is a very nice function from the viewpoint
of function analysis. It is sometimes called bell curve since its graph looks
like a bell, as shown in Figure 1.1.
Normal Convergence 3
and invented together with Stirling the well-known Stirling formula (it ac-
tually should be called De Moivre-Stirling formula)
√
n! = nn e−n 2πn(1 + o(1)).
√
Setting k = n/2 + nxk /2, where a ≤ xk ≤ b, we have
S − n √
n 2
1 nn e−n 2πn(1 + o(1))
P 1√ = xk = n · √ p
2 n
2 k k e−k (n − k)n−k e−(n−k) 2πk 2π(n − k)
1 2
= √ e−xk /2 (1 + o(1)).
2πn
Taking sum over k yields the integral of the right hand side of (1.2).
Given a random variable X, denote its distribution function FX (x) un-
der P . Let X, Xn , n ≥ 1 be a sequence of random variables. If for each
continuity point x of FX ,
FXn (x) → FX (x), n → ∞,
d
then we say Xn converges in distribution to X, and simply write Xn −→ X.
In this terminology, (1.3) is written as
S − np d
p n −→ N (0, 1), n → ∞,
np(1 − p)
where N (0, 1) stands for a standard normal random variable.
As the reader may notice, the Bernoulli law only deals with frequency
and probability, i.e., Bernoulli random variables. However, in practice peo-
ple are faced with a lot of general random variables. For instance, measure
length of a metal rod. Its length, µ, is intrinsic and unknown. How do
we get to know the value of µ? Each measurement is only a realization of
µ. Suppose that we measure repeatedly the metal rod n times and record
Pn
the observed values ξ1 , ξ2 , · · · , ξn . It is believed that k=1 ξk /n give us
a good feeling of how long the rod is. It turns out that a claim similar to
the Bernoulli law is also valid for general cases. Precisely speaking, assume
that ξ is a random variable with mean µ. ξn , n ≥ 1 is a sequence of i.i.d.
Pn
copy of ξ. Let Sn = k=1 ξk . Then
Sn P
−→ µ, n → ∞. (1.4)
n
This is called the Khinchine law of large numbers. It is as important as the
Bernoulli law. As a matter of fact, it provides a solid theoretic support for
a great deal of activity in daily life and scientific research.
The proof of (1.4) is completely different from that of (1.1) since we do
not know the exact distribution of ξk . To prove (1.4), we need to invoke
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 4
Normal Convergence 5
d
Theorem 1.1. (i) Xn −→ X if and only if ψXn (t) → ψX (t) for any t ∈ R.
(ii) If ψXn converges pointwise to a function ψ, and ψ is continuous at
t = 0, then ψ(t) is a characteristic function of some random variable, say
d
X, and so Xn −→ X.
Having the preceding preparation, we can easily obtain the following CLT
for sums of independent random variables.
Normal Convergence 7
Theorem 1.3. Under Feller condition, the ξn satisfies the CLT, that is
Pn
Sn − k=1 µk d
√ −→ N (0, 1), n → ∞
Bn
if and only if Lindeberg condition holds.
Next turn to the martingale CLT. First, recall some notions and basic
properties of martingales. Assume that An , n ≥ 0 is a sequence of non-
decreasing sub-sigma fields of A. Let Xn , n ≥ 0 be a sequence of random
variable with Xn ∈ An and E|Xn | < ∞. If for each n ≥ 1
E(Xn |An−1 ) = Xn−1 a.e.
we call {Xn , An , n ≥ 0} a martingale. If An = σ{X0 , X1 , · · · , Xn }, we
simply write {Xn , n ≥ 0} a martingale. If {Xn , An , n ≥ 0} is a martingale,
setting dn := Xn −Xn−1 , then {dn , An , n ≥ 0} forms a martingale difference
sequence, namely
E(dn |An−1 ) = 0 a.e.
Conversely, given a martingale difference sequence {dn , An , n ≥ 0}, we can
form a martingale {Xn , An , n ≥ 0} by
Xn
Xn = dk .
k=1
Normal Convergence 9
The interested reader is referred to Hall and Heyde (1980) for many other
limit theorems related to martingales.
Let E be a set of at most countable points. Assume that Xn , n ≥ 0 is
a random sequence with state space E. If for any states i and j, any time
n ≥ 0,
P (Xn+1 = j|Xn = i, Xn−1 = in−1 , · · · , X0 = i0 )
= P (Xn+1 = j|Xn = i)
= P (X1 = j|X0 = i), (1.10)
then we call Xn , n ≥ 0 a time-homogenous Markov chain. Condition (1.10),
called the Markov property, implies that the future is independent of the
past given its present state.
Denote pij the transition probability:
pij = P (Xn+1 = j|Xn = i).
The matrix P := (pij ) is called the transition matrix. It turns out that
both the X0 and the transition matrix P will completely determine the
(n)
law of a Markov chain. Denote the n-step transition matrix P(n) = pij ,
(n)
where pij = P (Xn = j|X0 = i). Then a simple chain rule manipulation
shows
P(n+m) = P(n) P(m) .
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 10
τi = min{n ≥ 1; Xn = i}.
π = πP,
πi pij = πj pji ,
Normal Convergence 11
This is a type of law of large numbers for Markov chains. See Serfozo (2009)
for a proof.
Let L02 be the subspace of L2 (π) consisting of functions f : E 7→ R with
P
Eπ f := i∈E f (i)πi = 0. We shall give a sufficient condition under which
Pn−1
the linear sum Sn (f ) := i=0 f (Xi ) satisfies the CLT. To this end, we
introduce the transition operator TX
T g(i) = g(j)pij .
j∈E
Trivially, T g(i) = E g(X1 )X0 = i . Assume that there exists a function g
such that
f = g − T g. (1.11)
Then it easily follows from the martingale CLT
Sn (f ) d
−→ N 0, σf2
√
n
with limit variance
σf2 = kgk22 − kT gk22 ,
where k · k2 denotes the L2 -norm.
If f is such that
∞
X
T nf (1.12)
n=0
is convergent in L2 (π), then the solution of the equation (1.11) do exist.
P∞
Indeed, g = n=0 T n f just solves the equation.
It is too restrictive to require that the series in (1.12) is L2 -convergent.
An improved version is
Theorem 1.6. Let Xn , n ≥ 0 is an ergodic Markov chain with stationary
distribution π. Assume f ∈ L02 satisfies the following two conditions:
(i)
L
T n f −→
2
0, n → ∞;
(ii)
∞
X 1/2
kT n f k22 − kT n+1 f k22 < ∞.
n=0
Then we have
Sn (f ) d
−→ N 0, σf2
√
n
with limit variance
n
X
k
2
n+1
X
2
σf2 T k f
2 .
= lim
T f 2−
n→∞
k=0 k=0
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 12
and
( 2
√ 1 e−(log x) /2 (1 + sin(2π log x)), x > 0
pY (x) = 2π x
0, x ≤ 0.
Then it is easy to check that X and Y have all moments finite and EX k =
EY k for any k ≥ 1.
Normal Convergence 13
m2k+1 (X) = 0.
(ii) Assume X is a Poisson random variable with parameter λ, then for
k≥1
EX(X − 1) · · · (X − k + 1) = λk .
(iii) Assume X is a random variable with density function given by
1√
4 − x2 , |x| ≤ 2,
ρsc (x) = 2π (1.15)
0, otherwise,
then for k ≥ 1
1 2k
m2k (X) = k+1 k ,
m2k+1 (X) = 0.
(iv) Assume X is a random variable with density function given by
( q
1 4−x
x , 0 < x ≤ 4,
ρM P (x) = 2π (1.16)
0, otherwise,
then for k ≥ 1
1 2k
mk (X) = .
k+1 k
We remark that ρsc and ρM P (see Figures 1.2 and 1.3) are often called
Wigner semicircle law and Marchenko-Pastur law in random matrix litera-
ture. They are respectively the expected spectrum distributions of Wigner
random matrices and sample covariance matrices in the large dimensions.
It is now easy to verify that these moments satisfy the Carleman condition
(1.14). Therefore normal distribution, Poisson distribution, Wigner semi-
circle law and Marchenko-Pastur law are all uniquely determined by their
moments.
Normal Convergence 15
where λ runs through the set of all partitions of {1, 2, · · · , n}, B ∈ λ means
one of the blocks into which the set if partitioned and |B| is the size of the
block.
Since knowledge of the moments of a random variable is interchangeable
with knowledge of its cumulants, Theorem 1.7 can be reformulated as
τ1 (Xn ) → 0, τ2 (Xn ) → 1
and for k ≥ 3
τk (Xn ) → 0,
d
then Xn −→ N (0, 1).
To conclude this section, we will make a brief review about Lindeberg
replacement strategy by reproving the Feller-Lévy CLT. To this end, we
need an equivalent version of convergence in distribution, see Section 1.4
below.
Ef (Xn ) −→ Ef (X), n → ∞.
Normal Convergence 17
Then
S T
n n
Ef √ − Ef √
n n
n h
X 1 1 i
= Ef √ (Rn,k + ξk ) − Ef √ (Rn,k + ηk )
n n
k=1
Xn h 1 1 i
= Ef √ (Rn,k + ξk ) − Ef √ Rn,k
n n
k=1
Xn h 1 1 i
− Ef √ (Rn,k + ηk ) − Ef √ Rn,k . (1.18)
n n
k=1
√
Applying the Taylor expansion of f at Rn,k / n, we have by hypothesis
1 1
Ef √ (Rn,k + ξk ) − Ef √ Rn,k
n n
1 ξ 1 1 ξ2
k
= Ef 0 √ Rn,k √ + Ef 00 √ Rn,k k
n n 2 n n
3
1 ξ
+ Ef (3) (R∗ ) 3/2 k
6 n
1 1 1
= Ef 00 √ Rn,k + 3/2 Ef (3) (R∗ )ξk3 , (1.19)
2n n 6n
√ √
where R∗ is between (Rn,k + ξk )/ n and Rn,k / n.
Similarly, we also have
1 1
Ef √ (Rn,k + ξk ) − Ef √ Rn,k
n n
1 1 1
= Ef 00 √ Rn,k + 3/2 Ef (3) (R∗∗ )ηk3 , (1.20)
2n n 6n
√ √
where R∗∗ is between (Rn,k + ξk )/ n and Rn,k / n.
Putting (1.19) and (1.20) back into (1.18) yields
S T
n n
Ef √ − Ef √
n n
n
X 1 1
= Ef (3) (R∗ )ξk3 + 3/2 Ef (3) (R∗∗ )ηk3 .
6n3/2 6n
k=1
Noting kf (3) k∞ < ∞ and E|ξk |3 < ∞ and E|ηk |3 < ∞, we obtain
S T
n n
= O n−1/2 .
Ef √ − Ef √
n n
The assertion is now concluded.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 18
Proof of Theorem 1.2 To apply Theorem 1.9, we need to use the trun-
cation technique. For any constant a > 0, define
Hence by (1.21),
Pn ¯
k=1 (ξk (an ) − µk (an )) d
√ −→ N (0, 1). (1.22)
n
Remark 1.2. The Lindeberg replacement strategy makes clear the fact
that the CLT is a local phenomenon. By this we mean that the structure
of the CLT does not depend on the behavior of any fixed number of the
increments. Only recently was it successfully used to establish the Four
Moment Comparison theorem for eigenvalues of random matrices, which
in turn solves certain long-standing conjectures related to universality of
eigenvalues of random matrices. See Tao and Vu (2010).
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 19
Normal Convergence 19
The Stein method was initially conceived by Stein (1970, 1986) to provide
errors in the approximation by the normal distribution of the distribution
of the sum of dependent random variables of a certain structure. However,
the ideas presented are sufficiently abstract and powerful to be able to
work well beyond that purpose, applying to approximation of more general
random variables by distributions other than normal. Besides, the Stein
method is a highly original technique and useful in quantifying the error in
the approximation of one distribution by another in a variety of metrics.
This subsection serves as a basic introduction of the fundamentals of the
Stein method. The interested reader is referred to nice books and surveys,
say, Chen, Goldstein and Shao (2010), Ross (2011).
A basic starting point is the following Stein equation.
Lemma 1.2. Assume that ξ is a random variable with mean zero and vari-
ance σ 2 . Then ξ is normal if and only if for every bounded continuously
differentiable function f (kf k∞ , kf 0 k∞ < ∞),
Eξf (ξ) = σ 2 Ef 0 (ξ). (1.24)
Note
1 −z2 /2
1 − Φ(z) ∼ e , z → ∞.
z
< π/2 and kfz0 k∞ ≤ 2. By hypothesis, it
p
It is not hard to see kfz k∞
follows
P (ξ ≤ z) − Φ(z) = Efz0 (ξ) − Eξfz (ξ) = 0.
We now conclude the proof.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 20
where τk is the kth culumant of ξ, the remainder term admits the bound
1 + (3 + 2q)q+2
εq ≤ cq kf (q+1) kE|ξ|q+2 , cq ≤ .
(q + 1)!
As the reader may notice, if we replace the indicator function 1(−∞,z] by
a smooth function in the preceding differential equation (1.25), then its
solution will have a nicer regularity property. Let H be a family of 1-
Lipschitz functions, namely
H = h : R 7→ R, |h(x) − h(y)| ≤ |x − y| .
Normal Convergence 21
We omit the proof, which can be found in Chen, Goldstein and Shao (2010).
It is easily seen that for any random variable W of interest,
Eh(W ) − Eh(ξ) = Efh0 (W ) − EW fh0 (W ).
Given two random variables X and Y , the Wasserstein distance is defined
by
dW (X, Y ) = sup |Eh(X) − Eh(Y )|.
h∈H
To illustrate the use of the preceding Stein method, let us take a look at
the normal approximation of sums of independent random variables below.
If EW 2 = 1, then
r
1 E(|W 0 − W |2 |W ) 2 1 0 3
dW (W, ξ) ≤ √ E 1− + E W − W .
2π 2τ 3τ
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 23
Normal Convergence 23
Rx
Proof. Given f ∈ G, define F (x) = 0
f (t)dt. Then it obviously follows
EF (W 0 ) − EF (W ) = 0. (1.32)
= Ef (W )E W 0 − W |W
= −τ EW f (W ).
Hence we have
1 1
EW f (W ) = Ef 0 (W )(W 0 − W )2 + Ef 00 (W ∗ )(W 0 − W )3 .
2τ 6τ
Subtracting Ef 0 (W ) yields
1
EW f (W ) − Ef 0 (W ) = −Ef 0 (W ) 1 − E (W 0 − W )2 W
2τ
1
+ Ef (W )(W 0 − W )3 .
00 ∗
6τ
Thus it follows from the Cauchy-Schwarz inequality and the fact kf 0 k∞ ≤
p
2/π and kf 00 k∞ ≤ 2,
r r
2
EW f (W ) − Ef 0 (W ) ≤ 2 E 1 − 1 E (W 0 − W )2 W
π 2τ
1 0 3
+ E W − W .
3τ
The proof is complete.
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 24
and
n
1 1 X 2
E (Wn0 − Wn )2 Wn = + 2 E
ξi |Wn .
n n i=1
Hence we have
1 2
E 1− E (Wn0 − Wn )2 Wn
2τ
n
1 1 1 X 2 2
=E 1− + 2E ξi Wn
2τ n n i=1
n 2
1 X 2 2
= E (ξ i − Eξi )|Wn
4n2 i=1
n
1 X 2 2
≤ 2
E ξi − Eξi2
4n i=1
n n
1 X 2 2 2 1 X 4
≤ 2 E(ξi − Eξi ) ≤ 2 Eξ .
4n i=1 n i=1 i
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 25
Normal Convergence 25
Finally, note
3 1 3
E Wn0 − Wn = E ξI − ξI0
n3/2
n
8 1X
≤ 3/2
E|ξi |3 .
n n i=1
for any z outside the support of µ. In particular, it is well defined for all
complex numbers z in C \ R. Some elementary properties about sµ (z) are
listed below. Set z = a + iη.
(i)
sµ (z) = sµ (z).
So we may and do focus on the upper half plane, namely η > 0 in the
sequel.
(ii)
1
|sµ (z)| ≤ .
|η|
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 26
(iii)
Z
1
Im(sµ (z)) = Im(z) dµ(x).
R |x − z|2
So Im(sµ (z)) has the same sign as Im(z).
(iv)
1
sµ (z) = − (1 + o(1)), z → ∞.
z
(v) If mk (µ) := R xk dµ(x) exists and are finite for every k ≥ 0, then it
R
follows
∞
1 X mk (µ)
sµ (z) = − .
z zk
k=0
So sµ (z) is closely related to the moment generating function of µ.
(vi) sµ (z) is holomorphic outside the support of µ.
Example 1.7. (i) Let µ := µsc be a probability measure on R with density
function ρsc given by (1.15). Then its Stieltjes transform is
√
z z2 − 4
ssc (z) = − + . (1.33)
2 2
(ii) Let µ := µM P be a probability measure on R with density function
given by (1.16). Then its Stieltjes transform is
1 1p
sM P (z) = − + z(z − 4).
2 2z
Theorem 1.13. Let µ be a probability measure with Stieltjes transform
sµ (z). Then for any µ-continuity points a, b (a < b)
1 b sµ (λ + iη) − sµ (λ − iη)
Z
µ(a, b) = lim dλ
η→0 π a 2i
1 b
Z
= lim Im(sµ (λ + iη))dλ.
η→0 π a
Normal Convergence 27
d
Note that Zη −→ X as η → 0. This implies
lim P (Zη ∈ (a, b)) = P (X ∈ (a, b)).
η→0
The reader is referred to Anderson, Guionnet and Zeitouni (2010), Bai and
Silverstein (2010), Tao (2012) for its proof and more details.
To conclude, let us quickly review the Riesz transform for a continual
diagram. Let
2 √
u arcsin u2 + 4 − u2 , |u| ≤ 2,
Ω(u) = π (1.34)
|u|, |u| > 2
Normal Convergence 29
The reader is referred to Billingsley (1999a) for the proof and more details.
In addition, (ii) can be replaced by
(ii0 ) for any bounded infinitely differentiable function f ,
Ef (Xn ) → Ef (X), n → ∞.
It can even be replaced by
(ii00 ) for any continuous function f with compact support,
Ef (Xn ) → Ef (X), n → ∞.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 30
d
In the special cases S = R and Rk , Xn ⇒ X is equivalent to Xn −→ X. In
the case S = R∞ , Xn ⇒ X if and only if for each k ≥ 1
Xn ◦ πk−1 ⇒ X ◦ πk−1 , n→∞
where πk denotes the projection from R∞ to Rk .
The case of C[0, 1] is more interesting and challenging. Assume Xn ⇒
X. Then for any k ≥ 1 and any k points t1 , t2 , · · · , tk ∈ [0, 1]
Xn ◦ πt−1
1 ,··· ,tk
⇒ X ◦ πt−1
1 ,··· ,tk
, n→∞ (1.36)
where πt1 ,··· ,tk is a projection from C[0, 1] to Rk , i.e., πt1 ,··· ,tk (x) =
(x(t1 ), · · · , x(tk )).
However, the condition (1.36) is not a sufficient condition for Xn to
weakly converge to X. We shall require additional conditions. The Xn
is said to be weakly relatively compact if every subsequence has a further
convergent subsequence in the sense of weak convergence. According to the
subsequence convergence theorem, Xn is weakly convergent if all the limit
variables are identical in law. Another closely related concept is uniform
tightness. The Xn is uniformly tight if for any ε > 0, there is a compact
subset Kε in (S, ρ) such that
P (Xn ∈
/ Kε ) < ε, for all n ≥ 1.
The celebrated Prohorov’s theorem tells that the Xn must be weakly rel-
atively compact if Xn is uniformly tight. The converse is also true in a
separable complete metric space. A major point of this theorem is that the
weak convergence of probability measures rely on how they concentrate in
a compact subset in a metric space. In C[0, 1], the so-called Ascoli-Arzelà
lemma completely characterizes a relatively compact subset: K ⊆ C[0, 1]
is relatively compact if and only if
(i) uniform boundedness:
sup sup |x(t)| < ∞,
t∈R x∈K
Note that under condition (ii), (i) can be replaced by the condition
(i0 )
sup |x(0)| < ∞.
x∈K
March 5, 2015 15:59 9197-Random Matrices and Random Partitions ws-book9x6 page 31
Normal Convergence 31
To illustrate how to use the above general framework, we shall state and
prove the Donsker invariance principle.
Proof. We need to verify the conditions (1.36), (1.37) and (1.38). In-
deed, (1.36) directly follows from the Feller-Lévy CLT. (1.37) is trivial
since Xn (0) = 0, and (1.38) follows from the Lévy maximal inequality for
sums of independent random variables. The detail is left to the reader.
We remark that the random process constructed in (1.39) is a polygon
√
going through points (k/n, Sk / n). It is often referred to as a partial sum
process. The Donsker invariance principle has found a large number of
applications in a wide range of fields. To apply it, we usually need the
following mapping theorem. Let (S1 , ρ1 ) and (S2 , ρ2 ) be two metric spaces,
h : S1 7→ S2 a measurable mapping. Assume that X, Xn , n ≥ 1 is a
sequence of S1 -valued random elements and Xn ⇒ X. It is natural to ask
under what hypothesis the h(Xn ) still weakly converges to h(X). Obviously,
if h is continuous, then we have h(Xn ) ⇒ h(X). Moreover, the same still
holds if h is a measurable mapping and PX (Dh ) = 0 where Dh is the set
of all discontinuity points of h. As a simple example, we can compute the
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 32
√
limiting distribution of max1≤k≤n Sk / n. Indeed, let h(x) = sup0≤t≤1 x(t)
where x ∈ C[0, 1]. Then h is continuous and
1
h(Xn ) = √ max Sk , h(B) = sup B(t).
n 1≤k≤n 0≤t≤1
Hence it follows
1 d
√ max Sk −→ sup B(t), n → ∞.
n 1≤k≤n 0≤t≤1
Another example is to compute the R 1 limiting distribution of the weighted
Pn
sum k=1 kξk /n3/2 . Let h(x) = 0 x(t)dt where x ∈ C[0, 1]. Then h is
continuous and
n Z 1
1 X
h(Xn ) = 3/2 kξk + op (1), h(B) = B(t)dt,
n k=1 0
where op (1) is negligible. Hence it follows
n Z 1
1 X d
kξ k −→ B(t)dt.
n3/2 k=1 0
More interesting examples can be found in Billingsley (1999a). In addi-
tion to R∞ and C[0, 1], one can also consider weak convergence of random
processes in D[0, 1], C(0, ∞) and D(0, ∞).
As the reader might notice, in proving the weak convergence of Xn in
C[0, 1], the most difficult part is to verify the uniform tightness condition
(1.38). A weaker version than (1.38) is stochastic equicontinuity: for every
ε > 0 and η > 0, there is a δ > 0 such that
sup P (|Xn (s) − Xn (t)| > η) < ε, for all n ≥ 1. (1.40)
|s−t|<δ
Although (1.40) does not guarantee that the process Xn converges weakly,
we can formulate a limit theorem for comparatively narrow class of func-
tionals of integral form.
Theorem 1.16. Suppose that X = (X(t), 0 ≤ t ≤ 1) and Xn = (Xn (t), 0 ≤
t ≤ 1) is a sequence of random processes satisfying (1.36) and (1.40). Sup-
pose that g(t, x) is a continuous function and there is a nonnegative function
h(x) such that h(x) ↑ ∞ as |x| → ∞ and
|g(t, x)|
lim sup sup = 0.
a→∞ 0≤t≤1 |x|>a h(x)
Chapter 2
2.1 Introduction
33
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 34
Theorem 2.1. The joint probability density for the unordered eigenvalues
of a Haar distributed random matrix in Un is
1 Y
eiθj − eiθk 2 ,
pn eiθ1 , · · · , eiθn =
n
(2.1)
(2π) n!
1≤j<k≤n
Interestingly, these properties hold for general n. To see this, let for any
set of n-tuple complex numbers (x1 , · · · , xn )
∆(x1 , · · · , xn ) = det xj−1
k 1≤j,k≤n
.
Then the Vandermonde identity shows
Y
∆(x1 , · · · , xn ) = (xk − xj ).
1≤j<k≤n
Define
n
X sin nθ/2
Sn (θ) = e−i(n−1)θ/2 eikθ = , (2.2)
sin θ/2
k=1
where by convention Sn (0) = n. Thus we have by using the fact that the
determinant of a matrix and its transpose are the same
eiθj − eiθk 2 = ∆(eiθ1 , · · · , eiθn )2
Y
1≤j<k≤n
= det Sn (θk − θj ) 1≤j,k≤n .
This formula is very useful to computing some eigenvalue statistics. In order
to compute the m-dimensional marginal density, we also need a formula of
Gaudin (see Conrey (2005)), which states
Z 2π
Sn (θj − θ)Sn (θ − θk )dθ = 2πSn (θj − θk ).
0
As a consequence, we have
Z 2π
det Sn (θj − θk ) 1≤j,k≤n dθn = 2π det Sn (θj − θk ) 1≤j,k≤n−1 .
0
Repeating this yields easily
Z
1
pn,m eiθ1 , · · · , eiθm := pn eiθ1 , · · · , eiθn dθm+1 · · · dθn
n
n!(2π) [0,2π]n−m
Z
1
= n
det Sn (θj − θk ) n×n dθm+1 · · · dθn
n!(2π) [0,2π]n−m
(n − m)! 1
= m
det Sn (θj − θk ) 1≤j,k≤m .
n! (2π)
In particular, the first two marginal densities are
1
pn,1 eiθ =
, 0 ≤ θ ≤ 2π (2.3)
2π
and
1 1
pn,2 eiθ1 , eiθ2 = 1 − (Sn (θ1 − θ2 ))2 .
2
(2.4)
n(n − 1) (2π)
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 36
Γ( 21 βn + 1)
Zn,β = .
[Γ( 12 β + 1)]n
The family of probability measures defined by (2.5) is called Circular β En-
semble (CβE). The CUE corresponds to β = 2. Viewed from the opposite
perspective, one may say that the CUE provides a matrix model for the
Coulomb gas at the inverse temperature β = 2. In Section 2.5 we shall see
a matrix model for general β.
In this chapter we will be particularly interested in the asymptotic be-
haviours of various eigenvalue statistics as n tends to infinity. Start with
the average spectral measures. Let eiθ1 , · · · , eiθn be eigenvalues of a Haar
distributed unitary matrix Un . Put them together as a probability measure
on T:
n
1X
νn = δeiθk .
n
k=1
Theorem 2.2. As n → ∞,
νn ⇒ µ in P,
The proof is basically along the line of Diaconis and Shahshahani (1994).
We need a second-order moment estimate as follows.
This is a kind of law of large numbers, and is very similar to the Khinchine
law of large numbers for sums of i.i.d. random variables in standard prob-
ability theory, see (1.4). One cannot see from such a first-order average the
difference between eigenvalues and sample points chosen at random from
the unit circle T. However, a significant feature will appear in the second-
order fluctuation, which is the main content of the following sections.
σ = ϑτ ϑ−1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 39
Since character is a class function, then we can rewrite (2.10) and (2.11) as
1 X
hχ, ψi = |K|χ(K)ψ(K) = δχ,ψ ,
n!
K
where we mean by χ(K) and ψ(K) that χ and ψ act in K respectively, |K|
denotes the number of K and the sum is over all conjugacy classes of Sn .
This implies that the modified character table
r
|K|
U= χ(K)
n! χ,K
Theorem 2.4.
X n!
χ(K)χ(L) = δK,L , (2.12)
χ
|K|
λ1 + λ2 + · · · + λi ≥ µ1 + µ2 + · · · + µi
Theorem 2.5.
χλ (1n ) = dimS λ = dλ
and
X X
χ2λ (1n ) = d2λ = n!. (2.13)
λ7→n λ7→n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 44
We define
pλ = pλ1 pλ2 · · · pλl
for each partition λ = (λ1 , λ2 , · · · , λl ). Note that pλ , λ 7→ n do not form a
Z-basis of Λn . For instance,
1 1
h2 = p2 + p21
2 2
does not have integral coefficients when expressed
in terms of the pλ . In
general, for any partition λ = 1r1 , 2r2 , · · · of n, define
∞
Y
zλ = iri ri !. (2.15)
i=1
and
X
en = (−1)n−l(λ) zλ−1 pλ .
λ7→n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 46
Theorem 2.6.
X
pλ = χρ (λ)sρ . (2.16)
ρ7→|λ|
It turns out that Schur polynomials are orthonormal with respect to this
inner product, which is referred to as Schur orthonormality. In particular,
we have
Theorem 2.7.
When |µ| = |ν| ≤ n, the sum is taken over all partitions of |µ|, and so the
character relation (2.12) of the second kind shows
hpµ , pν i = zµ δµ,ν , (2.19)
where zµ is defined as in (2.15).
so that fˆ0 is the average of f over T. Since f is real, then fˆ−l = fˆl .
In this section we shall focus on the fluctuation of linear eigenvalue
Pn
statistic k=1 f eiθk around the average nfˆ0 .
P∞
Theorem 2.8. If f ∈ L2 (T, dµ) is such that l=1 l|fˆl |2 < ∞, then
n
d
X
f eiθk − nfˆ0 −→ N 0, σf2 , n → ∞
(2.20)
k=1
P∞
where σf2 =2 l=1 l|fˆl |2 .
This theorem goes back to Szegö as early as in 1950s, and is now known
as Szegö’s strong limit theorem. There exist at least six different proofs
with slight different assumptions on f in literature, and the most classical
one uses the orthogonal polynomials on the unit circle T. Here we prefer to
prove the theorem using the moment method of Diaconis and Evans (2001),
Diaconis (2003). The interested reader is referred to Simon (2004) for other
five proofs. See also a recent survey of Deift, Its and Krasovsky (2012) for
extensions and applications.
Lemma 2.4. Suppose that Z = X + iY is a complex standard normal
random variable, then for any non-negative integers a and b
EZ a Z̄ b = a!δa,b .
Proof. Z can clearly be written in polar coordinates as follows:
Z = γeiθ ,
where γ and θ are independent, θ is uniform over [0, 2π], and γ has density
2
function 2re−r , r ≥ 0. It easily follows
EZ a Z̄ b = Eγ a+b eiθ(a−b)
= Eγ a+b Eeiθ(a−b)
= Eγ 2a δa,b = a!δa,b ,
as desired.
Lemma 2.5. (i) Suppose that Zl , l ≥ 1 is a sequence of i.i.d. complex
standard normal random variables. Then for any m ≥ 1 and any non-
negative integers a1 , a2 , · · · , am and b1 , b2 , · · · , bm
Ym X n m X
a l Y n bl
E eilθk e−ilθk
l=1 k=1 l=1 k=1
m √ m √
Y al Y b
=E lZl lZ l l (2.21)
l=1 l=1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 49
Pm Pm
whenever n ≥ max l=1 lal , l=1 lbl .
(ii) For any integer j, l ≥ 1,
Xn X n
E eilθk e−ijθk = δj,l min(l, n). (2.22)
k=1 k=1
l=1 k=1
and
m X
Y n b l
e−ilθk = pµ (eiθ1 , · · · , eiθn )
l=1 k=1
where λ = 1a1 , 2a2 , · · · , mam and µ = 1b1 , 2b2 , · · · , mbm .
According to the orthogonality relation in (2.19), we have
Ym Xn m X
a l Y n bl
E eilθk e−ilθk
l=1 k=1 l=1 k=1
= hpλ , pµ i
Ym
= δλ,µ lal al ! (2.23)
l=1
whenever |λ|, |µ| ≤ n. Now we can get the identity (2.21) using Lemma 2.4.
Turn to (2.22). We immediately know from (2.23) that the expectation
is zero if j 6= l, while the case j = l has been proven in (2.6).
As an immediate consequence, we have the following
P∞
Proof of Theorem 2.8. Since l=−∞ |fˆl |2 < ∞, then we can express f
in terms of Fourier series
X∞
f eiθ = fˆl e−ilθ ,
l=−∞
from which it follows
Xn ∞
X n
X
f eiθk = fˆl e−ilθk
whenever n ≥ mn (a + b).
If mn /n → 0, then for any non-negative integer numbers a and b, the
assumption n ≥ mn (a+b) can be guaranteed for sufficiently large n. Now we
can conclude the claim (2.25) by letting n → ∞ in (2.26). The proof is now
complete.
(ii)
mn n
1 X X d
√ fˆl e−ilθk −→ NC (0, 1). (2.29)
Bn l=1 k=1
According to (2.22),
X ∞ n
X 2 ∞
X
E fˆl e−ilθk = |fˆl |2 min(l, n)
l=mn +1 k=1 l=mn +1
Xn ∞
X
= l|fˆl |2 + n |fˆl |2 .
l=mn +1 l=n+1
Summing by parts,
∞ ∞
X X 1
2 |fˆl |2 = (Bl+1 − Bl )
l+1
l=n+1 l=n
∞
X Bl Bn
= − .
l(l + 1) n + 1
l=n
Thus (2.29) is valid from (2.27) and (2.30). The proof is now complete.
To conclude this section, we shall look at two interesting examples. The
first one is the distribution of values taken by the logarithm of charac-
teristic polynomial of a random unitary matrix. Recall the characteristic
polynomial of a matrix Un is defined by the determinant
det(zIn − Un ).
Fix z = eiθ0 and assume Un is from the CUE. Since eiθ0 is almost surely
not an eigenvalue of Un , then
n
Y
det eiθ0 In − Un = eiθ0 − eiθk 6= 0.
k=1
It is fascinating that the logarithm of det eiθ0 In −Un after properly scaled
weakly converges to a normal distribution, analogous to Selberg’s result on
the normal distribution of values of the logarithm of the Riemann zeta
function. This was first observed by Keating and Snaith (2000), which
argued that the Riemann zeta function on the critical line could be modelled
by the characteristic polynomial of a random unitary matrix.
Theorem 2.11. As n → ∞,
1 d
√ log det(eiθ0 In − Un ) − inθ0 −→ NC (0, 1),
log n
where log denotes the usual branch of the logarithm defined on C \ {z :
Re(z) ≤ 0}.
k=1
and so
n ∞ n
X eiθk X 1 X ilθk
log 1 − =− e .
r lrl
k=1 l=1 k=1
The proof is very similar to that of Theorem 2.8. Let mn = n/log n so that
mn → ∞ and mn /n → 0. We shall establish the following statements:
(i)
∞ n
1 X 1 X ilθk P
√ e −→ 0; (2.32)
log n l=m +1 l k=1
n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 55
(ii)
mn n
1 X 1 X ilθk d
√ e −→ NC (0, 1). (2.33)
log n l=1 l k=1
According to (2.22), it holds
∞ n n ∞
X 1 X ilθk 2 X 1 X 1
E e = +n
l l l2
l=mn +1 k=1 l=mn +1 l=n+1
= O(log log n),
which together with the Markov inequality directly implies (2.32).
To prove (2.33), note for any non-negative integers a and b,
mn n mn n
X 1 X ilθk a X 1 X −ilθk b
E e e
l l
l=1 k=1 l=1 k=1
mn
X 1 a
= a!δa,b
l
l=1
= (1 + o(1))(log n)a a!δa,b ,
as desired.
The second example of interest is the numbers of eigenvalues lying in an
arc. For 0 ≤ a < b < 2π, write Nn (a, b) for the number of eigenvalues eiθk
with θk ∈ [a, b]. Particularly speaking,
n
X
Nn (a, b) = 1(a, b) (θk ).
k=1
This section is aimed to establish a five diagonal sparse matrix model for the
CUE and to provide an alternate approach to asymptotic normality of the
characteristic polynomials and the number of eigenvalues lying in an arc.
We first introduce basic notions of orthogonal polynomials and Verblunsky
coefficients associated to a finitely supported measure on the unit circle,
and quickly review some well-known facts, including the Szegö recurrence
equations and Verblunsky’s theorem. The measure we will be concerned
with is the spectral measure induced by a unitary matrix and a cyclic
vector. Two matrices of interest to us are upper triangular Hessenberg
matrix and CMV five diagonal matrix, whose Verblunsky coefficients can
be expressed in a simple way. Then we turn to a random unitary matrix
distributed with Haar measure. Particularly interesting, the associated
Verblunsky coefficients are independent Θv -distributed complex random
variables. Thus as a consequence of Verblunsky’s theorem, we naturally get
a five diagonal matrix model for the CUE. Lastly, we rederive Theorems 2.11
and 2.12 via a purely probabilistic approach: use only the classical central
limit theorems for sums of independent random variables and martingale
difference sequences.
Assume we are given a finitely supported probability measure dν on
exactly n points eiθ1 , eiθ2 , · · · , eiθn with masses ν1 , ν2 , · · · , νn , where νi > 0
Pn
and i=1 νi = 1. Let L2 (T, dν) be the of square integrable functions on T
with respective to dν with the inner product given by
Z
f eiθ g eiθ dν.
hf, gi =
T
Applying the Gram-Schmidt algorithm to the ordered set 1, z, · · · , z n−1 ,
we can get a sequence of orthogonal polynomials Φ0 , Φ1 , · · · , Φn−1 , where
Φ0 (z) = 1, Φk (z) = z k + lower order.
Define the Szegö dual by
Φ∗k (z) = z k Φk (z̄ −1 ).
Namely,
k
X k
X
Φk (z) = cl z l ⇒ Φ∗k (z) = c̄k−l z l .
l=0 l=0
As Szegö discovered, there exist complex constants α0 , α1 , · · · , αn−2 ∈ D,
where D := {z ∈ C, |z| < 1}, such that for 0 ≤ k ≤ n − 2
Φk+1 (z) = zΦk (z) − ᾱk Φ∗k (z) (2.35)
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 58
and
Φ∗k+1 (z) = Φ∗k (z) − αk zΦk (z). (2.36)
Expanding z n in this basis shows that there exists an αn−1 , say αn−1 =
eiη ∈ T, 0 ≤ η < 2π, such that if letting
Φn (z) = zΦn−1 (z) − ᾱn−1 Φ∗n−1 (z), (2.37)
then
Φn (z) = 0 in L2 (T, dν).
Define
p
ρk = 1 − |αk |2 , 0 ≤ k ≤ n − 1,
then it follows from recurrence relations (2.35) and (2.36)
k−1
Y
kΦ0 k = 1, kΦk k = ρl , k ≥ 1. (2.38)
l=0
This theorem is now called Verblunsky’s theorem (also called Favard’s the-
orem for the circle). The reader is referred to Simon (2004) for the proof
(at least four proofs are presented).
It is very expedient to encode the Szegö recurrence relation (2.35) and
(2.36). Let
Φk (z)
Bk (z) = z .
Φ∗k (z)
It easily follows
1 − ᾱk B̄k (z)
B0 (z) = z, Bk+1 (z) = zBk (z) , z ∈ T, (2.39)
1 − αk Bk (z)
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 59
which shows that the Bk (z) can be completely expressed in terms of the
Verblunsky coefficients αk .
Note in view of (2.39), Bk is a finite Blaschke product of degree k + 1.
Define a continuous function ψk (θ) : [0, 2π) 7→ R via
Bk eiθ = eiψk (θ) .
(2.40)
ψk (θ) is the absolute Präfer phase of Bk , so the set of points
iθ
e : Bn−1 (eiθ ) = ᾱn−1 = eiθ : ψn−1 (θ) ∈ 2πZ + η
L
and the entries Hij depend on the Verblunsky coefficients αk and ρk in a
complicated way. This makes difficult this task. Moreover, the numerical
computations of zeros of high degree orthogonal polynomials becomes a
nontrivial problem due to the Hessenberg structure of HnL .
To overcome this difficulty, Cantero, Moral, and Velázquez (2003)
used a simple and ingenious idea. Applying the Gram-Schmidt proce-
−1 2 −2
dure
to the first
n of the ordered set 1, z, z , z , z , · · · rather than
1, z, · · · , z n−1 , we can get a sequence of orthogonal Laurent polynomi-
als, denoted by χk (z), 0 ≤ k ≤ n−1. We will refer to the χk as the standard
right orthonormal L-polynomial with respect to the measure dν. Interest-
ingly, the χk can be expressed in terms of the orthonormal polynomial φk
and its Szegö dual φ∗k as follows:
onal Laurent polynomial, denoted by χk∗ . We call the χk∗ the standard
left orthogonal L-polynomial. It turns out that the χk and χk∗ are closely
related to each other through the equation:
Define
ᾱk ρk
Ξk =
ρk −αk
ρ2k−2 ᾱ2k−1 −α2k−2 ᾱ2k−1
χ2k−1 (z) χ2k−2 (z)
z =
χ2k (z)ρ2k−2 ρ2k−1 −α2k−2 ρ2k−1 χ2k−1 (z)
ρ2k−1 ᾱ2k ρ2k−1 ρ2k χ2k (z)
+ .
−α2k−1 ᾱ2k −α2k−1 ρ2k χ2k+1 (z)
Construct now the n × n block diagonal matrices
L = diag(Ξ0 , Ξ2 , Ξ4 , · · · ), M = diag(Ξ−1 , Ξ1 , Ξ3 , · · · )
and define
Cn = ML, Cnτ = LM. (2.43)
It is easy to check that L and M are symmetric unitary matrices, and so
both Cn and Cnτ are unitary.
A direct manipulation of matrix product shows that Cn is a five diagonal
sparse matrix. Specifically speaking, if n = 2k, then Cn is equal to
ᾱ0 ρ0 0 0 0 ··· 0 0
ρ0 ᾱ1 −α0 ᾱ1 ρ1 ᾱ2 ρ1 ρ2 0 ··· 0 0
ρ ρ −α ρ −α ᾱ −α ρ 0 ··· 0 0
0 1 0 1 1 2 1 2
0 0 ρ2 ᾱ3 −α2 ᾱ3 ρ3 ᾱ4 · · · 0 0
,
0 0 ρ2 ρ3 −α2 ρ3 −α3 α4 · · · 0 0
.. .. .. .. .. .. .. ..
.
. . . . . . .
0 0 0 0 0 · · · −αn−3 ᾱn−2 −αn−3 ρn−2
0 0 0 0 0 · · · ρn−2 ᾱn−1 −αn−2 ᾱn−1
while if n = 2k + 1, then Cn is equal to
ᾱ0 ρ0 0 0 ··· 0 0 0
ρ0 ᾱ1 −α0 ᾱ1 ρ1 ᾱ2 ρ1 ρ2 · · · 0 0 0
ρ ρ −α ρ −α ᾱ −α ρ · · · 0 0 0
0 1 0 1 1 2 1 2
0 0 ρ2 ᾱ3 −α2 ᾱ3 · · · 0 0 0
.
0 0 ρ2 ρ3 −α2 ρ3 · · · 0 0 0
.. .. .. .. .. .. .. ..
.
. . . . . . .
0 0 0 0 · · · ρn−3 ᾱn−2 −αn−3 ᾱn−2 ρn−2 ᾱn−1
0 0 0 0 · · · ρn−3 ρn−2 −αn−3 ρn−2 −αn−2 ᾱn−1
The multiplication operator Π : f (z) 7→ zf (z) can be explicitly expressed
in the basis of χk , 0 ≤ k ≤ n − 1 as follows. If n = 2k, then
χ0 χ0 0
χ1 χ1 0
..
Π ... = Cn ... +
. , (2.44)
χn−2 χn−2 0
Φn
χn−1 χn−1 z k−1 kΦn−1 k
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 62
while if n = 2k + 1, then
χ0 χ0 0
χ1
χ1
0
..
.. ..
Π .
= Cn
+ .
. . (2.45)
χn−3
χn−3
0
χn−2 ρn−2 zk−1 Φn
χn−2 kΦn−1 k
χn−1 χn−1 −αn−2 zk−1 Φ n
kΦn−1 k
det(zIn − Cn ) = Φn (z).
which implies
det(zIn − Cn ) Φn (z)
= .
det(zIn−1 − Cn,n−1 ) Φn−1 (z)
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 63
where Bk = (x1 , · · · , xk ) : x21 + · · · + x2k < 1 .
In particular, X1 is Beta(1/2, n/2)-distributed and X1 + iX2 is Θn -
distributed.
Proof. (2.47) is actually a direct consequence of the following change-of-
variables formula using matrix volume:
Z Z
f (v)dv = (f ◦ φ)(u)Jφ (u)du (2.48)
V U
where V ⊆ Rn and U ⊆ Rm with n ≥ m, f is integrable on V, φ : U 7→ V
is sufficiently well-behaved function,
dv
and du denote respectively the
volume element in V ⊆ Rn , and Jφ (u) is the volume of Jacobian matrix
Jφ (u).
To apply (2.48) in our setting, let
φk (x1 , · · · , xn ) = xk , 1≤k≤n
and
1/2
φn+1 (x1 , · · · , xn ) = 1 − x21 − · · · − x2n .
n n
So the S is the graph of B under the mapping φ = (φ1 , · · · , φn+1 ). The
Jacobian matrix of φ is
1 0 0 ··· 0 0
0
1 0 ··· 0 0
Jφ =
..
.
0 0 0 . 1 0
0 0 0 ··· 0 1
∂φn+1 ∂φn+1 ∂φn+1 ∂φn+1 ∂φn+1
∂x1 ∂x2 ∂x3 ··· ∂xn−1 ∂xn
This is an n + 1 × n rectangular matrix, whose volume is computed by
q
Jφ = det(J τ Jφ )
φ
−1/2
= 1 − x21 − · · · − x2n .
Hence according to (2.48), we have
Ef (X1 , · · · , Xk )
Z
= f (x1 , · · · , xk )ds
Sn
Γ( n+1
2 )
Z
−1/2
= (n+1)/2
f (x1 , · · · , xk ) 1 − x21 − · · · − x2n dx1 · · · dxn
2π Bn
Γ( n+1
2 )
Z
= f (x1 , · · · , xk )dx1 · · · dxk
2π (n+1)/2 Bk
Z
−1/2
· 1 − x21 − · · · − x2n dxk+1 · · · dxn (2.49)
x2k+1 +···+x2n <1−x21 −···−x2k
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 66
Remark 2.2. Lemma 2.9 can also be proved by using the following well-
known fact: let g1 , · · · , gn+1 be a sequence of i.i.d. standard normal random
variables, then
d 1
(X1 , · · · , Xn+1 ) = 2 2
(g1 , · · · , gn+1 ).
(g1 + · · · + gn+1 )1/2
To keep notation consistent, Z is said to be Θ1 -distributed if Z is uniform
on the unit circle.
The key to proving Theorem 2.15 is the Householder transform, which will
transfer unitary matrix into an upper triangular Hessenberg form. Write
Un = (uij )n×n . Let w = (w1 , w2 , · · · , wn )τ where
u21 1 |u21 | 1/2
w1 = 0, w2 = − − ,
|u21 | 2 2α
ul1
wl = − , l≥3
(2α2 − 2α|u21 |)1/2
where α > 0 and
α2 : = |u21 |2 + |u31 |2 + · · · + |un1 |2
= 1 − |u11 |2 .
Trivially, it follows
w∗ w = 1.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 67
Define
···
1 0 0
0
Rn = In − 2ww∗ = . .
.. Vn−1
0
This is a reflection through the plane perpendicular to w. It is easy to
check
Rn−1 = Rn∗ = Rn
and
u11 (u12 , u13 , · · · , u1n )Vn−1
u21
u31
Rn Un Rn = ,
Vn−1 . Vn−1 Un,n−1 Vn−1
..
un1
where Un,n−1 is the n − 1 × n − 1 submatrix of Un by deleting the first row
and the first column.
Take a closer look at the first column. The first element of Rn Un Rn is
unchanged, u11 ; while the second is
u21
u32
u21
1 − 2w2 w2∗ , −2w2 w3∗ , · · · , −2w2 wn∗ . = α ,
.. |u21 |
un1
and the third and below are zeros. So far we have described one step of
the usual Householder algorithm. To make the second entry nonnegative,
we need to add one further conjugation. Let Dn differ from the identity
matrix by having (2, 2)-entry e−iφ with φ chosen appropriately and form
Dn Rn Un Rn Dn∗ . Then we get the desired matrix
u 11 (u12 , u13 , · · · , u1n )V n−1
p1 − |u |2
11
0
.
.. Vn−1 Un,n−1 Vn−1
.
0
Proof of Theorem 2.15. We shall apply the above refined Householder
algorithm to a random unitary matrix Un . To do this, we need the following
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 68
realization of Haar measure: choose the first column at random from the
unit sphere Sn−1 ; then choose the second column from the unit sphere of
vectors orthogonal to the first; then the third column and so forth. In this
way we get a Haar matrix because it is invariant under left multiplication
by any unitary matrix.
Now the first column of Un is a random vector from the unit sphere
Sn−1 . After applying the above refined Householder algorithm, the new
first column take the form (ᾱ0 , ρ0 , 0, · · · , 0)τ where ᾱ0 = u11 the original
(1, 1) entry of Un and so is by Lemma 2.9 Θ2n−1 -distributed, while ρ0 =
p
1 − |α0 |2 as desired. The other columns are still orthogonal to the first
column and form a random orthogonal basis for the orthogonal complement
of the first column. Remember Haar measure is invariant under both right
and left multiplication by a unitary.
For the subsequent columns the procedure is similar. Assume the (k −
1)th column is
ᾱk−2 ρ0 ρ1 · · · ρk−3
−ᾱk−2 α0 ρ1 · · · ρk−3
..
.
−ᾱk−2 αk−3
.
ρk−2
0
..
.
0
Let
ρ0 ρ1 · · · ρk−2
−α0 ρ1 · · · ρk−2
..
.
−αk−3 ρk−2
X= ,
−αk−2
0
..
.
0
then X is a unit vector orthogonal to the first k − 1 columns. Namely X is
an element of the linear vector space spanned by the last n−k+1 columns in
Un . Its inner product with the kth column, denoted by ᾱk−1 , is distributed
as the entry of a random vector from 2(n−k + 1)-sphere and is independent
of α0 , α1 , · · · , αk−2 . This implies that αk−1 is Θ2(n−k)+1 -distributed.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 69
We now multiply the matrix at hand from left by the appropriate re-
flection and rotation to bring the kth columns into the desired form. Note
neither these operations alter the top k rows and so the inner product of the
kth column with X is unchanged. But now the kth column is uniquely de-
termined; it must be ᾱk−1 X +ρk−1 ek+1 , where ek+1 = (0, · · · , 0, 1, 0 · · · 0)τ .
We then multiply on the right by RD∗ , but this leaves the first k column
unchanged, while orthogonally intermixing the other columns. In this way,
we obtain a matrix whose first k column confirm to the structure HnU . While
the remaining columns form a random basis for the orthogonal complements
of the span of those k columns.
In this way, we can proceed inductively until we reach the last column.
It is obliged to be a random orthonormal basis for the one-dimensional space
orthogonal to the preceding n − 1 columns and hence a random unimodular
multiple, say ᾱn−1 , of X. This is why the last Verblunsky coefficient is
Θ1 -distributed.
We have now conjugated Un to a matrix in the form of Hessenberg as
in Lemma 2.7. Note the vector e1 is unchanged under the action of each of
the conjugating matrices, then
αk (Un , e1 ) = αk HnU , e1 = αk .
then
1 0
Un := Vn (2.51)
0 Un−1
is distributed with Haar measure dµn on Un .
Lemma 2.11. Let Un be a random matrix from the CUE with the Verblun-
sky coefficients α0 , α1 , · · · , αn−1 . Then
n−1
d
Y
det(In − Un ) = (1 − αk ). (2.53)
k=0
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 71
−wn1
∗
= −w11 det(Un−1 − In−1 ). (2.56)
Substituting (2.56) in (2.55), we get
1 0
det In − Vn = −w11 det(In−1 − Un−1 ).
0 Un−1
Observe w11 = v11 − 1 and v11 ∼ Θ2n−1 -distributed. Thus it follows
d
det((In − Un ) = (1 − α0 ) det(In−1 − Un−1 ).
Proceeding in this manner, we have
n−1
d
Y
det(In − Un ) = (1 − αk ),
k=0
as required.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 72
Remark 2.3. The identity (2.53) can also be deduced using the recurrence
relation of orthogonal polynomials Φk (z). Indeed, according to Theorem
2.16,
d
det(In − Un ) = det(In − Cn ). (2.57)
On the other hand, by Lemma 2.6 and (2.37)
det(In − Cn ) = Φn (1)
= Φn−1 (1) − ᾱn−1 Φ∗n−1 (1).
Note by (2.40)
Φ∗n−1 (1)
= e−iψn−1 (0) .
Φn−1 (1)
Hence we have
det(In − Cn ) = Φn−1 (1) 1 − ᾱn−1 e−iψn−1 (0) .
which together with (2.57) implies (2.53). It is worth mentioning that the
identity (2.58) is still valid for the CβE discussed in next section.
Proof. (i) directly follows from (2.47), while a simple computation easily
yields (ii).
Proof of Theorem 2.11. Without loss of generality, we may and do
assume θ0 = 0. According to Lemma 2.11, it suffices to prove the following
asymptotic normality
n−1
1 X d
√ log(1 − αk ) −→ NC (0, 1). (2.59)
log n k=0
Since |αk | < 1 almost surely for k = 0, 1, · · · , n − 2, then
∞
X 1
log(1 − αk ) = − αkl .
l
l=1
and
n−2
X π2
E|αk |4 ≤ .
3
k=0
So, it follows
1 P
√ Zn,3 −→ 0. (2.64)
log n
Finally, for Zn,4 , note αn−1 = eiη is uniformly distributed on T, then almost
surely αn−1 is not equal to 1, and so log(1 − αn−1 ) < ∞. Hence it follows
1 P
√ Zn,4 −→ 0. (2.65)
log n
Gathering (2.60), (2.63), (2.64) and (2.65) together implies (2.59).
Turn to Theorem 2.12. Let αk = αk (Un , e1 ), the Verblunsky coefficients
associated to (Un , e1 ), and construct the Bk and ψk as in (2.39) and (2.40),
then
e : Bn−1 (eiθ ) = e−iη = eiθ : ψn−1 (θ) ∈ 2πZ + η
iθ
In this way, it suffices to show that asymptotically, ψn−1 (b) and ψn−1 (a)
follow a joint normal law. This is what Killip and Nenciu (2008) employed
in the study of general CβE.
Then we have
EΥ(a, α) = 0.
E Υ̃(a, α)Υ̃(b, α)
= 4E|α|2 sin(a + argα) sin(b + argα)
4(v − 1) 1 2π 3
Z Z
(v−3)/2
= r 1 − r2 sin(θ + a) sin(θ + b)drdθ
2π 0 0
4
= cos(b − a).
v+1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 76
Similarly, we have
16(v − 1) 1 2π 5
Z Z
(v−3)/2 4
E Υ̃(a, α)4 = r 1 − r2 sin (θ + a)drdθ
2π 0 0
48
= .
(v + 1)(v + 3)
Applying Plancherel’s theorem to the power series formula for Υ gives
∞
2 X 2
E|α|2l .
E Υ(a, α) − Υ̃(a, α) =
l2
l=2
π2
≤2 − 1 E|α|4
6
(2.67) easily follows from Lemma 2.12.
Lastly, combining (2.67) and (2.69) implies (2.70).
Θk+1 = Θk + δ + γk
Proof. Note
n
X
eiδ − 1 ak eiΘk
k=1
n
X
ak ei(Θk +δ) − eiΘk
=
k=1
Xn n
X
iΘk+1 iΘk
ak eiΘk+1 − ei(Θk +δ) .
= ak e −e −
k=1 k=1
Then (2.71) easily follows using summation by parts and the fact 1−eiγk ≤
|γk |.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 77
k=0
and
n−2
1 X 4
2 E Υ̃ ψk (a), αk ) → 0.
log n k=0
The goal of this section is to extend the five diagonal matrix representation
to the CβE.
Recall that the CβE represents a family of probability measures on
n points of T with density function pn,β (eiθ1 , · · · , eiθn ) defined by (2.5).
As observed in Section 2.1, pn,2 describes the joint probability density of
eigenvalues of a unitary matrix chosen from Un according to Haar measure.
Similarly, pn,1 (pn,4 ) describes the joint probability density of eigenvalues of
an orthogonal (symplectic) matrix chosen from On (Sn ) according to Haar
measure. However, no analog holds for general β > 0.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 79
The following five diagonal matrix model discovered by Killip and Nen-
ciu (2004) plays an important role in the study of CβE.
The rest of this section is to prove the theorem. The proof is actually an
ordinary use of the change of variables in standard probability theory. Let
HnU be as in (2.46), then we have by Lemmas 2.7 and 2.8
αk (Cn , e1 ) = αk HnU , e1 = αk .
So it suffices to prove the claim for HnU . Denote the ordered eigenvalues of
HnU by eiθ1 , · · · , eiθn . Then there must be a unitary matrix Vn = (vij )n×n
such that
iθ1
0 ··· 0
e
0 eiθ2 · · · 0
∗
HnU = Vn . .. Vn . (2.73)
.. . .
.. . . .
0 0 · · · eiθn
Vn may be chosen to consist of eigenvectors v1 , v2 , · · · , vn . We also further
require that v11 := q1 > 0, v12 := q2 > 0, · · · , v1n := qn > 0, thus Vn is
uniquely determined. It easily follows
q12 + q22 + · · · + qn2 = 1 (2.74)
because of orthonormality of Vn .
The following lemma gives an elegant identity between the eigenvalues
and eigenvectors and the Verblunsky coefficients.
Lemma 2.15.
n n−2
e j − eiθk 2 =
Y Y Y n−l−1
ql2
iθ
1 − |αl |2
.
l=1 1≤j<k≤n l=0
and
q12 ···
0 0
0 q22 ··· 0
Q= .
.. .. ..
..
. . .
0 0 · · · qn2
then it follows
n
e j − eiθk 2 = det(AQA∗ ).
Y Y
ql2
iθ
l=1 1≤j<k≤n
On the other hand, define
Φ0 (eiθ1 ) Φ0 (eiθ2 ) · · · Φ0 (eiθn )
Φ1 (eiθ1 ) Φ1 (eiθ2 ) · · · Φ1 (eiθn )
B= ,
.. .. .. ..
. . . .
iθ1 iθ2 iθn
Φn−1 (e ) Φn−1 (e ) · · · Φn−1 (e )
where Φ0 , Φ1 , · · · , Φn−1 are monic orthogonal polynomials associated to the
Verblunsky coefficients α0 , α1 , · · · , αn−1 . Then it is trivial to see
det(A) = det(B).
In addition, from the orthogonality property of the Φl , it follows
kΦ0 k2 0 · · ·
0
0 kΦ1 k2 · · · 0
BQB ∗ = . ,
. ..
.. .. . 0
0 0 · · · kΦn−1 k2
Pn
where kΦl k2 = j=1 qj2 |Φl (eiθj )|2 .
Hence according to (2.38),
n−1
Y
det(AQA∗ ) = det(BQB ∗ ) = kΦl k2
l=0
n−2
2 n−l−1
Y
= 1 − |αl |
l=0
just as required.
A key ingredient to the proof of Theorem 2.17 is to look for a proper change
of variables and to compute explicitly the corresponding determinant of the
Jacobian. For any |t| < 1, it follows from (2.73)
(1 − teiθ1 )−1 · · ·
0
In − tHnU
−1
= Vn .. .. .. ∗
Vn . (2.75)
. . .
iθn −1
0 · · · (1 − te )
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 81
n
X
ᾱ0 = qj2 eiθj
j=1
n
X
∗ + ρ20 ᾱ1 = qj2 ei2θj
j=1
Xn
∗ + ρ20 ρ21 ᾱ2 = qj2 ei3θj (2.76)
j=1
.. ..
. .
n
X
∗ + ρ20 ρ21 · · · ρ2n−2 ᾱn−1 = qj2 einθj
j=1
where the ∗ denotes terms involving only variables already having appeared
on the left hand side of the preceding equations.
In this way, we can naturally get a one-to-one mapping from
(α0 , α1 , · · · , αn−1 ) to (eiθ1 , · · · , eiθn , q1 , · · · , qn−1 ). Recall that αk , 0 ≤ k ≤
n − 2 has an independent real and imaginary part, while αn−1 , eiθj have
unit modulus. We see that the number of variables is equal to 2n − 1. In
particular, let αk = ak + ibk and define J to be the determinant of the
Jacobian matrix for the change of variables, namely
Vn−2 V
k=0 dak ∧ dbk dᾱn−1
J= Vn−1 Vn
l=1 dql j=1 dθj
n−1
X
qn dqn = − qj dqj . (2.77)
j=1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 82
4
Q
1≤j<k≤n (xj − xk )
D(x1 , · · · , xn ) = cn Qn 2n−3
j=1 xj
In the cofactor, we add n−3 times the first column to the (n−1)th column.
−(2n−7)
Then we see the coefficient of x2 is given by a cofactor of the following
2 × 2 matrix !
−(n−4) −(n−4) −(n−4) −(n−4)
x2 − xn x2 − (n − 1)xn
−(n−3) −(n−3) −(n−3) .
x2 − xn −(n − 3)xn
Proceeding in this manner, we see the coefficient of x−5 k−1 is given by the
determinant
of the n + 1 × n + 1 matrix
xk − xn · · · xn−1 − xn xk ··· xn
x−1 − x−1 · · · x−1 − x−1 −x−1 ··· −x−1
k n n−1 n k n
2 2 2 2 2 2
xk − xn · · · xn−1 − xn 2xk ··· 2xn
.. .. .. .. .. .. .
. . . . . .
n−1 n−1 n−1 n−1 n−1 n−1
xk − xn · · · xn−1 − xn (n − 1)xk · · · (n − 1)xn
cn = (−1)(n−1)(n−2)/2 ,
as desired.
Lemma 2.17.
Qn−2
l=0 1 − |αl |2
|J| = Qn .
qn l=1 ql
Proof. Note
It easily follows
1 −i
dᾱk ∧ dαk = det dak ∧ dbk
1 i
= 2idak ∧ dbk . (2.83)
Then
n
Γ( β2 )n
Z
1 Y β−1
qj dq1 · · · dqn−1 = ,
∆n qn j=1 2n−1 Γ( βn2 )
q
2
where qn = 1 − q12 − · · · − qn−1 .
Proof. We only consider the case of n ≥ 2 since the other case is trivial.
Start with n = 2. Using the change of variable, we have
Z 2 Z 1
1 Y β−1 β/2−1 β−1
qj dq1 = 1 − q12 q1 dq1
∆2 q2 j=1 0
1 1
Z
β/2−1
= (1 − q1 )β/2−1 q1 dq1
2 0
Γ( β2 )2
= .
2Γ(β)
Assume by induction that the claim is valid for some n ≥ 2. It easily follows
Z n+1
1 Y β−1
qi dq1 · · · dqn
∆n+1 qn+1 i=1
Z n
β/2−1 Y
= 1 − qn2 − q12 − · · · − qn−1
2
qiβ−1 dq1 · · · dqn
q12 +···+qn−1
2 2
≤1−qn i=1
Z 1 Z n−1
β/2−1 Y
= 1 − qn2 qnβ−1 dqn qiβ−1
0 q12 +···+qn−1
2 2
≤1−qn i=1
2
q12 qn−1 β/2−1
· 1− − · · · − dq1 · · · dqn−1 . (2.84)
1 − qn2 1 − qn2
Making a change of variable, the inner integral becomes
β/2−1 n−1
Z Y β−1
(1 − qn2 )(n−1)β/2 1 − q12 − · · · − qn−1
2
qi dq1 · · · dqn−1 ,
∆n i=1
Γ( β2 )n+1
= .
2n Γ( (n+1)β
2 )
We conclude by induction the proof.
Proof of Theorem 2.17. As remarked above, (2.73) naturally induces a
one-to-one mapping from eiθ1 , · · · , eiθn , q1 , · · · , qn−1 to (a0 , b0 , · · · , an−2 ,
bn−2 , αn−1 ). Let fn,β and hn,β be their respective joint probability density
functions. Then it follows by Lemmas 2.17 and 2.15
fn,β eiθ1 , · · · , eiθn , q1 , · · · , qn−1
Theorem 2.18. Let eiθ1 , · · · , eiθn be chosen on the unit circle according to
the CβE. Then as n → ∞
(i) for any θ0 with 0 ≤ θ0 < 2π,
n
1 X d
q log 1 − ei(θj −θ0 ) −→ NC (0, 1);
2
β log n j=1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 88
where Nn (a, b) denotes the number of the angles θj lying in the arc between
a and b.
Chapter 3
3.1 Introduction
89
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 90
where x = (x1 , · · · , xn ) ∈ Rn .
This theorem, due to Weyl, plays an important role in the study of GUE.
Its proof can be found in the textbooks Anderson, Guionnet and Zeitouni
(2010), Deift and Gioev (2009). (3.2) should be interpreted as follows. Let
f : Hn 7→ R be an invariant function, i.e., f (H) = f (U HU ∗ ) for each
H ∈ Hn and unitary matrix U . Then
Z
Ef (H) = f (x)pn (x)dx.
Rn
It is worthy to remark that there are two factors in the righthand side of
(3.2). One is the product of n standard normal density functions, while the
other is the square of Vandermonde determinant. The probability that two
eigenvalues neighbor each other very closely is very small. Hence intuitively
speaking, eigenvalues should locate more neatly than i.i.d. normal random
points in the real line. It is the objective of this chapter that we shall take
a closer look at the arrangement of eigenvalues from global behaviours.
In order to analyze the precise asymptotics of pn (x), we need to intro-
duce Hermite orthogonal polynomials and the associated wave functions.
Let hl (x), l ≥ 0 be a sequence of monic orthogonal polynomials with respect
2
to the weight function e−x /2 with h0 (x) = 1. Then
2 dl 2
hl (x) = (−1)l ex /2 l e−x /2
dx
[l/2]
X xl−2i
= l! (−1)i i , l ≥ 1. (3.3)
i=0
2 i!(l − 2i)!
Define
2
ϕl (x) = (2π)−1/4 (l!)−1/2 hl (x)e−x /4
(3.4)
so that we have Z
∞
ϕl (x)ϕm (x)dx = δl,m , ∀l, m ≥ 0.
−∞
Now a simple matrix manipulation directly yields
1 ··· 1
1
Y x1 x2 · · · xn
(xj − xi ) = det .
.. . . .
.. . ..
1≤i<j≤n .
xn−1
1 xn−1
2 · · · xn−1
n
h0 (x2 ) · · · h0 (xn )
h0 (x1 )
h1 (x2 ) · · · h1 (xn )
h1 (x1 )
= det . (3.5)
.. .. .. ..
. . . .
hn−1 (x1 ) hn−1 (x2 ) · · · hn−1 (xn )
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 91
Proposition 3.1.
1
pn (x) = det Kn (xi , xj ) n×n , (3.6)
n!
where Kn is defined by
n−1
X
Kn (x, y) = ϕl (x)ϕl (y). (3.7)
l=0
Such a expression like (3.6) turns out to be very useful in the study of
asymptotics of eigenvalues. In fact, the GUE is one of the first examples
of so-called determinantal point processes (see Section 3.3 below for more
details). A nice observation about the kernel Kn is the following:
Z ∞
Kn (x, z)Kn (z, y)dz = Kn (x, y).
−∞
(iv) boundedness
κ := sup kϕl k∞ < ∞. (3.11)
l≥0
(ii) Given s > 0, there exist positive constants c1 and c2 such that for each
n≥1
s c1 3/2
p̄n 2 + 3/2 ≤ 1/3 e−c2 s , (3.12)
n n s
s c1 3/2
p̄n −2 − 3/2 ≤ 1/3 e−c2 s .
n n s
(iii) Given |x| > 2 + δ with δ > 0, there exist positive constants c3 and c4
such that for each n ≥ 1
3/2
p̄n (x) ≤ c3 e−c4 n|x| .
Proof. To start with the proof of (i). Note it follows from (ii) and (iii)
of Lemma 3.1
1 √ √
p̄n (x) = √ Kn nx, nx
n
√ √ √ √
= ϕ0n ( nx)ϕn−1 ( nx) − ϕn ( nx)ϕ0n−1 ( nx)
√ √ √ √ √
= nϕn−1 ( nx)2 − n − 1ϕn ( nx)ϕn−2 ( nx). (3.13)
Fix a δ > 0. For each x such that |x| ≤ 2 + δ, we have by Lemma 3.1 (i)
√ √ 2 π
nϕn−1 ( nx)2 = cos 2
nα(θ) − + O(n−1 )
π(4 − x2 )1/2 4
and
√ √ √
nϕn ( nx)ϕn−2 ( nx)
2 θ π θ π
= 2 1/2
cos nα(θ) + − cos nα(θ) − − + O(n−1 ),
π(4 − x ) 2 4 2 4
where x = 2 cos θ, α(θ) = θ − sin 2θ/2 and the convergence is uniform in x.
Now a simple algebra yields
1 p
p̄n (x) = 4 − x2 + O(n−1 ),
2π
as desired.
Proceed to prove (ii). Note that it follows by Lemma 3.1
d √
Kn (x, x) = n ϕ00n (x)ϕn−1 (x) − ϕn (x)ϕ00n−1 (x)
dx √
= − nϕn (x)ϕn−1 (x).
Since Kn (x, x) vanishes at ∞, we have
Z ∞
√
Kn (x, x) = n ϕn (u)ϕn−1 (u)du,
x
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 94
Now we are ready to state and prove the celebrated Wigner semicircle
law for the GUE. Define the empirical spectral distribution for normalized
eigenvalues by
n
1X
Fn (x) = 1(λk ≤√nx) , −∞ < x < ∞. (3.15)
n
k=1
Proposition 3.2 gives the limit of the mean spectral density. In fact, we can
further prove the following
Proof. The statement (3.16) means that for any bounded continuous
function f ,
Z ∞ Z ∞
P
f (x)dFn (x) −→ f (x)ρsc (x)dx, n → ∞. (3.17)
−∞ −∞
and
n
1 X λ
k
V ar f √ → 0. (3.19)
n n
k=1
Fix a small δ > 0 and let δn = sn /n2/3 satisfy δn n2/3 → ∞ and δn n1/2 → 0.
Write the integral on the righthand side of (3.20) as the sum of integrals
Ik , k = 1, 2, 3, 4, over the sets A1 = {x : |x| < 2−δ}, A2 = {x : 2−δ ≤ |x| <
2 − δn }, A3 = {x : 2 − δn ≤ |x| < 2 + δn } and A4 = {x : 2 + δn ≤ |x| < ∞}.
March 5, 2015 15:59 9197-Random Matrices and Random Partitions ws-book9x6 page 96
≤ 4κ2 n1/2 δn → 0.
To estimate the integral over A4 , we use Proposition 3.2 (ii) to get
Z Z ∞
s s
p̄n (x)dx ≤ p̄n 2 + 2/3 + p̄n −2 − 2/3 ds.
I4 sn n n
→ 0.
Combining the above four estimates together yields
λ Z 2−δ
1
lim Ef √ = f (x)ρsc (x)dx + O(δ).
n→∞ n −2+δ
Letting δ → 0, we can conclude the proof of (3.18).
Next we turn to the proof of (3.19). Observe
Xn λ
k
V ar f √
n
k=1
h λ 2 λ 2 i
1 1
= n Ef √ − Ef √
n n
h λ λ λ 2 i
1 2 1
+ n(n − 1) Ef √ f √ − Ef √ . (3.21)
n n n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 97
Note
λ 2 Z ∞
1
Ef √ = f (x)2 p̄n (x)dx
n −∞
Z ∞
1 √ √
= √ f (x)2 Kn ( nx, nx)dx
n −∞
Z ∞Z ∞
√ √
= f (x)2 Kn ( nx, ny)2 dxdy (3.22)
−∞ −∞
and
λ λ
1 2
Ef √ f √
n n
Z ∞Z ∞
√ √
= f (x)f (y)npn,2 ( nx, ny)dxdy
−∞ −∞
Z ∞Z ∞
1 √ √
= f (x)f (y) det Kn ( nx, ny) dxdy
n − 1 −∞ −∞
1 ∞ √ √
Z 2
= f (x)Kn ( nx, nx)dx
n − 1 −∞
Z ∞Z ∞
1 √ √
− f (x)f (y)Kn ( nx, ny)2 dxdy. (3.23)
n − 1 −∞ −∞
Substituting (3.22) and (3.23) into (3.21) yields
X n λ
k
V ar f √
n
k=1
Z ∞Z ∞
1 √ √
= (f (x) − f (y))2 Kn ( nx, ny)2 dxdy
2 −∞ −∞
n ∞ ∞ f (x) − f (y) 2
Z Z
=
2 −∞ −∞ x−y
√ √ √ √ 2
ϕn ( nx)ϕn−1 ( ny) − ϕn ( ny)ϕn−1 ( nx) dxdy.
Since f is 1-Lipschitz function, it follows by the orthogonality of ϕl
X n λ
k
V ar f √
n
k=1
Z ∞Z ∞
n √ √ √ √ 2
≤ ϕn ( nx)ϕn−1 ( ny) − ϕn ( ny)ϕn−1 ( nx) dxdy
2 −∞ −∞
= 1, (3.24)
which implies the claim (3.19).
Now we conclude the proof of Theorem 3.2.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 98
Proof. (i) trivially follows form the fact Gn (Hn − z) = In . To prove (ii),
let U = (uij )n×n be a unitary matrix such that
λ1 0 · · · 0
0 λ2 · · · 0
∗
Hn = U . . . . U . (3.25)
.. .. . . ..
0 0 · · · λn
Then
(λ1 − z)−k ···
0 0
0 (λ 2 − z)−k ··· 0
∗
Gkn = U U .
.. .. .. ..
. . . .
0 0 · · · (λn − z)−k
Hence we have
n
X
(Gkn )ij = uil (λl − z)−k u∗lj ,
l=1
from which it follows
n
(Gn )ij ≤ 1 1
X
|uil ||u∗lj | ≤ k .
k
|η|k |η|
l=1
We conclude (ii).
Finally, (iii) easily follows from the Sherman-Morrison equation
1 1 1 1
− = −δ A .
Hn + δA − z Hn − z Hn + δA − z Hn − z
The next lemma collects some important properties that will be used below
about Gaussian random variables.
Lemma 3.4. Assume that g1 , g2 , · · · , gm are independent centered normal
random variables with Egk2 = σk2 . Denote σ 2 = max1≤k≤m σk2 .
(i) Stein equation: if F : Rm 7→ C is a differentiable function, then
∂F
Egk F (g1 , · · · , gm ) = σk2 E (g1 , · · · , gm ).
∂gk
(ii) Poincaré-Nash upper bound: if F : Rm 7→ C is a differentiable function,
then
2
E F (g1 , · · · , gm ) − EF (g1 , · · · , gm ) ≤ σ 2 E|∇F |2
Now we can use the above two lemmas to get a rough estimate for Emn (z)
and V ar(mn (z)).
Proposition 3.3. For each z with Imz 6= 0, it follows
(i)
Emn (z) = msc (z) + O(n−2 ), (3.26)
where msc (z) denotes the Stieltjes transform of ρsc , namely
z 1p 2
msc (z) = − + z − 4;
2 2
(ii)
E|mn (z) − Emn (z)|2 = O(n−2 ). (3.27)
Proof. Start with the proof of (3.27). Note that mn is a function of
independent centered normal random variables {Hii , 1 ≤ i ≤ n} and
{ReHij , ImHij , 1 ≤ i < j ≤ n}. We use the Poincaré-Nash upper bound
to get
n
1 hX ∂mn 2 X ∂mn 2
E|mn − Emn |2 ≤ E + E
n i=1 ∂Hii ∂ReHij
i<j
X ∂mn 2 i
+ E . (3.28)
i<j
∂ImHij
It easily follows from the differential relations in Lemma 3.3 (iii) that
n
∂mn 1X
=− Gli Gil ,
∂Hii n
l=1
n
∂mn 1X
=− (Gli Gjl + Glj Gil ),
∂ReHij n
l=1
n
∂mn 1 X
= −i (Gli Gjl − Glj Gil ).
∂ImHij n
l=1
In turn, according to Lemma 3.3 (ii), we have (3.27).
Proceed to the proof of (3.26). First, use the matrix identity to get
Emn
n
1X
= EGii
n i=1
n n
1X 1 1X
= E − + Gik Hki
n i=1 z z
k=1
n n
1 1 X 1 X
=− + EGii Hii + EGik ReHki + iImHki . (3.29)
z zn i=1 zn
i6=k
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 101
and so we have
n
1 X Emn
1 + O(n−1 ) .
EGik Gki = − (3.40)
n z + 2Emn
i,k=1
zEmn + (Emn )2
=− (1 + o(1))
(z + 2Emn )3
1 + o(1)
= ,
(z + 2Emn )3
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 104
where in the last step we used the fact zEmn + (Emn )2 = −1 + o(1).
Next, we turn to prove (3.37). Since the proof is very similar to (3.36),
we only give some main steps. It follows by the matrix identity
We now have
Applying once again the matrix identity and Stein equation, we obtain
n
1 X
EGkl (z1 )Gli (z1 )Gik (z2 )
n
i,k,l=1
n
1 X
=− EGkl (z1 )Glk (z1 )
z2 n
k,l=1
n n
1 1 X 1 X
− E Gkl (z1 )Glk (z1 ) E Gkl (z1 )Glk (z2 )
z2 n n
k,l=1 k,l=1
n
1 1 X
− Em (z
n 1 ) EGkl (z1 )Gli (z1 )Gik (z2 )
z2 n n
i,k,l=1
n
1 1 X
− Emn (z2 ) EGkl (z1 )Gli (z1 )Gik (z2 ) + o(1).
z2 n n
i,k,l=1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 106
In combination, we have
Emn (z1 )mn (z2 ) − Emn (z1 )Emn (z2 )
1 + o(1) Emn (z1 )
=−
z2 + 2Emn (z2 ) z2 + Emn (z2 ) + Emn (z1 )
1 Emn (z1 ) 1
× 1− . (3.44)
z1 + 2Emn (z1 ) z2 + Emn (z2 ) + Emn (z1 ) n2
To simplify the righthand side of (3.44), we observe the following asymptotic
formulae
Emn (z) = msc (z)(1 + o(1))
and
Emn (z1 ) Emn (z2 )
= (1 + o(1)).
z2 + Emn (z2 ) + Emn (z1 ) z1 + Emn (z2 ) + Emn (z1 )
Thus a simple calculus now easily yields
Emn (z1 )mn (z2 ) − Emn (z1 )Emn (z2 )
1 + o(1) 2
=p p p p
z1 − 4 z2 − 4 z1 − 4 + z22 − 4 + z1 − z2
2 2 2
2 1
×p 2
z1 − 4 + z2 − 4 − (z1 − z2 ) n2
p
2
1 + o(1) 2 1
=p
z12 − 4 z22 − 4 z1 z2 − 4 + (z12 − 4)(z22 − 4) n2
p p
p
1 + o(1) z1 z2 − 4 − (z12 − 4)(z22 − 4) 1
= p 2 ,
(z1 − z2 )2 n2
p
2 z1 − 4 z22 − 4
as desired.
We have so far proved a kind of law of large numbers for mn (z) and provided
a precise estimate of Emn (z) and V ar(mn (z)). Having these, one may ask
how mn (z) fluctuates around its average. In the rest of this section we will
deal with such a issue. It turns out that mn (z) asymptotically follows a
normal fluctuation. Moreover, we have
Theorem 3.4. Define a random process by
ζn (z) = n(mn (z) − Emn (z)), z ∈ C \ R.
Then there is a Gaussian process Ξ = {Ξ(z), z ∈ C \ R} with the covariance
structure
1 z1 z2 − 4
Cov Ξ(z1 ), Ξ(z2 ) = 2
p p − 1
2(z1 − z2 ) z12 − 4 · z12 − 4
such that
ζn ⇒ Ξ, n → ∞.
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 107
Similarly,
∂(m (z ) − m (z )) 2
n 1 n 2
∂ReHij
n
4 X 1 1
≤ 2 2
− 2
n (λk − z1 ) (λk − z2 )
k,l=1
1 1
× − Re(uik u∗jk )Re(uil u∗jl ) (3.47)
(λl − z̄1 )2 (λl − z̄2 )2
and
∂(m (z ) − m (z )) 2
n 1 n 2
∂ImHij
n
4 X 1 1
≤ 2 −
n (λk − z1 )2 (λk − z2 )2
k,l=1
1 1
× − Im(uik u∗jk )Im(uil u∗jl ). (3.48)
(λl − z̄1 )2 (λl − z̄2 )2
Substituting (3.46)-(3.48) into (3.45) yields
n
4 X 1 1
E|mn (z1 ) − mn (z2 )|2 ≤ 3 2
− 2
n (λk − z1 ) (λk − z2 )
i,j,k,l=1
1 1
× − uik u∗jk u∗il ujl .
(λl − z̄1 )2 (λl − z̄2 )2
Note by orthonormality
Xn Xn
uik u∗il = δk,l , ujk u∗jl = δk,l
i=1 j=1
n
X
uik u∗jk u∗il ujl = n.
i,j,k,l=1
We have
n 2
4 X
2 1 1
E|mn (z1 ) − mn (z2 )| ≤ 3 −
n (λk − z1 )2 (λk − z2 )2
k=1
4
≤ 2 6 |z1 − z2 |2 ,
n η
from which we can establish the uniform tightness for mn (z).
Proceed to proving finite dimensional distribution convergence. Fix
z1 , z2 , · · · , zq ∈ C \ R. It is enough to prove
d
(ζn (z1 ), · · · , ζn (zq )) −→ (Ξ(z1 ), · · · , Ξ(zq )), n → ∞.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 109
Note by (3.34)
1 + 3(Emn )2 Em
n 1
1− = −(z + 2Em n ) − + o(1).
z2 z2 z
In combination with (3.39), it is now easy to see that (3.49) holds true, as
desired.
To conclude this section, let us turn to linear eigenvalue statistics. This
is a very interesting and well studied object in the random matrix theory.
Let f : R → R be a real valued measurable function. A linear eigenvalue
statistic with test function f is defined by
n
1X
Tn (f ) = f (λi ), (3.52)
n i=1
where the λi ’s are eigenvalues of normalized GUE matrix Hn .
As shown in Theorem 3.2, if f is bounded and continuous, then
Z 2
P
Tn (f ) −→ f (x)ρsc (x)dx, n → ∞.
−2
This is a certain weak law of large numbers for eigenvalues. From a proba-
bilistic view, the next natural issue is to take a closer look at the fluctuation.
Under what conditions could one have asymptotic normality? As a matter
of fact, this is usually a crucial problem in the statistical inference theory.
As an immediate application of Theorem 3.4, we can easily derive a
central limit theorem for a class of analytic functions.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 111
Therefore we have
Z Z
2 1
σf = f (z1 )f (z2 )Cov(Ξ(z1 ), Ξ(z2 ))dz1 dz2
(2πi)2 |z1 |=1 |z2 |=1
Z 2Z 2
xy − 4
Z Z
1 1
= √ dxdy
4π 2 −2 −2 (x − y)2 x2 − 4 y 2 − 4 (2πi)2 |z1 |=1 |z2 |=1
p
1 1 1 1
f (z1 )f (z2 ) − − dz1 dz2
z1 − x z1 − y z2 − x z2 − y
Z 2Z 2
1 xy − 4 (f (x) − f (y))2
= 2
√ dxdy.
(x − y)2
p
4π −2 −2 x2 − 4 y 2 − 4
The proof is now complete.
It has been an interesting issue to study fluctuations of linear eigenvalue
statistics for as wide as possible class of test functions. In Theorem 3.5,
the analyticity hypothesis was only required to use the Cauchy integral
formula. This condition can be replaced by other regularity properties. For
instance, Lytova and Pastur (2009) proved that Theorem 3.5 is valid for
a bounded continuous differentiable test function with bounded derivative.
Johansson (1998) studied the global fluctuation of eigenvalues to manifest
the regularity of eigenvalue distribution. In particular, assume that f :
R 7→ R is not too large for large values of x:
(i) f (x) ≤ L(x2 + 1) for some constant L and all x ∈ R;
(ii) |f 0 (x)| ≤ q(x) for some polynomial q(x) and all x ∈ R;
(iii) For each x0 , there exists an α > 0 such that f (x)ψx0 (x) ∈ H 2+α , where
H 2+α is standard Sobolev space and ψx0 (x) is an infinitely differentiable
function such that |ψx0 (x)| ≤ 1 and
1 |x| ≤ x0 ,
ψx0 (x) =
0 |x| > x0 + 1.
Then (3.53) is also valid with σf2 given by.
Z 2Z 2 0 √
2 1 f (x)f (y) 4 − x2
σf = dxdy. (3.55)
4π 2 −2 −2 (y − x) 4 − y 2
p
Here we note that the righthand sides of (3.54) and (3.55) are equal.
In this section we are further concerned with such an interesting issue like
how many eigenvalues locate in an interval. Consider the standard GUE
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 113
Proposition 3.4.
Z b
ENn (a, b) = n ρsc (x)dx + O(1) (3.56)
a
and
1
V ar Nn (a, b) = 2
log n 1 + o(1) . (3.57)
2π
Proof. (3.56) is trivial since the average spectral density function p̄n (x)
converges uniformly to ρsc (x) in [a, b].
To prove (3.57), note the following variance formula
Z b
√ √ √
V ar(Nn (a, b)) = n Kn nx, nx dx
a
Z b Z b √ √ 2
−n Kn nx, ny dxdy
a a
Z bZ ∞ √ √ 2
= n Kn nx, ny dxdy
a b
Z bZ a √ √ 2
+n Kn nx, ny dxdy
a −∞
=: I1 + I2 .
We shall estimate the integrals I1 and I2 below. The focus is upon I1 , since
I2 is completely similar. A change of variables easily gives
Z ∞ Z b−a
√ √ 2
I1 = n dv Kn n(b − u), n(b − u + v) du
b−a 0
Z b−a v √ √
Z
2
+n dv Kn n(b − u), n(b − u + v) du
0 0
=: I1,1 + I1,2 .
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 114
and
√
kϕl k∞ ≤ κ, kϕ0l k∞ ≤ lκ.
So,
1
Z n
Z v √ √ 2
n dv Kn n(b − u), n(b − u + v) du = O(1).
0 0
For the integral over (1/n, b − a), we use Lemma 3.2 to get
√ √ √ √
n1/2 ϕn ( nx)ϕn−1 ( ny) − ϕn ( ny)ϕn−1 ( nx)
2 1 h θ1 π π
= 2 1/4 2 1/4
cos nα(θ1 ) + − cos nα(θ2 ) −
π (4 − x ) (4 − y ) 2 4 4
θ2 π π i
− cos nα(θ2 ) + − cos nα(θ1 ) − + O(n−1 )
2 4 4
1 (4 − xy)1/2
= + O(n−1 ).
2π (4 − x2 )1/4 (4 − y 2 )1/4
Thus with x = b − u and y = b − u + v,
Z b−a Z v
√ √ 2
n dv Kn n(b − u), n(b − u + v) du
1
n 0
b−a v
4 − xy
Z Z
1
= dv du
4π 2 1
n 0 v 2 (4 − x2 )1/2 (4 − y 2 )1/2
Z b−a Z v
1
+O(n−1 ) dv du.
1
n 0 v2
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 115
Trivially, it follows
Z b−a Z v
1
dv du = O(log n).
1
n 0 v2
We also note
4 − xy
sup ≤ Ca,b
x,y∈(a,b) (4 − x2 )1/2 (4 − y 2 )1/2
for some positive constant Ca,b . Then for any ε > 0
Z b−a Z v
4 − xy
dv du 2 2 )1/2 (4 − y 2 )1/2
= O(| log ε|). (3.58)
ε 0 v (4 − x
On the other hand, it is easy to see
4 − xy
= 1 + O(ε2 ), 0 < v < ε.
(4 − x2 )1/2 (4 − y 2 )1/2
Hence we have
Z ε Z v
4 − xy
dv du 2 2 )1/2 (4 − y 2 )1/2
= (1 + O(ε2 ))(log n + | log ε|). (3.59)
1
n 0 v (4 − x
Combining (3.58) and (3.59) together yields
Z b−a Z v
4 − xy
dv du 2
1
n 0 v (4 − x )1/2 (4 − y 2 )1/2
2
Z ε Z v
4 − xy
= dv du 2 2 )1/2 (4 − y 2 )1/2
n
1
0 v (4 − x
Z b−a Z v
4 − xy
+ dv du 2 2 )1/2 (4 − y 2 )1/2
ε 0 v (4 − x
= (1 + O(ε2 ))(log n + | log ε|).
√Pn
We remark that the linear eigenvalue statistic i=1 f (λi / n) has variance
at most 1 whenever f is a 1-Lipschtiz test function, see (3.24). On the other
hand, the counting function is not a 1-Lipschitz function. The Proposition
3.4 provides a log n-like estimate for the size of variance of Nn (a, b).
Having the Proposition, one would expect the asymptotic normal fluc-
tuations for Nn (a, b). Below is our main result of this section.
The rest of this section is devoted to the proof of Theorem 3.6. In fact, we
shall prove the theorem under a more general setting. To do this, we need to
introduce some basic definitions and properties about determinantal point
processes. Recall a point process X on R is a random configuration such
that any bounded set contains only finitely many points. The law of X
is usually characterized by the family of integer-valued random variables
{NX (A), A ∈ B}, where NX (A) denotes the number of X in A. Besides,
the correlation function is becoming a very useful concept in describing the
properties of point processes. The so-called correlation function was first
introduced to study the point process by Macchi (1975). Given a point
process X , its k-point correlation function is defined by
1
ρk (x1 , · · · , xk ) = lim P (xi − δ, xi + δ) ∩ X 6= ∅, 1 ≤ i ≤ k ,
δ→0 (2δ)k
where x1 , · · · , xk ∈ R. Here we only considered the continuous case, the
corresponding discrete case will be given in Chapter 4.
It turns out that the correlation functions is a powerful and nice tool in
computing moments of NX (A). In fact, it is easy to see
Z
↓k
E NX (A) = ρk (x1 , · · · , xk )dx1 · · · dxk , (3.60)
A⊗k
Also, by (3.60) and the Hadamard inequality for nonnegative definite matrix
Z
↓k
E NX (A) = det KX (xi , xj ) dx1 · · · dxk
A⊗k
Z k k
≤ KX (x1 , x1 ))dx1 = ENX (A) .
A
Therefore, we have
∞
X (et − 1)k k
EetNX (A) ≤ ENX (A) < ∞.
k!
k=0
R
For any compact set D, ENX (D) < ∞ since D
KX (x, x)dx < ∞, so
EetNX (D) < ∞ for all t ∈ R.
X l
Y
= (−1)l−1 (l − 1)! E(Nn )↓|Gj |
G j=1
l
X k! Y
= Q (−1)l−1 (l − 1)! E(Nn )↓τj ,
τi !mτi !
τ 7→k j=1
where τ = (τ1 , · · · , τl ) →
7 k is an integer partition of k, mτi stands for the
multiplicity of τi in τ . We can derive from (1.17)
∞ ∞
X βk X E(Nn )↓k
(et − 1)k = log (et − 1)k
k! k!
k=1 k=0
∞
X γk
= log EetNn = tk . (3.62)
k!
k=1
Equivalently,
k−1
X
γk = βk + bk,j γj (3.63)
j=1
and so
Z
k k
βk + (−1) (k − 1)!γ1 = (−1) (k − 1)! KXn (x1 , x1 )dx1
In
Z
− KXn (x1 , x2 ) · · · KXn (xk , x1 )dx1 · · · dxk .
⊗k
In
Then it follows
βk + (−1)k (k − 1)!γ1 = (−1)k (k − 1)! trKIn − trKIkn
k
X
= (−1)k (k − 1)! trKIl−2 KIn − KI2n .
n
l=2
= k!γ2 ,
which gives (3.64). Now we conclude the proof of the theorem.
As the reader may see, Theorem 3.7 is of great universality for determi-
antal point processes in the sense that there is almost no requirement on
the kernel function. However, the theorem itself does not tell what the ex-
pectation and variance of NXn (In ) look like. To have numerical evaluation
of expectation and variance, one usually needs to know more information
about the kernel function. In the case of GUE, the kernel function is given
by Hermite orthogonal polynomials so that we can give precise estimates
of expectation and variance. This was already done in Proposition 3.4.
It is believed that Theorem 3.7 would have a wide range of applications.
We only mention the work of Gustavsson (2005), in which he studied the
kth greatest eigenvalue λ(k) of GUE model and used Theorem 3.7 to prove
the λ(kn ) after properly scaled has a Gaussian fluctuation around its av-
erage as kn /n → a ∈ (0, 1). He also dealt with the case of kn → ∞ and
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 122
and so
∞ Z
X
Kf (x) = f (y)φk (y)dy Kφk (x)
k=1 I
Z ∞
X
= f (y) qk φk (y)φk (x)dy
I k=1
Z
= f (y)K I (x, y)dy.
I
P∞ ∞
Y
Eet k=1 ξk
= Eetξk
k=1
Y∞
= 1 + qk (et − 1)
k=1
∞
X X
= 1+ qi1 · · · qik (et − 1)k .
k=1 1≤i1 <···<ik <∞
For k ≥ 1,
q1 φ1 (x1 ) q2 φ2 (x1 ) · · · qn φn (x1 ) · · ·
q1 φ1 (x2 ) q2 φ2 (x2 ) · · · qn φn (x2 ) · · ·
K I (xi , xj ) k×k
=
.. .. .. ..
. . ··· . .
q1 φ1 (xk ) q2 φ2 (xk ) · · · qn φn (xk ) · · ·
φ1 (x1 ) φ1 (x2 ) · · · φ1 (xk )
φ2 (x1 ) φ2 (x2 ) · · · φ2 (xk )
× ... .. ..
. · · · .
φn (x1 ) φn (x2 ) · · · φn (xk )
··· ··· ··· ···
=: AB. (3.68)
Then according to the Cauchy-Binet formula
X
det K I (xi , xj ) k×k =
det(Ak Bk ), (3.69)
1≤i1 <···<ik
In this section we are concerned with the asymptotic behaviors of the log-
arithm of the determinant of the GUE matrix. Let An = (zij )n×n be the
standard GUE matrix as given in the Introduction, denote its eigenvalues
by λ1 , λ2 , · · · , λn . Then we have
Theorem 3.8. As n → ∞,
1 1 1
d
q log | det An | − log n! + log n −→ N (0, 1). (3.71)
1 2 4
2 log n
Proposition 3.6. As n → ∞,
1 1 1
d
q log | det Mn | − log n! + log n −→ N (0, 1). (3.74)
1 2 4
4 log n
where the χi is a chi random variable with index i, and all chi random
variables are independent.
Indeed, let y1 , y2 , · · · , yn denote row vectors of Mn . Then the abso-
lute value of the determinant of Mn is equal to the volume of parallelnoid
consisting of vectors y1 , y2 , · · · , yn . In turn, the volume is equal to
|y1 | · |(I − P1 )y2 | · · · |(I − Pn−1 )yn |, (3.76)
where Pi is an orthogonal projection onto the subspace spanned by vectors
{y1 , y2 , · · · , yi }, 1 ≤ i ≤ n − 1. Note Pi is an idempotent projection
with rank i, so Pi yi+1 is an i-variate complex standard normal random √
vector and√is independent of {y1 , y2 , · · · , yi }. Then letting χ2n = 2|y1 |,
χ2(n−i) = 2|(I − Pi )yi+1 |, 1 ≤ i ≤ n − 1 conclude the desired identity.
Second, note the χi has a density function
21−i/2 i−1 −x2 /2
x e , x>0
Γ(i/2)
then it is easy to get
Γ((i + k)/2)
Eχki = 2k
Γ(i/2)
and the following asymptotic estimates
1 1 1
+ O i−2 , V ar(log χi ) = + O i−2 .
E log χi = log i −
2 2i 2i
In addition, for each positive integer k ≥ 1
E(log χi − E log χi )2k = O i−k .
(3.74) now directly follows from the classic Lyapunov CLT for sums of
independent random variables.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 127
The proof of Proposition 3.6 is simple and elegant. The hypothesis that all
entries are independent plays an essential role. It is no longer true for An
since it is Hermitian. We need to adopt a completely different method to
prove Theorem 3.8. Start with a tridiagonal matrix representation of An .
Let an , n ≥ 1 be a sequence of independent real standard normal random
variables, bn , n ≥ 1 a sequence of independent random variables with each
bn distributed like χn . In addition, assume an ’s and bn ’s are all independent.
For each n ≥ 1, construct a tridiagonal matrix
an bn−1 0 0 ··· 0
n−1 an−1 bn−2 0 · · · 0
b
0 bn−2 an−2 bn−3 · · · 0
Dn = .. .. . . . . .. . (3.77)
0 . . . . .
0 · · · b2 a2 b1
0
0 0 0 · · · b1 a1
Lemma 3.7. The eigenvalues of Dn are distributed according to (3.2). In
particular,
d
det An = det Dn . (3.78)
where wn = (w1 , w2 , · · · , wn )τ .
It is easy to check that Vn is a unitary matrix and
z11 (z , z , · · · , z1n )Vn,n−1
∗ 12 13
z12
∗
z13
Vn An Vn =
Vn,n−1 . Vn,n−1 An,n−1 Vn,n−1
..
∗
z1n
α |zz12
z11 21 |
0
∗
= α |zz12 | .
21 Vn,n−1 An,n−1 Vn,n−1
0
To make the second entry in the first column nonnegative, we need to add
one further configuration. Let Rn differ from the identity matrix by having
(2,2)-entry e−iφ with φ chosen appropriately and form Rn Vn An Vn Rn∗ . Then
we get the desired matrix
z11 α 0
α ,
Vn,n−1 An,n−1 Vn,n−1
0
where α2 = |z21 |2 + |z31 |2 + · · · + |zn1 |2 .
Define an = z11 , bn−1 = α. Vn,n−1 is also a unitary matrix and is
independent of An,n−1 , so Vn,n−1 An,n−1 Vn,n−1 is an n − 1 × n − 1 GUE
matrix. Repeating the preceding procedure yields the desired matrix Dn .
The proof is complete.
According to (3.78), it suffices to prove (3.71) for log | det Dn | below. Let
dn = det Dn . It is easy to see the following recurrence relations
dn = an dn−1 − b2n−1 dn−2 (3.79)
dn−1 = an−1 dn−2 − b2n−2 dn−3 .
(3.80)
√ √
Let en = dn / n! and cn = (b2n − n)/ n. Note cn−k is asymptotically
normal as n − k → ∞. So we deduce from (3.79) and (3.80)
an cn−1 1
en = √ en−1 − 1 + √ − en−2 + 1 (3.81)
n n 2n
an−1 cn−2 1
en−1 = √ en−2 − 1 + √ − en−3 + 2 , (3.82)
n n 2n
where and in the sequel 1 , 2 denote a small negligible quantity, whose
value may be different from line to line.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 129
Lemma 3.8. As n → ∞,
log rn + 14 log n d
q −→ N (0, 1). (3.84)
1
2 log n
Lemma 3.9. Let θn ∈ (0, 2π) be such that (e2n , e2n−1 ) = rn (cos θn , sin θn ).
Then as n → ∞,
d
θn −→ Θ, (3.87)
Proof. Let us first look at the difference between θn and θn−1 . Rewrite
(3.83) as
cos θn cos θn−1
rn = rn−1 (−I2 + Dn ) ,
sin θn sin θn−1
where
1 1
Dn = − √ Sn,1 + Sn,2 +
n n
where is a negligible matrix. It then follows
rn cos θn cos θn−1 cos θn−1
=− + Dn . (3.88)
rn−1 sin θn sin θn−1 sin θn−1
Note (cos θn−1 , sin θn−1 ) and (− sin θn−1 , cos θn−1 ) form an orthonormal
basis. It is easy to see
cos θn−1 cos θn−1 − sin θn−1
Dn = xn + yn , (3.89)
sin θn−1 sin θn−1 cos θn−1
where the coefficients xn and yn are given by
cos θn−1
xn = (cos θn−1 , sin θn−1 )Dn
sin θn−1
and
cos θn−1
yn = (− sin θn−1 , cos θn−1 )Dn .
sin θn−1
Substituting (3.89) back into (3.88) yields
rn cos θn cos θn−1 − sin θn−1
= (−1 + xn ) + yn .
rn−1 sin θn sin θn−1 cos θn−1
Thus it is clear that
yn
tan(θn − θn−1 ) = ,
−1 + xn
which in turn leads to
yn
θn − θn−1 = arctan .
−1 + xn
Next we estimate xn and yn . There is a constant ς > 0 such that
xn = OP (n−ς ), yn = OP (n−ς ).
In addition, for some ι > 1
E xn θn−1 = O(n−ι ), E yn θn−1 = O n−ι
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 132
1
E yn2 θn−1 = + O n−ι .
2n
Using the Taylor expansions of arctan x and (1 + x)−1 we obtain
θn − θn−1 = −(yn − xn yn ) + O yn3 ,
and so
k2
eikθn = eikθn−1 1 − ikyn + ikyn xn − yn2 + O yn2 (xn + yn ) .
2
Hence we have
k2
Eeikθn = Eeikθn−1 1 − ikyn + ikyn xn − yn2 + O yn2 (xn + yn ) .
2
Moreover, using conditioning argument we get
k2
Eeikθn = Eeikθn−1 1 − + O n−ι .
(3.90)
4n
Let mn → ∞ and mn /n → 0. Then repeatedly using (3.90) to get
n n
Y k 2 ikθmn X
Eeikθn = 1− Ee +O l−ι
4l
l=mn +1 l=mn +1
→ 0, n→∞
Qn 2
where in the last limit we used ι > 1 and the fact that l=mn +1 1− k4l → 0
Pn 2
since l=mn +1 1 − k4l → ∞. Thus we complete the proof of (3.87).
The analog is valid for e2n−1 . Therefore we have proven Theorem 3.8.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 133
uniformly on any fixed compact set in the half plane Reα > −1/2. Here
1 Z α 1
C(α) = 1 exp 2 log Γ(s + )ds + α2
Γ(α + 2 ) 0 2
2 G(α + 1)2
= 22α , (3.94)
G(2α + 1)
where G(α) is Barnes’s G-function. The remainder term εα,n = O(log n/n)
is analytic in α.
The proof is omitted. For α positive integers (3.93) has been found by
Brézin and Hikami (2000), Forrester and Frankel (2004). For such α,
√
E| det(An − z n)|2α can be reduced to the Hermite polynomials and their
derivatives at the points z. However, it is not the case for noninteger α.
In order to obtain (3.93), Krasovsky (2007) used Riemann-Hilbert problem
approach to compute asymptotics of the determinant of a Hankel matrix
whose support is supported on the real line and possesses power-like singu-
larity.
Proof of Theorem 3.10. Start with computing expectation and variance
√ √ 2α
of log det(An −z n). For simplicity, write M (α) for E det(An −z n)
below, and set
M (α) = A(α)B(α),
where
2 2
z 2 α /2 n αn+α (z2 /2−1)αn
A(α) = 2αn 1 − e
4 2
and
B(α) = C(α)(1 + εα,n ).
It obviously follows
√ 2
E log det(An − z n) = M 0 (0)
and
√ 2
E log | det(An − z n)|2 = M 00 (0).
and
n z2 α2 z 2
M 00 (α) = 2 n log n + 2α log + −1 n+ 1− A(α)B 0 (α)
2 2 2 4
n z2 α2 z 2 2
+ n log n + 2α log + −1 n+ 1− A(α)B(α)
2 2 2 4
n 1 z 2
+ 2 log + 1− A(α)B(α) + A(α)B 00 (α).
2 2 4
It is easy to see M (0) = A(0) = 1, and so B(0) = 1. Furthermore, by the
analyticity of B(α) for Reα > −1/2 and using Cauchy’s theorem, we have
B 0 (0) = C 0 (0) + O(n−1 log n), B 00 (0) = C 00 (0) + O(n−1 log n).
Similarly, it follows from (3.94)
G0 (α + 1) G0 (2α + 1)
C 0 (α) = C(α) 4 log 2α + 2 −2 .
G(α + 1) G(2α + 1)
Note
G0 (α + 1) 1 1 Γ0 (α)
= log(2π) + − α + α
G(α + 1) 2 2 Γ(α)
1 1 π2
= log(2π) − − (γ + 1)α + α + O(α3 ),
2 2 6
where γ is the Euler constant. So we have
C 0 (0) = 0, C 00 (0) = 4 log 2 + 2(γ + 1).
In combination, we obtain
z2
M 0 (0) = n log n + ( − 1)n + B 0 (0)
2
and
z2 2 z2
M 00 (0) = n log n + ( − 1)n + 2 n log n + ( − 1)n B 0 (0)
2 2
n 1 z 2
+ 2 log + 1− + B 00 (0).
2 2 4
This in turn gives
√ 1 z2
E log det(An − z n) = n log n + ( − 1)n + o(1) (3.95)
2 2
and
√
V ar log | det(An − z n)|
1 z2 1 1 5
= log n − − log 2 + γ + + o(1). (3.96)
2 32 2 2 8
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 136
In the last section of this chapter, we will turn to the study of the Hermite
β Ensemble (HβE), which is a natural extension of the GUE. By the HβE
we mean an n-point process in the real line R with the following joint
probability density function
n
Y Y 2
pn,β (x1 , · · · , xn ) = Zn,β |xi − xj |β e−xj /2 , (3.98)
1≤i<j≤n j=1
Lemma 3.10.
Qn−1 i
Y (n) (n) yi
λi − λj = Qi=1
n .
1≤i<j≤n i=1 qi
Hence it follows
n−1 n Y n−2
n−1
(n) (n−1) 2(n−1) (n−2) (n−1)
YY Y
λi − λj = (−1)n−1 yn−1 λi − λj
j=1 i=1 j=1 i=1
n−1
Y
= (−1)n(n−1)/2 yl2l . (3.101)
l=1
that is
n
(n)
X Y
qi2
Pn−1 (λ) = λl −λ .
i=1 l6=i
Hence it follows
n n−1
(n−1) (n)
Y Y
λi − λj
j=1 i=1
n
(n) (n)
Y Y
= qj2 λi − λj
j=1 i6=j
n
(n) (n) 2
Y Y
= (−1)n(n−1)/2 qj2 λi − λj . (3.102)
j=1 1≤i<j≤n
Lemma 3.11.
Qn−1
i=1 yi
J= Q n . (3.103)
qn i=1 qi
Proof. Observe the following identity
n
−1 X qi2
In − λXn 11 = (n)
.
i=1 1 − λλi
Use the Taylor expansion of (1 − x)−1 to get
∞ ∞ X n
(n)k
X X
k k
qi2 λi λk .
1+ λ Xn 11 =
k=1 k=0 i=1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 140
1 2n−1 Y 1 2 n−1
n Y βi−1 Qn−1 yi
− 2 λi i=1
= e yi n
(2π)n/2 n−1 βi
Q
q i=1 qi
Q
i=1 Γ( )
2 i=1 i=1 n
n n
1 2n−1 Y
− 12 λ2i
Y
β 1
Y
= e (λi − λj ) qiβ−1 . (3.106)
(2π)n/2 n−1 βi q
Q
i=1 Γ( )
2 i=1 1≤i<j≤n
n i=1
We see from (3.106) that (λ1,β , λ2,β , · · · , λn,β ) and q = (q1 , q2 , · · · , qn−1 )
are independent and so the joint probability density pn,β (λ) of (λ1,β , λ2,β ,
· · · , λn,β ) can be obtained by integrating out the variable q
pn,β (λ1 , · · · , λn )
n
1 2n−1 Y 2 Y
= n/2 n−1 βi
e−λi /2 (λi − λj )β
(2π)
Q
i=1 Γ( )
2 i=1 1≤i<j≤n
Z n
1 Y
· qiβ−1 dq1 · · · dqn
q:qi >0, i=1 qi2 =1 qn i=1
Pn
n
1 Γ( β2 )n Y Y 2
= n/2 n βi
(λ i − λ j )β
e−λi /2 , (3.107)
(2π)
Q
i=1 Γ( 2 ) 1≤i<j≤n i=1
where we used Lemma 2.18 to compute the integral in the second equation.
Finally, to obtain the joint probability density of unordered eigenvalue,
we only need to multiply (3.107) by the factor 1/n!. The proof is now con-
cluded.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 142
The HβE is a rich and well studied model in random matrix theory.
It possesses many nice properties similar to the GUE. In particular, there
have been a lot of new advances in the study of asymptotic behaviours of
point statistics since the tridiagonal matrix model was discovered. We will
below quickly review some results related to limit laws without proofs. The
interested reader is referred to original papers for more information.
√ √
To enable eigenvalues
q asymptotically fall in the interval (−2 n, 2 n),
we consider Hn,β =: β2 Dn,β . Denote by λ1,β , λ2,β , · · · , λn,β the eigenval-
ues of Hn,β , the corresponding empirical distribution function is
n
1X √
Fn,β (x) = 1 .
n i=1 (λi,β ≤ nx)
Dumitriu and Edelman (2002) used moment methods to prove the Wigner
semicircle law as follows
d
Fn,β −→ ρsc in P.
In particular, for each bounded continuous function f ,
n Z 2
1 X λi,β P
f √ −→ f (x)ρsc dx.
n i=1 n −2
Based on variational analysis, Ramı́rez, Rider and Virág (2011) proved that
√
under the assumption n1/6 (2 n − un ) → a ∈ R,
d
Nn,β (x) −→ NAiryβ (x), x ∈ R,
where Airyβ is defined as −1 times the point process of eigenvalues of the
stochastic with parameter β, and NAiryβ (x) is the number of points between
0 and x.
In the same spirit, Valkó and Virág (2009) considered the eigenvalues
around any location away from the spectral edge. Let un be a sequence of
√
real numbers so that n1/6 (2 n − |un |) → ∞. Define for x ∈ R
p
Nn,β (x) = ] 1 ≤ i ≤ n : 4n − u2n (λi,β − un ) fall between 0 and x ,
then
d
Nn,β (x) −→ NSineβ (x), x ∈ R,
where Sineβ is a translation invariant point process given by the Brownian
carousel.
As the reader may see, the point process from the HβE is no longer
determinantal except in special cases. Thus Theorem 3.7, the Costin-
Lebowitz-Soshnikov theorem, is not applicable. However, we can follow
the strategy of Valkó and Virág (2009) to prove the central limit theorem
for the number of points of the HβE lying in the right side of the origin,
see and (2010).
Theorem 3.12. Let Nn (0, ∞) be the number of points of the HβE lying in
the right side of the origin. Then it follows
1 n d
q Nn (0, ∞) − −→ N (0, 1). (3.108)
1
log n 2
βπ 2
We remark that the number Nn (0, ∞), sometimes called the index, is a key
object of interest to physicists. Cavagna, Garrahan and Giardina (2000)
calculated the distribution of the index for GOE by means of the replica
method and obtained Gaussian distribution with asymptotic variance like
log n/π 2 . Majumdar, Nadal, Scardicchio and Vivo (2009) further computed
analytically the probability distribution of the number Nn [0, ∞) of positive
points for HβE using the partition function and saddle point analysis. They
computed the variance log n/βπ 2 + O(1), which agrees with the correspond-
ing variance in (3.108), while they thought the distribution is not strictly
Gaussian due to an usual logarithmic singularity in the rate function.
The rest part of the section will prove Theorem 3.12. The proof relies
largely on the new phase evolution of eigenvectors invented by Valkó and
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 144
p
Virág (2009). Let Hn,β = 2/βDn,β . We only need to consider the number
of positive eigenvalues of Hn,β . A key idea is to derive from the tridiagonal
matrix model a recurrence relation for a real number Λ to be an eigenvalue,
which yields an evolution relation for eigenvectors. Specifically, let sj =
p
n − j − 1/2. Define
d11 0 0 · · · 0
0 d22 0 · · · 0
On = . . . . . ,
.. .. .. . . ..
0 0 0 · · · dnn
where
bn−1−i
d11 = 1, dii = di−1,i−1 , 2 ≤ i ≤ n.
si−1
Let
an−i
Xi = √ , 0≤i≤n−1
β
and
b2n−1−i
Yi = − si , 0 ≤ i ≤ n − 2.
βsi+1
Then
X0 s0 + Y0 0 ··· 0
s1 X1 s1 + Y1 · · · 0
On−1 Hn,β On = 0
s2 X2 · · · 0
. .. .. .. ..
.. . . . .
0 0 0 · · · Xn−1
obviously have the same eigenvalues as Hn,β . However, there is a significant
difference between these two matrices. The rows between On−1 Hn,β On are
independent of each other, while Hn,β is symmetric so that the rows are
not independent.
Assume that Λ is an eigenvalue of On−1 Hn,β On , then by definition there
exists a nonzero eigenvector v = (v1 , v2 , · · · , vn )τ such that
On−1 Hn,β On v = Λv.
Without loss of generality, we can assume v1 = 1. Thus, Λ is an eigenvalue
if and only there exists an eigenvector vτ = (1, v2 , · · · , vn ) such that
X0 s0 + Y0 0 ··· 0 1 1
s1 X1 s1 + Y1 · · · 0 v2 v2
0
s2 X2 · · · 0 v3 = Λ v3 .
. .. .. .. . . .
.. . .. .. ..
. .
0 0 0 · · · Xn−1 vn vn
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 145
and
Λ
Rl,Λ = Q(π)A 1, Wl , 0 ≤ l ≤ n − 1.
sl
With these notations, the evolution of r in (3.109) becomes
rl+1 = rl• Rl,Λ , 0≤l ≤n−1
and Λ is an eigenvalue if and only if
∞• R0,Λ · · · Rn−1,Λ = 0.
For 0 ≤ l ≤ n define
ϕ̂l,Λ = π∗ R0,Λ · · · Rl−1,Λ , ϕ̂ −1 −1
l,Λ = 0∗ Rn−1,Λ · · · Rl,Λ ,
then
ϕ̂l,Λ = ϕ̂
l,Λ mod 2π.
The following lemma summarizes nice properties about ϕ̂ and ϕ̂ , whose
proof can be found in Valkó and Virág (2012).
Note that for any finite λ, Ll,λ and Wl become infinitesimal in the n → ∞
limit while Jl does not. Let
1
Tl = A , −Re(ρl ) ,
Im(ρl )
then
−1
Jl = Q(−2arg(ρl ))Tl
Ql = Q(2arg(ρ0 )) . . . Q(2arg(ρl ))
and
Zl,λ = i• S−1
l,λ − i
= i• T−1
l+1 (Lλ Wλ )
−1
Tl − i
= vl,λ + Vl ,
where
λ ρl+1 − ρl Xl + ρl+1 Yl
vl,λ = − √ √ + , Vl = √ .
2 n0 n0 − l Im(ρl ) n0 − l
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 148
Then
∆ϕl,λ = ash(Sl,λ , −1, z η̄)
i(1 + z̄η)2 2
= Re −(1 + z̄η)Z − Z + O(Z 3 )
4
2
ImZ
= −ReZ + + η terms + O(Z 3 ),
4
where we used Z = Zl,λ , η = ηl and z = eiϕl,λ .
√
Lemma 3.14. Assume λ = λn = o( n). For l ≤ n0 , we have
1 1
osc1 + O (n0 − l)−3/2 ,
E ∆ϕl,λ |ϕl,λ = x = bn + (3.111)
n0 n0
1 1
E (∆ϕl,λ )2 |ϕl,λ = x = osc2 + O (n0 − l)−3/2 , (3.112)
an +
n0 n0
(ii)
φl,∞ = (l + 1)π, φ
l,∞ = −2nπ + 3lπ;
(iii)
φ
l,0 = φ̂l,0 + lπ, φ̂ −1 −1
l,0 = 0∗ Rn−1 · · · Rl .
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 149
φ̂ −1
l,∞ = φ̂l+1,∞∗ Rl,∞ ,
where R−1 −1
l,∞ = Wl A (1, −∞) Q(−π).
By the angular shift formula, we have
φ̂ −1
l,∞ = φ̂l+1,∞∗ Wl A (1, −∞) − π
iφ̂
= φ̂ −1
l+1,∞ + ash Wl A (1, −∞) , −1, e
l+1,∞ −π
!
eiφ̂l+1,∞ ◦ Wl−1 A (1, −∞)
= φ̂
l+1,∞ + arg[0,2π)
−1◦ Wl−1 A (1, −∞)
!
eiφ̂l+1,∞
−arg[0,2π) −π
−1
= φ̂
l+1,∞ + arg[0,2π) (1) − arg[0,2π) (−e
iφ̂l+1,∞
) − π,
φ̂
n−1,∞ = −2π, φ̂
n−2,∞ = −4π, · · · , φ̂l,∞ = −2(n − l)π.
Next we turn to the proof of (ii) and (iii). Since x = 0, then ρl = i, and so
Tl is the identity transform and Ql−1 = Q(lπ) for each 0 < l ≤ n0 . Thus
we have by the fact that Q is a rotation,
where 0 ≤ λ ≤ ∞. Similarly, ϕ
l,λ = ϕ̂l,λ + lπ.
Proof of Theorem 3.12. Let x = 0 and n0 = n − 1/2. Taking l = bn0 c =
n − 1, we have by (3.110)
and
ϕ −1
n−1,0 = 0∗ Rn−1,0 + (n − 1)π.
Also, it follows
0∗ R−1 P
√ n−1,0 −→ 0.
log n
In combination, we need only to prove
ϕ d 4
√n−1,0 −→ N 0, .
log n β
To this end, we shall use the following CLT for Markov chain. Recall that
π = ϕ0,0 , ϕ1,0 , · · · , ϕn−1,0 forms a Markov chain. Let
zl+1 = ∆ϕl,0 − E ∆ϕl,0 |ϕl,0 .
Then z1 , z2 , · · · , zn−1 forms a martingale difference sequence. The martin-
gale CLT implies: if the following conditions are satisfied:
(i)
n−1
X
Bn := Ezl2 → ∞, (3.115)
l=1
(ii)
n−1
1 X P
E zl2 |ϕl−1,0 −→ 1, (3.116)
Bn
l=1
then we have
n−1
1 X d
√ zl −→ N (0, 1).
Bn l=1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 151
4 4
ERe (−1)l+1 e−iϕl,0
= +
β(n0 − l) n0 − l
+O (n0 − l)−3/2 .
Chapter 4
4.1 Introduction
153
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 154
We remark that on the one hand, for many problems it suffices to consider
F(z) as a formal power series in z; on the other hand, much asymptotic
work requires that F(z) be an analytic function of the complex variable z.
The asymptotic theory starts 150 years after Euler, with the first cel-
ebrated letters of Ramanujan to Hardy in 1913. In a celebrated series of
memoirs published in 1917 and 1918, Hardy and Ramanujan found (and
was perfected by Radamacher) very precise estimates for p(n). In particu-
lar, we have
Theorem 4.1.
1 √
p(n) = √ e2c n 1 + o(1) ,
(4.3)
4 3n
√
where and in the sequel c = π/ 6.
The complete proof of Theorem 4.1 can be found in §2.7 of the book Post-
nikov (1988). Instead, we prefer to give a rough sketch of the proof, without
justifying anything. First, using the Cauchy integral formula for the coef-
ficients of a power series, we obtain from (4.1) and (4.2)
Z π
1
r−n e−inθ F reiθ dθ.
p(n) =
2π −π
Choose θn > 0 and split the integral expression for p(n) into two parts:
Z Z
1
r−n e−inθ F reiθ dθ.
p(n) = +
2π |θ|≤θn |θ|>θn
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 155
Up to this point,
0 r has been a free parameter.−vWe now choose r so that
n + log F(r) = 0, i.e., we must choose r = e to satisfy the equation
∞
X k
n= .
ekv − 1
k=1
Note
∞ ∞
kX 1 X kv
= v kv
ekv − 1 v2 e −1
k=1 k=1
Z ∞
1 x
≈ 2 dx
v 0 ex − 1
c2
= 2.
v
√
Thus we must take v so that n ≈ c /v 2 , i.e., v ≈ c/ n. For such a choice,
2
F(r)
Z
2 00
p(n) ≈ n
e−θ (log F (r)) /2 dθ
2πr |θ|≤θn
F(r) ∞ −θ2 (log F (r))00 /2
Z
≈ e dθ
2πrn −∞
F(r)
= p , (4.4)
r n 2π log(F(r))00
using the classical normal integral. To evaluate the value of (4.4), we need
the following lemma, see Postnikov (1988).
Lemma 4.1. Assume Rez > 0 and z → 0, staying within some angle lying
in the right half-plane. Then
c2 1 z
log F(e−z ) = + log + O(|z|). (4.5)
z 2 2π
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 156
√
As a consequence, it easily follows with r = e−c/ n
√ 1 1
log F(r) ≈ c n + log (4.6)
4 24n
and
2
(log F(r))00 ≈ n3/2 . (4.7)
c
Substituting (4.6) and (4.7) into (4.4) yields the desired result (4.3).
Another effective elementary device for studying partitions is graphical
representation. To each partition λ is associated its Young diagram (shape),
which can be formally defined as the set of point (i, j) ∈ Z2 such that
1 ≤ j ≤ λi . In drawing such diagrams, by convention, the first coordinate
i (the row index) increases as one goes downwards, the second coordinate
j (the column index) increases as one goes from the left to the right and
these points are left justified. More often it is convenient to replace the
nodes by unit squares, see Figure 2.1.
Such a representation is extremely useful when we consider applications
of partitions to plane partitions or Young tableaux. Sometimes we prefer
the representation to be upside down in consistency with Descartes coordi-
nate geometry.
The conjugate of a partition λ is the partition λ0 whose diagram is the
transpose of the diagram λ, i.e., the diagram obtained by reflection in the
main diagonal. Hence the λ0i is the number of squares in the ith column of
λ, or equivalently,
∞
X
λ0i = rk . (4.8)
k=i
In particular, λ01 = l(λ) and λ1 = l(λ0 ). Obviously, λ00 = λ.
We have so far defined the set Pn of partitions of n and known how to
count its size p(n). Now we want to equip a probability measure on this
set. As we will see, this set bears many various natural measures. The first
natural measure is certainly uniform, i.e., choose at random a partition
with equal probability. Let Pu,n be the uniform measure defined by
1
Pu,n (λ) = , λ ∈ Pn , (4.9)
p(n)
where the subscript u stands for Uniform. The primary goal of this chapter
is to study the asymptotic behaviours of a typical partition under (Pn , Pu,n )
as its size n → ∞. The first remarkable feature is that a typical Young
diagram properly scaled has a limit shape. To be precise, define
X
ϕλ (t) = rk , t ≥ 0. (4.10)
k≥t
March 5, 2015 15:59 9197-Random Matrices and Random Partitions ws-book9x6 page 157
In particular,
R∞ ϕλ (i) = λ0i , and ϕλ (t) is a nonincreasing step function such
that 0 ϕλ (t)dt = n.
Theorem 4.2. Under (Pn , Pu,n ) we have as n → ∞
1 √
P
sup √ ϕλ ( nt) − Ψ(t) −→ 0 (4.11)
a≤t≤b n
where 0 < a < b < ∞ and
Z ∞
e−cu 1
Ψ(t) = −cu
du = − log(1 − e−ct ). (4.12)
t 1 − e c
We remark that the curve Ψ(t) was first conjectured by Temperley (1952),
who studied the number of ways in which a given amount of energy can be
shared out among the different possible states of an assembly. The rigorous
argument was given by Vershik (1994, 1996). In fact, Vershik and his school
has been recognized to be the first group who started a systematic study
of limit shapes of various random geometric objects. We also note that the
curve Ψ(t) has two asymptotic lines: s = 0 and t = 0, see Figure 4.1.
To understand the limit (4.14), we remark the following nice fact. Let
η1 , η2 , · · · be a sequence of random variables with (4.14) as their joint
distribution functions, then for x1 > x2 > · · · > xm
P ηm = xm ηm−1 = xm−1 , · · · , η1 = x1 = p(xm−1 , xm ).
Hence the ηk ’s form a Markov chain with p(x, y) as the transition density.
√
Next, let us turn to the bulk case, i.e., treat ϕλ ( nt) where 0 < t < ∞.
Define
1 √ √
Xn (t) = 1/4 ϕλ ( nt) − nΨ(t) , t > 0.
n
An interesting result is the following central limit theorem due to Pittel
(1997).
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 159
Note that ectn /2 /n1/4 goes to zero under the assumption (4.17). We have
so far seen many interesting probability limit theorems for random uniform
partitions. In the next two sections we shall provide rigorous proofs. A
basic strategy is as follows. First, we will in Section 4.2 construct a larger
probability space (P, Qq ) where 0 < q < 1 is a model parameter, under
which the rk ’s are independent geometric random variables. Thus we can
P
directly apply the classical limit theory to the partial sums k rk . Second,
we will in Section 4.3 transfer to the desired space (Pn , Pu,n ) using the fact
Pu,n is essentially the restriction of Qq to Pn . It is there that we develop
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 160
However, we are not able to find a good way to verify such a condition.
Instead, we shall in Section 4.4 state and prove a weaker stochastic
equicontinuity condition: for any ε > 0
lim lim sup Pu,n |Xn (t) − Xn (s)| > ε = 0.
δ→0 n→∞ |t−s|≤δ
Qq (λ ∈ P : rk = j) = (1 − q k )q jk , j = 0, 1, 2, · · · .
∞
Y
Qq (λ) = q krk (1 − q k ),
k=1
as desired.
This lemma will play a fundamental important role in the study of random
uniform partitions. It will enable us to apply the classical limit theorems
for sums of independent random variables. Denote by Eq expectation with
respect to Qq . As a direct consequence, we have
qk qk
Eq rk = , V arq (rk ) =
1 − qk (1 − q k )2
and
1 − qk
Eq z rk = .
1 − zq k
√
Under (P, Qq ), the size |λ| is itself a random variable. Let qn = e−c/ n
,
then it is easy to see
∞
X
µn := Eqn |λ| = kEqn rk
k=1
∞ ∞ √
X kqnk X ke−ck/ n
= k
= √
1 − qn 1 − e −ck/ n
k=1 k=1
Z ∞
ue−cu √
= n −cu
du + O( n)
0 1 − e
√
= n + O( n), (4.20)
Similarly,
∞
X
σn2 := V arqn (|λ|) = k 2 V arqn (rk )
k=1
∞ ∞ √
X k 2 qnk X k 2 e−ck/ n
= = √
(1 − qnk )2 (1 − e−ck/ n )2
k=1 k=1
Z ∞
3/2 u2 e−cu
= n du + O(n)
0 (1 − e−cu )2
2
= n3/2 + O(n). (4.21)
c
Fristedt (1993) obtained the following refinement.
We have
q k (eikx/σn − 1) q k (eikx/σn − 1) 1 qn2k (eikx/σn − 1)2
log 1 − n k
=− n +
1 − qn 1 − qnk 2 (1 − qnk )2
q 3k |eikx/σn − 1|3
+O n
(1 − qnk )3
ix kqnk x2 k 2 qnk
=− · + ·
σn 1 − qnk 2σn2 (1 − qnk )2
k 3 q 3k
n
+O .
(1 − qnk )3
Taking summation over k yields
∞ ∞ ∞
X q k (eikx/σn − 1) ix X kqnk x2 X k 2 qnk
log 1 − n = − +
1 − qnk σn 1 − qnk 2σn2 (1 − qnk )2
k=1 k=1 k=1
∞
1 X k 3 qn3k
+O 3 . (4.24)
σn (1 − qnk )3
k=1
It follows by (4.21),
∞
1 X k 3 qn3k
= O n−1/4 ,
3 k 3
(4.25)
σn (1 − qn )
k=1
1/3
Let ρ(n) = πσn . Then for |x| < ρ(n),
∞
1X 2q k (1 − cos kx/σn )
log |fn (x)| = − log 1 + n
2 (1 − qnk )2
k=1
1 X c1 x2
≤− log 1 + 2/3
2 2/3 2/3 σn
σn /2<k≤σn
X
≤ −c2 x2 σn−2/3
2/3 2/3
σn /2<k≤σn
≤ −c3 x2 .
Thus we get
2
|fn (x)| ≤ e−c3 x , |x| ≤ ρ(n).
n o
2/3
For |x| > ρ(n), let Sx = k : k ≤ σn , cos kx/σn ≤ 0 . Then
1 X 2q k (1 − cos kx/σn )
log |fn (x)| ≤ − log 1 + n
2 (1 − qnk )2
k∈Sx
≤ −c4 σn2/3 ,
which implies
2/3
|fn (x)| ≤ e−c4 σn = o σn−1 .
sup (4.27)
ρ(n)<|x|≤πσn
Next we shall estimate the integral in the right hand side of (4.26).
Split the interval (−πσn , πσn ) into two disjoint subsets: {|x| ≤ ρ(n)} and
{ρ(n) < |x| ≤ πσn }, and evaluate the integral value over each one. Since
2
e−ixn/σn fn (x) → e−x /2 , then by the control convergence theorem,
Z Z ∞ √
2
e−ixn/σn fn (x)dx −→ e−x /2 dx = 2π.
|x|≤ρ(n) −∞
Proof. We first prove the convergence in (4.28) for each fixed t > 0.
Indeed, it follows by (4.20)
1 √ 1 X 1 X qnk
Eqn √ ϕλ ( nt) = √ Eqn rk = √
n n √ n √ 1 − qnk
k≥ nt k≥ nt
Z ∞ −cu
e
= du + o(1) = Ψ(t) + o(1).
t 1 − e−cu
Similarly, by (4.21)
1 √
V arqn √ ϕλ ( nt) = O n−1/2 .
n
Therefore according to the Markov inequality, we immediately have
1 √
P
√ ϕλ ( nt) − Ψ(t) −→ 0. (4.29)
n
Turn to the uniform convergence. Fix 0 < a < b < ∞. For any ε > 0, there
is an m ≥ 0 and a = t0 < t1 < · · · < tm < tm+1 = b such that
√ √
Proof. Let An,x = n(log n/c + x)/c. Since λ1 is the largest part of
λ, then it is easy to see
√
c n
Qqn √ λ1 − log ≤ x = Qqn rk = 0, ∀k ≥ An,x .
n c
It follows by Lemma 4.2
Y Y
1 − qnk .
Qqn (rk = 0) = (4.32)
k≥An,x k≥An,x
→ e−x2 − e−x1 , n → ∞.
Also, according to the proof of Theorem 4.9,
\ −x
lim Qqn {rk = 0} = e−e , x ∈ R.
n→∞
k>An,x
Since two summands in the right hand side of (4.34) are independent and
each converges weakly to a normal random variable, then x1 Xn (s)+x2 Xn (t)
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 168
2
must converges weakly to a normal random variable with variance σs,t given
by
Z t Z ∞
e−cu e−cu
x21 −cu )2
du + (x 1 + x2 )2
du
s (1 − e t (1 − e−cu )2
Z ∞ Z ∞
e−cu 2 e−cu
= x21 du + x 2 du
s (1 − e−cu )2 t (1 − e−cu )2
Z ∞
e−cu
+2x1 x2 du.
t (1 − e−cu )2
Therefore
d
Xn (s), Xn (t) −→ G(s), G(t) ,
where (G(s), G(t)) is jointly normally distributed with covariance
Z ∞
e−cu
Cov(G(s), G(t)) = du.
t (1 − e−cu )2
The m-dimensional case can be analogously proved.
To conclude this section, we investigate the asymptotic behaviours when tn
tends to ∞.
Theorem 4.12. Assume that tn is a sequence of positive numbers such
that
1
tn → ∞, tn − log n → −∞. (4.35)
2c
Then under (P, Qqn ),
ectn /2 √ √ d
1/4
ϕλ ( ntn ) − nΨ(tn ) −→ N (0, 1). (4.36)
n
√
Proof. First, compute mean and variance of ϕλ ( ntn ). According to
the definition of ϕλ ,
√ X X qnk
Eqn ϕλ ( ntn ) = Eqn rk =
√ √ 1 − qnk
k≥ ntn k≥ ntn
Z ∞
√ e−cu
du 1 + O(n−1/2 )
= n −cu
tn 1 − e
√
= nΨ(tn ) 1 + O(n−1/2 )
and
√ X X qnk
V arqn ϕλ ( ntn ) = V arqn (rk ) =
√ √ (1 − qnk )2
k≥ ntn k≥ ntn
Z ∞
√ e−cu
du 1 + O(n−1/2 )
= n −cu 2
tn (1 − e )
√ −ctn −1/2
= ne 1 + O(n ) .
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 169
√
The condition (4.35) guarantees ne−ctn → ∞. Next, we verify the Linde-
berg condition. For any ε > 0,
X X X X
j 2 qnjk (1 − qnk ) ≤ j 2 qnjk
√ √
k≥ ntn j≥εσn k≥ ntn j≥εσn
X X
= j2 qnjk
√
j≥εσn k≥ ntn
√
jd ntn e
X q n
= j2 .
j≥εσn
1 − qnj
This section is devoted to the proofs of main results given in the Intro-
duction. A basic strategy is to use conditioning argument on the event
{|λ| = n}. The following lemma due to Vershik (1996) characterizes the
relations between grand ensembles and small ensembles.
X
≤ Qqn (Wn = wn ) − Qqn Wn = wn |λ| = n
wn ∈Bn
Q |λ| = nW = w
X q n n
= Qqn (Wn = wn ) n − 1. (4.40)
Qqn (|λ| = n)
wn ∈Bn
It follows from (4.38) that the right hand side of (4.40) tends to 0. Thus
we conclude the desired assertion (4.39).
Proof. We will construct Bn such that (i) and (ii) of Lemma 4.4 holds.
First, observe that there is an an such that
X k 2 qnk
= o(a2n ), an = o(n3/4 ). (4.42)
(1 − qnk )2
k∈Kn
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 171
Define
n X X o
Bn = (xk , k ∈ Kn ) : kxk − kEqn rk ≤ an .
k∈Kn k∈Kn
Then by (4.42)
X X
Qqn ◦ Wn−1 Bnc = Qqn
krk − kEqn rk > an
k∈Kn k∈Kn
P
V arqn k∈Kn krk
≤ 2
an
1 X k 2 qnk
= → 0.
a2n (1 − qnk )2
k∈Kn
= σn2 (1 + o(1)) → ∞.
Hence as in Theorem 4.7 one can prove that under (P, Qqn )
P
/ n k(rk − Eqn rk )
qk∈K P −→ N (0, 1)
/ n krk )
V arqn ( k∈K
and so
1 X
k(rk − Eqn rk ) −→ N (0, 1).
σn
k∈K
/ n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 172
Note that
∞
1 X X
n− Eqn rk − k(xk − Eqn rk ) → 0
σn
k=1 k∈Kn
where
n ectn /2 X √ o
B = (xk , k ∈ Kn ) : 1/4 xk − nΨ(tn ) ≤ x .
n k∈K n
∞ ∞
X qnk X qnk
m(k2 ) = , σ 2 (k2 ) =
1 − qnk (1 − qnk )2
k=k2 k=k2
and
2 −1
kX ∞
kqnk X kqnk
s(k1 , k2 ) = , s(k2 ) = .
(1 − qnk )2 (1 − qnk )2
k=k1 k=k2
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 174
The proof of Proposition 4.1 will consist of several lemmas. Note that for
any 0 < q < 1 and z1 , z2
Y Y r Y Y
Eq z1rk z2k = Eq z1rk Eq z2rk
k1 ≤k<k2 k≥k2 k1 ≤k<k2 k≥k2
Y 1 − qk Y 1 − qk
= .
1 − z1 q k 1 − z2 q k
k1 ≤k<k2 k≥k2
where z = (z1 , z2 ).
Note that the above equation (4.49) is still valid for all complex number
q with |q| < 1. Hence using the Cauchy integral formula yields
Z π
Y rk
Y r 1
r−n e−inθ F z; reiθ dθ, (4.50)
Eu,n z1 z2 =
k
2πp(n) −π
k1 ≤k<k2 k≥k2
Lemma 4.6.
k=1
1 1
= F (z; r) exp Re − .
1 − reiθ 1−r
Also, it is easy to see
1 1 2r(1 + r)(cos θ − 1)
Re iθ
− = .
1 − re 1−r (1 − r)3 + 2r(1 − r)(1 − cos θ)
The proof is complete.
and
∞
X ke−tk
s(2) (t, z) = .
(1 − ze−tk )(1 − e−tk )
k=k2
√
−n e2c n −1/4
r F (z; r) = 1 + O(n )
(24n)1/4
u
1 u21 2
· exp 1/4 m(k1 , k2 ) + 1/2 σ (k1 , k2 )
n 2n
u
2 u22 2
· exp 1/4
m(k2 ) + 1/2 σ (k2 )
n c 2n
2
· exp − 2 u1 s(k1 , k2 ) + u2 s(k2 ) .
4n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 176
Proof. Let
2 −1
kX ∞
c2 1 − e−tk X 1 − e−tk
H(z; t) = nt + + log + log .
t 1 − z1 e−tk 1 − z2 e−tk
k=k1 k=k2
= O n1/2 .
(4.53)
∗
To evaluate the third term, note for any t̃ between τ and τ
2
Htt (z; τ ) = n3/2 + O n5/4
c
and
2
τ∗ X
τ∗ − τ = (zb − 1)s(b) (τ ∗ , zb )
2n
b=1
2
c X
ub s(b) + O n−1 .
= 7/4
2n b=1
We have
2 2
c X
Htt z; t̃ (τ ∗ − τ )2 = ub s(b) + O n−1/2 .
2
(4.54)
2n
b=1
Inserting (4.52)-(4.54) into (4.51) yields
√ u1 u1
H(z; τ ) = 2c n + 1/4 m(k1 , k2 ) + 1/2 σ 2 (k1 , k2 )
n 2n
u2 u2
+ 1/4 m(k2 ) + 1/2 σ 2 (k2 )
n 2n
c 2
− 2 u1 s + u2 s(2) + O n−1/4 .
(1)
4n
Finally, with the help of τ − τ ∗ = O n−3/4 , we have
1 τ 1
+ O n−1/4 ,
log = log 1/4
2 2π (24n)
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 178
and so
√
−τ ec n
1 + O(n−1/4 ) .
F(r) = F(e ) = 1/4
(24n)
To conclude the proof, we need only to note
c2
r−n F (z; r) = F(r) exp − + H(z; τ ) .
τ
To estimate the integral over (−π, π), we split the interval (−π, π) into two
subsets: |θ| ≤ θn and |θ| ≥ θn , where θn = n−3/4 log n. The following
lemma shows that the overall contribution to the value of integral in (4.50)
made by large θ’s is negligible.
1/4 1/4
Proof of Proposition 4.1. Let z1 = eu1 /n and z2 = eu2 /n , and
choose r = e−τ in (4.50). Combining (4.55) and (4.56) yields
Z π Z Z
r−n e−inθ F z; reiθ dθ = r−n e−inθ F z; reiθ dθ
+
−π |θ|≤θn |θ|≥θn
−n π
=r F (z, r) 1 + o(1) .
61/4 n3/4
Taking Theorem 4.1 and Lemma 4.7 into account, we conclude (4.48) as de-
sired.
√
Proof of Theorem 4.5. Take u1 = x1 and u2 = x1 + x2 and k1 = n t1
√
and k2 = n t2 . Note
Z t2
√ e−cu
σ 2 (k1 , k2 ) = n −cu )2
du(1 + o(1)),
t1 (1 − e
∞
√ e−cu
Z
2
σ (k2 ) = n du(1 + o(1)),
t2 (1 − e−cu )2
t2
ue−cu
Z
s(k1 , k2 ) = n du(1 + o(1)),
t1 (1 − e−cu )2
∞
ue−cu
Z
s(k2 ) = n du(1 + o(1)).
t2 (1 − e−cu )2
Substituting these into (4.48) of Proposition 4.1, we easily get (4.47), as
required. We conclude the proof.
In this section we shall first prove a theorem that may allow us to get the
distributional results in the case when the functional of λ depends primarily
on the moderate-sized parts. Then we shall use the functional central limit
theorem to prove the asymptotic normality for character ratios and the
log-normality for dλ .
Introduce an integer-valued function
l √n 1 m
kn (t) = log , t ∈ (0, 1).
c 1−t
Let
c √
t0 (n) = √ , t1 (n) = n−δ0 , t2 (n) = 1 − e−c/ n
2 n
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 181
We shall prove the theorem following the line of Pittel (2002) by applying
the Gikhman-Skorohod theorem, namely Theorem 1.16. A main step is to
verify the stochastic equicontinuity for Yn (t), 0 ≤ t ≤ 1. As the reader may
notice, a significant difference between Yn and Xn is that there is an extra
factor t in Yn besides parametrization. This factor is added to guarantee
that Yn satisfies the stochastic equicontinuity property and so the limit
process Y has continuous sample paths.
where the error term holds uniformly over nδ ≤ k1 < k2 ≤ ∞ with δ < 1/2.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 182
Y 1 − qnk
=
k1 ≤k<k2
1 − eu/n1/4 qnk
1/4
X 1 − eu/n qnk
= exp − log .
1 − qnk
k1 ≤k<k2
1 1/4 2 X qn2k
+ eu/n −1
2 (1 − qnk )2
k1 ≤k<k2
1/4 X qn3k
+O |eu/n − 1|3
.
(1 − qnk )3
k1 ≤k<k2
Note
u/n1/4 3 X qn3k |u|3 Z ∞ e−3cx
e − 1 k 3
= O 1/4 √ −cx 2
dx
(1 − qn ) n k1 / n (1 − e )
k1 ≤k<k2
= O n3/4−2δ ,
and the contribution proportional to u3 that comes from the first two terms
is of lesser order of magnitude. We conclude the proof.
where the error term holds uniformly over nδ ≤ k1 < k2 ≤ ∞ with δ > 3/8.
Indeed, the above equation holds for any complex number |q| < 1. Using
the Cauchy integral formula, we obtain
Z π
P
k1 ≤k<k2 rk
1
r−n e−inθ Fn x; reiθ dθ
Eu,n x =
2πp(n) −π
where
Y 1 − qk
Fn (x; q) := F(q) .
1 − xq k
k1 ≤k<k2
√
−c/ n
Choose r = qn = e , with the help of (4.5) we further obtain
P
rk
Y 1 − qnk
Eu,n x k1 ≤k<k2
≤ c8 .
1 − xqnk
k1 ≤k<k2
√
u/ n
Letting x = e , we have by (4.62)
u X u X
Eu,n exp 1/4 rk ≤ c8 exp 1/4 Eqn rk
n k1 ≤k<k
n k1 ≤k<k2
2
u2 X qnk
3/4−2δ
· exp + O n ,
2n1/2 (1 − qnk )2
k1 ≤k<k2
Proof. Observe that if t > t2 (n) then kn (t) > n so Yn (t) = 0; while
if t < t0 (n) then Yn (t) = Op (n1/4−δ0 log n). So it suffices to prove (4.64)
uniformly for t, s ∈ [t1 (n), t2 (n)]. Assume s < t. We use Lemma 4.11 to
obtain
us X
Eu,n exp 1/4 (rk − Eqn rk )
n
kn (s)≤k<kn (t)
u2 s2 X qnk
≤ c9 exp k 2
2n1/2 (1 − qn )
kn (s)≤k<kn (t)
u2
≤ c10 exp (t − s) (4.65)
2c
and
u(t − s) X
Eu,n exp (rk − Eqn rk )
n1/4
k≥kn (t)
u2 (t − s)2 X qnk
≤ c11 exp
2n1/2 (1 − qnk )2
k≥kn (t)
u2
≤ c12 exp (t − s) . (4.66)
2c
Note
t−s X
Yn (t) − Yn (s) = (rk − Eqn rk )
n1/4
k≥kn (t)
s X
− (rk − Eqn rk ).
n1/4
kn (s)≤k<kn (t)
Therefore we have
2 ρ
Pu,n |Yn (t) − Yn (s)| > ε ≤ e−cε /8δ + e−cn ε/8
uniformly for s, t ∈ [t1 (n), t2 (n)] with |t−s| ≤ δ. This verifies the stochastic
equicontinuity property (4.64).
We can analogously obtain
( 2
e−cx /8(t−s) , x ≤ nρ t(1 − t)/c,
Pu,n |Yn (t)| > x ≤ ρ
e−cn x/8 , x ≥ nρ t(1 − t)/c.
Therefore it follows by integral formula by parts
Z ∞
Eu,n |Yn (t)|m = m xm−1 Pu,n |Yn (t)| > x dx
0
≤ c15 tm/2 (1 − t)m/2 + n−mρ ,
as desired.
Proof of Theorem 4.13. Begin with the continuity of sample paths of
Y (t). Note
2
E Y (t) − Y (s) = EY (t)2 − 2EY (s)Y (t) + EY (s)2
1 1
= t − s − (t − s)2 − (l(t) − l(s))2
c 2
1
≤ (t − s).
c
Since Y (t) − Y (s) is Gaussian with zero mean, we have
4 3
E Y (t) − Y (s) ≤ 2 (t − s)2 .
c
This implies by Kolmogorov’s continuity criterion that there exists a sepa-
rable continuous version of Y (·) on [0, 1].
Turn to the proof of (ii). The asymptotic normality directly follows
from Theorem √ 4.5. Indeed, making a change of time parameter, we see
that tXn( cn log 1−t 1
), 0 < t < 1 converges weakly to tX( 1c log 1−t 1
),
0 < t < 1 in terms of finite dimensional distributions. If letting Y (t) =
tX( 1c log 1−t
1
), then a simple calculus shows that Y (t), 0 < t < 1 has the
desired covariance structure (4.60).
Finally, we show (iii). Fix g. Without loss of generality, we assume
that α, β > 1. Introduce εm = 1/m, m ≥ 1, and break up the integration
interval into three subsets: (0, εm ), (εm , 1 − εm ), and (1 − εm , 1). Let
Z 1 Z 1−εm
Zn = g(t, Yn (t))dt, Zn,m = g(t, Yn (t))dt
0 εm
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 186
and
Z 1 Z 1−εm
Z= g(t, Y (t))dt, Zm = g(t, Y (t))dt.
0 εm
Then using Lemma 4.12 and Theorem 1.16, it is not difficult to check the
following three statements:
(a) for any ε > 0,
lim lim Pn (|Zn − Zn,m | > ε) = 0;
m→∞ n→∞
(b) for each m ≥ 1,
d
Zn,m −→ Zm ;
(c) for any ε > 0,
lim P (|Zm − Z| > ε) = 0, m → ∞.
m→∞
Here we prefer to leave the detailed computation to the reader, see also
Pittel (2002).
Having (a), (b) and (c) above, Theorem 4.2, Chapter 1 of Billingsley
d
(1999a) guarantee Zn −→ Z, which concludes the proof.
To illustrate, we shall give two examples. The first one treats the char-
acter ratios in the symmetric group Sn . Fix a transposition τ ∈ Sn . Define
the character ratio by
χλ (τ )
γτ (λ) = , λ ∈ Pn
dλ
where χλ be an irreducible representation associated with the partition
λ 7→ n, dλ is the dimension of χλ , i.e., dλ = χλ (1n ).
The ratio function played an important role in the well-known analy-
sis of the card-shuffling problem performed by Diaconis and Shahshahani
(1981). In fact, Diaconis and Shahshahani proved that the eigenvalues for
this random walk are the character ratios each occurring with multiplicity
d2λ . Character ratios also play a crucial role in the work on moduli spaces
of curves, see Eskin and Okounkov (2001), Okounkov and Pandharipande
(2004). The following theorem can be found in the end of the paper of
Diaconis and Shahshahani (1981).
Theorem 4.14. Under (Pn , Pu,n ),
d
n3/4 γτ (λ) −→ N 0, στ2 ,
Note that the limit variable is a centered Gaussian random variable with
variance
Z 1Z 1
EY (s)Y (t) 1−s 1−t
log log dsdt,
0 0 s(1 − s)t(1 − t) s t
where EY (s)Y (t) is given by (4.60).
In combination, we obtain
d
n3/4 γτ (λ) −→ N 0, στ2 ,
as desired.
The second example we shall treat is dλ . It turns out that the logarithm
of dλ satisfies the central limit theorem. Introduce
Z ∞
log | log x|
κ(t) = dx. (4.67)
0 (1 − t − tx)2
Theorem 4.15. Under (Pn , Pu,n ),
1 1
d
−→ N 0, σd2 .
3/4
log d λ − n log n + An
n 2
Here A and σd2 are given by
1 ∞ y log y
Z
A = 1 − log c + 2 dy
c 0 ey − 1
and
Z 1 Z 1
1
σd2 = EY (s)Y (t)κ(s)κ(t)dsdt, (4.68)
c2 0 0
The theorem was first proved by Pittel (2002). The proof will use the
following classic identities (see also Chapter 5):
Q
1≤i<j≤l (λi − λj + j − i)
dλ = n! Q (4.69)
1≤i≤l (λi − i + l)!
and
n! n!
dλ = := Q , (4.70)
Hλ ∈λ h
Consequently, we obtain
log dλ − log n!
X X
log λ0i − λ0j + j − i − log λ0i − i + λ1 !
=
1≤i<j≤λ1 1≤i≤λ1
=: Mn − Nn . (4.71)
The bulk of the argument consists of computing Mn and Nn . We need
some basic estimates about λ0i − λ0j below. Let 0 < δ < 1/2 and define
K = (k1 , k2 ) : nδ ≤ k1 ≤ k2 − nδ .
Lemma 4.14. Given ε > 0 and 0 < a < δ < 1/2, there is an n0 (a, δ, ε) ≥ 1
such that
(i)
p
Qqn |λ0k1 − λ0k2 − m(k1 , k2 )| > σ(k1 , k2 ) ε log n ≤ n−ε/3
√
uniformly for (k1 , k2 ) ∈ K and k1 ≤ a n log n/c;
(ii)
Qqn |λ0k1 − λ0k2 − m(k1 , k2 )| > (σ(k1 , k2 ) + log n) ε log n ≤ n−ε/3
p
√
uniformly for (k1 , k2 ) ∈ K and k1 ≥ a n log n/c.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 190
Proof. Let η be small enough to ensure that eη qnk < 1 for all k ∈ [k1 , k2 ).
√
For instance, select |η| = σ(k1 , k2 ) ε log n. We have
η2
Eqn exp η(λ0k1 − λ0k2 ) = exp ηm(k1 , k2 ) + σ 2 (k1 , k2 )
2
3
X qn3k
· exp O |η| 3k
. (4.72)
(1 − qn )
k1 ≤k<k2
Note
X qn2k
η2 = O(n−a log n).
(1 − qn )2k
k1 ≤k<k2
Using (4.73) and the Markov inequality, and noting σ 2 (k1 , k2 )/m(k1 , k2 ) =
1 + o(1), we have
p
Qqn |λ0k1 − λ0k2 − m(k1 , k2 )| ≥ (σ(k1 , k2 ) + log n) ε log n ≤ n−ε/3 .
X
λ0k − m(k) log m(k) − k + λ1 ,
Rn =
ln ≤k<kn
δ
√
where ln = [n ] and kn = n log n/c .
Lemma 4.15. Under (Pn , Pu,n ), we have for 1/8 < δ < 1/4
Nn = Nn + Rn + op (n3/4 ). (4.75)
Proof. Let φ(x) = x log x − x. Then by the Stirling formula for factorial,
λ1
X
φ λ0k − k + λ1 + ∆(λ1 ),
Nn = (4.76)
k=1
λ1
X
N̄n = φ m(k) − k + λ1 .
k=1
We shall compare Nn with N̄n below. For this, we break up the sum in
(1) (2) (3)
(4.76) into Nn , Nn and Nn for k ∈ [1, ln ), k ∈ [ln , kn ] and k ∈ (kn , ∞)
(1) (2) (3)
respectively. We similarly define N̄n , N̄n and N̄n .
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 192
and so
φ λ0k − k + λ1 = Op n1/2 log2 n , φ m(k) − k + λ1 = Op n1/2 log2 n .
Then we have
Nn(1) = Op n1/2+δ log2 n , N̄n(1) = Op n1/2+δ log2 n ,
which implies
Nn(1) − N̄n(1) = Op n1/2+δ log2 n = op n3/4 .
(4.77)
(2) (2)
Turn to Nn − N̄n . With the help of (4.74), we expand φ(x) at x =
m(k) − k + λ1 :
φ λ0k − k + λ1 = φ m(k) − k + λ1 + λ0k − m(k) log m(k) − k + λ1
σ 2 (k) log n
+Op .
m(k) − k + λ1
It follows from Lemma 4.13 that the remainder term is controlled by
Op (log n). Hence
kn
X
Nn(2) N̄n(2) λ0k − m(k) log m(k) − k + λ1
− =
k=ln
k>kn
where x∗k
is between m(k) − k + λ1 and λ0k − k + λ1 .
It follows from (4.74)
Nn(3) − N̄n(3) = op n3/4 .
(4.79)
Putting (4.77)-(4.79) together yields
Nn − N̄n = Rn + op n3/4 .
X
Mn = log µ(i, j) + j − i ,
1≤i<j≤λ1
X
λ0k − m(k) v(yk ) + log(g(yλ1 ) − g(yk )) ,
Sn =
2ln ≤k≤λ1 −ln
√ Pj−1
where yk = ck/ n and µ(i, j) = l=i µl .
where
2ln λ1
X X 1
Sn(1) = λ0k − m(k)
m(j, k) + j − k
k=ln j=k+ln
and
λ1
X X 1
Sn(2) = λ0k − m(k)
.
m(j, k) + j − k
k=2ln ln ≤j≤λ1 ,|j−k|≥ln
(1)
It follows from Proposition 4.2 that with high probability Sn is of smaller
(2)
order than n3/4 . To study Sn , we need a delicate approximation (see pp.
200-202 of Pittel (2002) for lengthy and laborious computation):
X 1
= v(yk ) + log g(yλ1 ) − g(yk )
m(j, k) + j − k
ln ≤j≤λ1 ,|j−k|≥ln
and
X
log m(k1 , k2 ) + k2 − k1 − log µ(k1 , k2 ) + k2 − k1
1≤k1 <k2 ≤λ1
X 1
= Op n1/2 log2 n .
≤2
k2 − k1
1≤k1 <k2 ≤λ1
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 195
Therefore we need only prove that the right hand side of (4.88) are neg-
ligible. First, according to Lemma 5 of Pittel (2002), there is a constant
c6 > 0
1 c6
|v(x)| = c6 log x + , |v 0 (x)| ≤ . (4.89)
x x
We easily get
X
v(yk )λ0k = O nδ λ1 log n = Op n1/2+δ log2 n .
k>λ1 −ln
Thus we have
X
v(yk ) λ0k − m(k) = op n3/4 .
k>λ1 −ln
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 196
= n + Op n1/2 log n .
k∈[2l
/ n ,kn ]
which together with (4.90) in turn implies
X
λ0k − m(k) = Op n3/4 .
(4.91)
2ln ≤k≤kn
Besides, we have by (4.74)
X
λ0k − m(k) = Op n3/4 log n .
(4.92)
2ln ≤k≤kn
By the definition of g(x) and m(k),
√ √ √
n n
1 − e−cλ1 / n
g(yλ1 ) − g(yk ) = m(k) − k + λ1 +
c c
= m(k) − k + λ1 + Op (1).
Therefore for k ≤ kn
log g(yλ1 ) − g(yk ) − log m(k) − k + λ1
c
= log √ + Op n−1/2 log−1 n .
n
Thus by (4.91) and (4.92)
X
log(g(yλ1 ) − g(yk )) − log(m(k) − k + λ1 ) λ0k − m(k)
2ln ≤k≤kn
c X
λ0k − m(k)
= log √
n
2ln ≤k≤kn
X
−1/2
log−1 n
0
+Op n λk − m(k)
2ln ≤k≤kn
3/4
= op n .
The proof is complete.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 197
Moreover,
X
− log ν(i) − k + µ(k) − i + 1
(i,k)∈D
X
=− log `(i) − k + `(k) − i + 1
(i,k)∈D
X |`(i) − ν(i)| |`(k) − µ(k)|
+O +
min{ν(i), `(i)} − k + 1 min{µ(k), `(k)} − i + 1
(i,k)∈D
X
log `(i) − k + `(k) − i + 1 + Op n1/2 log2 n .
=−
(i,k)∈D
By virtue of Proposition 4.2, the whole order is actually Op (n3/4−δ log1/2 n).
Neglecting ∆Tn , we can equate Tn with the integral on the right side in
(4.95). Furthermore, we also extend the integration to [xn , ∞), where
√
n c
xn = − log 1 − √ .
c 2 n
Making the substitution
√
n 1
x= log
c 1−t
in the last integral and using the definition of the process Yn (t) we obtain
n3/4 1 v(− log(1 − t))
Z
Yn (t)dt + O n3/4−ε .
Tn =
c 0 t(1 − t)
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 200
converges in D% .
Now define the measure Pm,n on Pn by
sk (j)
Pm,n λ ∈ Pn : rk (λ) = j =
Zm,n
and
Q∞
k=1 sk (rk )
Pm,n (λ) = , λ ∈ Pn ,
Zm,n
where
∞
X Y
Zm,n = sk (rk ).
λ∈Pn k=1
Vershik (1996), Vershik and Yakubovich (2006) treat Qβ,q and Pβ,n as
generalized Bose-Einstein models of ideal gas; while in combinatorics and
number theory they are well known for a long time as weighted partitions.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 202
Having the limit shape, we will continue to further study the second order
fluctuation of Young diagrams around it. This will separately discussed
according to two cases: at the edge and in the bulk. First, let us look at
the asymptotic distribution of the largest part of a partition under Pβ,n .
The following result, due to Vershik and Yakubovich (2006), is an extension
of Erdös and Lehner’s theorem
Theorem 4.17.
An x −x
lim Pβ,n λ ∈ Pn : λ1 − ≤ = e−e ,
n→∞ hn hn
where
β+1 β+1 β+1
An = log n + β log log n + β log − log Γ(β + 2)ζ(β + 2).
β+2 β+2 β+2
d
Remark 4.3. As known to us, λ = λ0 , and so λ01 and λ1 have the same
asymptotic distribution under (Pn , Pu,n ). But such an elegant property is
no longer valid under (Pn , Pβ,n ). In fact, we have the following asymptotic
normality for λ01 instead of Gumbel distribution.
−(β+1)
Let σn2 = hn and define for k ≥ 1
∞ ∞
X e−hn j X e−hn j
µn,k = jβ , 2
σn,k = jβ ,
1 − e−hn j (1 − e−hn j )2
j=k j=k
where Xβ (t) is a normal random variable with zero mean and variance
1 2
2
κ2β (t) = σβ2 (t) − σβ+1 (t) ;
Γ(β + 3)ζ(β + 2)
(ii) for 0 < t1 < t2 < · · · < tm < ∞,
d
Xβ,n (t1 ), Xβ,n (t2 ), · · · , Xβ,n (tm ) −→ Xβ (t1 ), Xβ (t2 ), · · · , Xβ (tm ) ,
where Xβ (t1 ), Xβ (t2 ), · · · , Xβ (tm ) is a Gaussian vector with covariance
structure
2 2
σβ+1 (s)σβ+1 (t)
Cov Xβ (s), Xβ (t) = σβ2 (t) −
, s < t;
Γ(β + 3)ζ(β + 2)
(iii) Each separable version of Xβ is continuous in (0, ∞).
and
Z ∞ Z ∞
2
σβ,d = Cov Xβ (s), Xβ (t) log Ψβ (s) log Ψβ (t)dsdt.
0 0
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 205
1 Y ark
k
Pa,n (λ) = , λ ∈ Pn ,
Za,n rk !
k=1
Chapter 5
5.1 Introduction
207
March 5, 2015 15:59 9197-Random Matrices and Random Partitions ws-book9x6 page 208
Lemma 5.1.
√ √
Pp,n max{λ1 , λ01 } ≥ 2e n ≤ e−2e n . (5.6)
Lemma 5.2. As n → ∞
d2 √
Pp,n − log λ > 2c n → 0, (5.7)
n!
√
where c = π/ 6 as in Chapter 4.
√ √
For simplicity of notations, write ψn,λ (x) for ψλ ( nx)/ n. Then
1
log √ λi − i + λ0j − j + 1
n
Z Z
1
√ ψλ (x) − x + ψλ−1 (y) − y dxdy
= log
n
Z Z
1
log √ ψλ (x) − x + ψλ−1 (y) − y dxdy,
≥
n
−1
where stands for the (i, j)th unit square, ψn,λ denotes the inverse of ψn,λ
and the last inequality follows from the concavity property of logarithmic
function.
Hence we obtain
d2
− log λ ≥ − log n! + n log n
n! Z Z
−1
+2n log ψn,λ (x) − x + ψn,λ (y) − y dxdy
0≤y<ψn,λ (x)
=: nI(ψn,λ ) + n ,
where n = O(log n) and
Z Z
−1
I(ψn,λ ) = 1 + 2 log ψn,λ (x) − x + ψn,λ (y) − y dxdy.
0≤y<ψn,λ (x)
(i)
J(Ω) = 0; (5.10)
(ii)
Z ∞ Z ∞
1
J(Ψn,λ ) = − log |u − v|∆0n,λ (u)∆0n,λ (v)dudv
4 −∞ −∞
|u|
Z
+ Ψn,λ (u) − |u| arccosh du. (5.11)
|u|>2 2
Consequently,
Z ∞ Z ∞
1
J(Ψn,λ ) ≥ − log |u − v|∆0n,λ (u)∆0n,λ (v)dudv. (5.12)
4 −∞ −∞
Also, it follows
Z u Z 2
%0 (u − v)dv = %1 (2 + u), %0 (u − v)du = %1 (2 − v).
−2 v
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 214
To calculate the last three double integral in the right hand side of (5.13),
set
Z 2
H2 (u) = %1 (u − v)Ω00 (v)dv, −∞ < u < ∞.
−2
Then
Z 2
2 1
H200 (u) = √ dv = 0, −2 ≤ u ≤ 2
π −2 (u − v) 4 − v 2
and so for each −2 ≤ u ≤ 2
2
log |v|
Z
2
H20 (u) = H20 (0) = √ dv = 0.
π −2 4 − v2
This in turn implies
Z 2
H2 (u) = H2 (0) = − %1 (v)Ω00 (v)dv = 0 (5.14)
−2
Similarly,
Z 2 Z u Z 2
%0 (u − v)dv Ω0 (u)du = %1 (2 + u)Ω0 (u)du
−2 −2 −2
= %2 (4) − 2.
Again, by (5.14)
Z 2 Z 2
%0 (u − v)Ω0 (v)dv = %1 (2 − u) − %1 (2 + u) + %1 (u − v)Ω00 (v)dv
−2 −2
= %1 (2 − u) − %1 (2 + u).
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 215
Hence we have
Z 2Z 2 Z 2
%0 (u − v)Ω0 (u)Ω0 (v)dudv = %1 (2 − u) − %1 (2 + u) Ω0 (u)du
−2 −2 −2
= 4 − 2%2 (4).
In combination, we have proven (5.10).
Turn to the proof of (5.11). First, observe there is an a = an (may
depend on λ) such that [−a, a] contains the support of ∆n,λ (u). Hence we
have
1 a a
Z Z
J(Ψn,λ ) = 1 − %0 (u − v)dudv
4 −a −a
1 a u
Z Z
+ %0 (u − v)dv Ψ0n,λ (u)du
2 −a −a
1 a a
Z Z
− %0 (u − v)du Ψ0n,λ (v)dv
2 −a v
1 a a
Z Z
+ %0 (u − v)Ψ0n,λ (u)Ψ0n,λ (v)dudv.
4 −a −a
A simple calculus shows
Z aZ a
%0 (u − v)dudv = 2%2 (2a),
−a −a
Z a Z u Z a
%0 (u − v)dv Ψ0n,λ (u)du = %1 (a + u)Ψ0n,λ (u)du
−a −a −a
= ρ2 (2a),
Z a Z a Z a
%0 (u − v)du Ψ0n,λ (v)dv = %1 (a − v)Ψ0n,λ (v)dv
−a v −a
= −ρ2 (2a).
On the other hand, it follows
1 ∞ ∞
Z Z
− log |u − v|∆0n,λ (u)∆0n,λ (v)dudv
4 −∞ −∞
1 a a
Z Z
= %0 (u − v)Ω0 (u)Ω0 (v)dudv
4 −a −a
1 a a
Z Z
− %0 (u − v)Ω0 (u)Ψ0n,λ (v)dudv
2 −a −a
1 a a
Z Z
+ %0 (u − v)Ψ0n,λ (u)Ψ0n,λ (v)dudv.
4 −a −a
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 216
and similarly
Z a Z −2
%1 (a + v)Ω0 (v)dv = %2 (2a) − 2 + H2 (u)du.
−a −a
By (5.15),
Z a Z a
%0 (u − v)Ω0 (u)du Ψ0n,λ (v)dv
−a −a
Z a Z −2
= −2ρ2 (2a) + H2 (v)Ψ0n,λ (v)dv + H2 (v)Ψ0n,λ (v)dv.
2 −a
In combination, we get
J(Ψn,λ )
1 ∞ ∞
Z Z
=− log |u − v|∆0n,λ (u)∆0n,λ (v)dudv
4 −∞ −∞
1 a 1 −2
Z Z
0
+ H2 (u) Ψn,λ (u) − u du + H2 (u)(Ψn,λ (u) + u)0 dv.
2 2 2 −a
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 217
Lemma 5.4.
Z ∞ Z ∞
1
− log |u − v|f 0 (u)f 0 (v)dudv = kf k2s . (5.16)
−∞ −∞ 2
Proof. Denote by H(f ) the Hilbert transform, namely
Z ∞
f (u)
H(f )(v) = du.
−∞ v −u
Then it is easy to see
Z ∞
[)(ω) =
H(f ei2πωv H(f )(v)dv
−∞
= isgnω fb(ω),
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 218
To finish the proof, we need a key observation due to Vershik and Kerov
(1985)
Z ∞
2 1
|ω|fb(ω) dω = kf k2s .
−∞ 2
The proof is complete.
Combining (5.12) and (5.16) yields
1
J(Ψn,λ ) ≥ k∆n,λ k2s . (5.17)
8
Now we are ready to give
Proof of Theorem 5.2. In view of Lemma 5.5, we can and do consider
only the case in which the support of ∆n,λ is contained in a finite interval,
say, [−a, a]. We will divide the double integral into two parts:
Z aZ a Z a 2
2 ∆n,λ (u) − ∆n,λ (v) 2 ∆n,λ (u)
k∆n,λ ks = dudv + 4a 2 2
du,
−a −a u − v −a a − u
for two distinct points x1 and x2 . Thus, one cannot expect a kind of
weak convergence of the stochastic process. However, we will in Section
5.2 establish a functional central limit theorem, namely Kerov’s integrated
√ √
central limit theorem for ψλ ( nx) − nω(x).
Note
sin(k + 1)θ
uk (2 cos θ) =
sin θ
and
Z 2
uk (u)ul (u)ρsc(u) du = δk,l .
−2
This theorem is referred to the Kerov CLT since it was Kerov who first
presented it and outlined the main ideas of the proof in Kerov (1993). A
complete and rigorous proof was not given by Ivanov and Olshanski (2002)
until 2002. The proof uses essentially the moment method and involves a
lot of combinatorial and algebraic techniques, though the theorem is stated
in standard probability terminologies. We need to introduce some basic
notations and lemmas.
Begin with Frobenius coordinates. Let λ = (λ1 , λ2 , · · · , λl ) be a parti-
tion from P. Define
and
1 1
ai = λi − i + , bi = λ0i − i + , i = 1, 2, · · · , `, (5.20)
2 2
where ` := `(λ) is the length of the main diagonal in the Young dia-
gram of λ. The natural numbers {āi , b̄i , i = 1, · · · , `} are called the usual
Frobenius coordinates, while the half integer numbers {ai , bi , i = 1, · · · , `}
are called the modified Frobenius coordinates. We sometimes represent
λ = (a1 , a2 , · · · , a` |b1 , b2 , · · · , b` ).
Proof. First, observe the infinite series in (5.21) is actually finite because
λi = 0 when i is large enough. Second, the second product is an noncon-
tractible function since the numbers a1 , a2 , · · · , a` , −b1 , −b2 , · · · , −b` are
pairwise distinct. The identity (5.21) follows from a classical Frobenius
lemma, see Proposition 1.4 of Ivanov and Olshanski (2002).
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 222
Lemma 5.9. Fix k ≥ 1 and λ ∈ P. Then p]k (λ) equals the coefficient of
z −1 in the expansion of the function
1 1 ↓k Φ(z; λ)
− z− (5.33)
k 2 Φ(z − k; λ)
in descending powers of z about the point z = ∞.
mation yields
Φ z − n + 21 ; λ
1 ↓k
F (z) = − (z − n) .
k Φ z − n + 12 − k; λ
Note the residue at z = ∞ will not change under the shift z 7→ z + n − 1/2.
Consequently,
1 1 ↓k Φ(z; λ)
p]k (λ) = −Res − z− ,z = ∞ .
k 2 Φ(z − k; λ)
The proof is complete.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 226
We shall employ the following notation. Given a formal series A(t), let
[tk ]{A(t)} = the coefficient of tk in A(t).
Lemma 5.10. The functions p]k (λ) belong to the algebra A. In particular,
it can be described through the generators p̄1 , p̄2 , · · · as follows
n 1Y k
1
p]k (λ) = [tk+1 ] − 1 − (j − )t
k j=1 2
∞
X p̄j (λ)tj o
· exp 1 − (1 − kt)−j . (5.34)
j=1
j
The proof is left to the reader. See also Proposition 3.7 of Ivanov and
Olshanski (2002), which contains a general inversion formula.
We next extend p]k in (5.32) to any partition ρ on P. Let |ρ| = r, define
( n−r
] n↓r χλ (ρ,1
dλ
)
, λ ∈ Pn , n ≥ r,
pρ (λ) = (5.37)
0, λ ∈ Pn , n < r.
The following lemma lists some basic properties. The reader is referred to
its proof and more details in Kerov and Olshanski (1993), Okounkov and
Olshanski (1998), Vershik and Kerov (1985).
Lemma 5.12. (i) For any partition ρ, the function p]ρ is an element of A.
(ii) In the canonical grading,
p]ρ (λ) = p̄ρ (λ) + lower terms,
where λ = (1r1 , 2r2 , · · · ) and
Y
p̄ρ (λ) = p̄i (λ)ri .
i=1
We remark that the basis p]ρ is inhomogeneous both in the canonical grading
and weight grading. For each f ∈ A, let (f )ρ be the structure constants of
f in the basis of p]ρ . Namely,
X
f (λ) = (f )ρ p]ρ (λ).
ρ
Proof. This directly follows from the definition (5.37). Indeed, we need
only to show for each λ ∈ Pn with n ≥ |σ| since otherwise both sides are
equal to 0. When n = |σ| and λ ∈ Pn ,
p]σ (λ)p]1 (λ) = n · p]σ (λ) = |σ| · p]σ (λ).
Next assume n ≥ |σ| + 1. Setting k = |σ|, then
χλ (σ, 1n−k )
p]σ (λ)p]1 (λ) = n↓(k) · n .
dλ
Hence the claim follows from a simple relation
n↓k · n = n↓(k+1) + n↓k · k.
Proof. Set
X
p]σ p]τ = (p]σ p]τ )ρ p]ρ .
ρ
We claim that only partitions ρ with kρkN ≤ kσkN + kτ kN can really con-
tribute. Indeed, assume (p]σ p]τ )ρ 6= 0, and fix a set X of cardinality |ρ| and a
permutation s : X → X whose cycle structure is given by ρ. Then according
to Proposition 4.5 of Ivanov and Olshanski (2002) (see also Proposition 6.2
and Theorem 9.1 of Ivanov and Kerov (2001)), there must exist a quadruple
{X1 , s1 , X2 , s2 } such that
(i) X1 ⊆ X, X2 ⊆ X, X1 ∪ X2 = X;
(ii) |X1 | = |σ| and x1 : X1 7→ X1 is a permutation of cycle structure σ;
(iii) |X2 | = |τ | and x2 : X2 7→ X2 is a permutation of cycle structure τ ;
(iv) denoting by s̄1 : X → X and s̄2 : X → X the natural extensions of s1,2
from X1,2 to the whole X. I.e., s̄1,2 is trivial on X \ X1,2 , then s̄1 s̄2 = s.
Fix any such quadruple and decompose each of the permutations s, s1 , s2
into cycles. Let CN (s1 ) denote the set of all cycles of s1 , AN (s1 ) the subset
of those cycles of s1 that entirely contained in X1 \ X2 , BN (s1 ) the subset
of those cycles of s1 that have a nonempty intersection with X1 ∩ X2 . Then
CN (s1 ) = AN (s1 ) + BN (s1 ).
Define similarly CN (s2 ), AN (s2 ) and BN (s2 ), then we have
CN (s2 ) = AN (s2 ) + BN (s2 ).
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 229
Similarly again, let CN (s) denote the set of all cycles of s, BN (s) the subset
of those cycles of s that intersect both X1 and X2 . Then
CN (s) = AN (s1 ) + AN (s2 ) + BN (s).
The claimed inequality kρkN ≤ kσkN + kτ kN is equivalent to
|X| + |BN (s)| ≤ |X1 | + |BN (s1 )| + |X2 | + |BN (s2 )|. (5.41)
To prove (5.41), it suffices to establish a stronger inequality:
|BN (s)| ≤ |X1 ∩ X2 |. (5.42)
To see (5.42), it suffices to show each cycle ∈ o ∈ BN (s) contains a point
of X1 ∩ X2 . By the definition of BN (s), cycle o contains both points of X1
and X2 . Therefore there exist points x1 ∈ X1 ∩ c and x2 ∈ X2 ∩ o such that
sx1 = x2 . By (iv), it follows that either x1 or x2 lies in X1 ∩ X2 . Thus the
claim is true, as desired.
Now assume (p]σ p]τ )ρ 6= 0 and kρkN = kσkN + kτ kN , then both BN (s1 )
and BN (s2 ) are empty, which implies X1 ∩ X2 = ∅. Therefore ρ = σ ∪ τ .
Finally, by (5.38),
(p]σ p]τ )σ∪τ = 1.
It concludes the proof.
Lemma 5.15. (i) For any two partitions σ and τ with no common part,
p]σ p]τ = p]σ∪τ + lower terms,
where lower terms means terms with deg1 (·) < kσ ∪ τ k1 .
(ii) For any partition σ ∈ P and k ≥ 2, if rk (σ) ≥ 1, then
p]σ p]k = p]σ∪k + krk (σ)p](σ\k)∪1k + lower terms, (5.43)
where lower terms means terms with deg1 (·) < kσk1 + k.
Proof. (i) can be proved in a way similar to that of Lemma 5.14 with
minor modification. Turn to (ii). Set
X
p]σ p]τ = (p]σ p]τ )ρ p]ρ .
ρ
n↓r X
= χλ (ρ, 1n−r )χλ (e)
n!
λ∈Pn
↓r
n , ρ = (1r ),
=
0, otherwise.
The proof is complete.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 231
where (p̃k )ρ denotes the structure coefficient. Note by Lemmas 5.11 and
5.14 deg∞ (p̃k ) = k so that the summation in (5.47) is over all ρ with
kρk∞ ≤ k. Then it follows from (5.44)
X
Ep,n p̃k = (p̃k )ρ Ep,n p]ρ
kρk∞ ≤k
X
= (p̃k )1r n↓r .
2r≤k
March 3, 2015 14:1 9197-Random Matrices and Random Partitions ws-book9x6 page 232
Also, by (5.40),
Ep,n p]ρ p]σ = Ep,n p]ρ∪σ + Ep,n lower terms
where lower terms means a linear combination of p]τ ’s with kτ k∞ < kρk∞ +
kσk∞ .
Again by (5.48),
↓r
] n , ρ ∪ σ = (1r ),
Ep,n pρ∪σ =
0, otherwise.
In summary, we have
X
Ep,n p̃2k = (p̃k )1r1 (p̃k )1r2 n↓(r1 +r2 ) + Ep,n lower terms
2r1 ≤k,2r2 ≤k
In particular, we have
(2m)↓m 2
Ep,n p̃22m = n↓2m + O n2m−1
m!
and
Ep,n p̃22m+1 = O n2m .
Therefore it follows
( 2
(2m)↓m
Ep,n p̃2k , k = 2m,
→ m!
nk 0, k = 2m + 1.
The proof is now complete.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 233
Remark 5.1. One can derive the strong form of limit shape theorem, i.e.,
Theorem 5.2, using the equivalence between weak topology and uniform
topology. The interested reader is referred to Theorem 5.5 of Ivanov and
Olshanski (2002).
√
Inserting (5.51) back into (5.50) and dividing by ( knk/2 )l+1 , we get
n↓kl 1
ηkl ηk = ηkl+1 + l ηkl−1 + √ lower terms, (5.52)
n↓k(l−1) nk
( kn )l+1
k/2
[(k−1)/2]
X k ]
p̃k+1 (λ) = (k + 1) nj p (λ) + n(k+1)/2 p̃k+1 (Ω)
j=0
j k−2j
+lower terms with deg1 (·) ≤ k − 1. (5.55)
A nontrvial point in (5.55) is the switch between two weight filtrations. Its
proof is left to the reader. See also Proposition 7.3 of Ivanov and Olshanski
(2002) for details.
[(k−1)/2]
p]k (λ)
X k k−j
= (−1)j qk−2j .
nk/2 j=0
k−j j
[k/2] ∞ √ √
h Z
X k−j
(−1)j uk−2j Ψλ ( nu) − n|u| du
=
j=0
j −∞
Z ∞ √ i
− uk−2j n (Ω(u) − |u|) du
−∞
[k/2]
X k−j 2
= (−1)j
j (k + 2 − 2j)(k + 1 − 2j)
j=0
p̃
k+2−2j (λ) √
· (k+1−2j)/2 − np̃k+2−2j (Ω) .
n
By the definition of (5.53) and noting
1 k−j 1 k+1−j
= ,
k + 2 − 2j j k+1−j j
we further get
[k/2]
X
j 2 k+1−j
Xn,k (λ) = (−1) qk+1−2j (λ).
k+1−j j
j=0
By (5.56),
2p]k+1 (λ)
Xn,k (λ) =
(k + 1)n(k+1)/2
1
+ (k+1)/2 lower terms with deg1 (·) ≤ k.
n
Since the remainder terms of negative degree do not affect the asymptotics,
then we can use Theorem 5.5 to conclude the proof.
To conclude this section, we remark that Theorem 5.5 for character ra-
tios is of independent interest. Another elegant approach was suggested by
Hora (1998), in which a central limit theorem was established for adjacency
operators on the infinite symmetric group. Still, Fulman (2005, 2006) de-
veloped the Stein method and martingale approach to prove asymptotic
normality for character ratios.
March 3, 2015 14:29 9197-Random Matrices and Random Partitions ws-book9x6 page 237
Theorem 5.7. Assume 0 < x1 < · · · < xm < 2, then under (Pn , Pp,n ),
1 d
1
√ Ξn (xi ), 1 ≤ i ≤ m −→ (ξi , 1 ≤ i ≤ m), n → ∞
2π log n
where ξi , 1 ≤ i ≤ m are independent normal random variables.
Remark 5.2. (i) The work of this section, in particular Theorems 5.6
and 5.7, are motivated by Gustavsson (2005), in which he investigated
the Gaussian fluctuation of eigenvalues in the GUE. There is a surprising
similarity between Plancherel random partitions and GUE from the view-
point of asymptotics, though no direct link exists between two finite models.
(ii) Compared with the uniform random partitions, the normalizing con-
√
stant log n is much smaller than n1/4 , see Theorem 4.5. This means that
Plancherel Young diagrams concentrated more stably around their limit
shape.
(iii) The random process Ξn (x) weakly converges to a Gaussian white
noise in the finite dimensional sense. Thus one cannot expect a usual pro-
cess convergence for Ξn in the space of continuous functions on [0, 2].
March 31, 2015 12:3 9197-Random Matrices and Random Partitions ws-book9x6 page 238
(iv) As we will see, if xn → x where 0 < x < 2, then (5.57) still holds
for Ξn (xn ), namely
Ξn (xn ) d
−→ N 0, %2 (x) ,
√
1 n → ∞. (5.58)
2π
log n
(v) Since %(x) → ∞ as x → 0 or 2, the normal fluctuation is no longer
true at either 0 or 2. In fact, it was proved
√
ψλ (0) − 2 n d
−→ F2 , n → ∞
n1/6
where F2 is Tracy-Widom law.
Proof. Fix −2 < u < 2 and assume λ ∈ Pn . A key step is to express the
√ √
error Ψλ ( n u) − nΩ(u) in terms of ψλ and ω. Let local extrema consist
of two interlacing sequences of points
ǔ1 < û1 < ǔ2 < û2 < · · · < ǔm < ûm < ǔm+1 ,
where ǔi ’s are the local minima and ûi ’s are the local maxima of the function
√
Ψλ ( n ·). Without loss of generality, we may and will assume that u is
√
between ûk and ǔk+1 for some 1 ≤ k ≤ m. Denote by n(xn , xn ) and
√ √ √ √
n(x∗n , x∗n ) the projections of ( n u, Ψλ ( n u)) and n(u, Ω(u)) in the
line u = v, respectively. Then we obviously have
√ √ √
Ψλ ( n u) − n Ω(u) = 2 n(xn − x∗n ).
According to Theorem 5.2, it follows
P
xn − x∗n −→ 0, n→∞
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 239
and so
P 1
xn −→ x := (Ω(u) + u).
2
√ √
On the other hand, if we let 2hn be the distance between n(xn , xn ) and
√ √
( n u, Ψλ ( n u)), then we have
√ √
n(xn − x∗n ) = hn − n ω(x∗n )
√ √
= hn − n ω(xn ) + n(ω(xn ) − ω(x∗n )). (5.60)
Now using the Taylor expansion for the function ω at xn and solving equa-
tion (5.60), we obtain
√
√ hn − n ω(xn )
n(xn − x∗n ) = ,
1 − ω 0 (x̃n )
where x̃n is between xn and x∗n .
P
Since xn , x∗n −→ x ∈ (0, 2), then it holds
1 P 1 1 ω(x) − x
−→ = arccos .
1 − ω 0 (x̃n ) 1 − ω 0 (x) π 2
√
Hence it suffices to prove hn − n ω(xn ) after properly scaled converges in
distribution. Observe that
√ √
ψλ ( n xn + 1) ≤ hn ≤ ψλ ( n xn )
since u is between ûk and ǔk+1 .
P
Note xn −→ x ∈ (0, 2). Then for each subsequence {n0 } of integers
there exists a further subsequence {n00 } ⊆ {n0 } such that xn00 → x a.e.
Thus by (5.58) it holds
√ √
ψλ ( n00 xn00 ) − n00 ω(xn00 ) d
−→ N 0, %2 (x) .
1
√ 00
2π log n
By a standard subsequence argument, it holds
√ √
ψλ ( n xn ) − n ω(xn ) d
−→ N 0, %2 (x) .
1
√
2π log n
√
Similarly, since n ω(xn + √1n ) − ω(xn ) = Op (1),
√ √
ψλ ( n xn + 1) − n ω(xn ) d
−→ N 0, %2 (x) .
1
√
2π log n
In combination, we have
√
hn − n ω(xn ) d
−→ N 0, %2 (x) .
1
√
2π log n
We conclude the proof.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 240
Let us now turn to the proof of Theorems 5.6 and 5.7. A basic strat-
egy is to adapt the conditioning argument—the Poissonization and de-
Poissonization techniques. Define the Poissonized Plancherel measure Qp,θ
on P as follows:
d 2
λ
Qp,θ (λ) = e−θ θ|λ|
|λ|!
∞
X θn
= e−θ Pp,n (λ)1(λ∈Pn ) , λ ∈ P (5.61)
n=0
n!
where θ > 0 is a model parameter. This is a mixture of a Poisson random
variable with mean θ and the Plancherel measures. Let Ξθ (x) be given by
(5.57) with n replaced by θ, namely
√ √
Ξθ (x) = ψλ θ x − θ ω(x), x ≥ 0, λ ∈ P. (5.62)
Theorem 5.10. Assume 0 < x1 < · · · < xm < 2. Under (P, Qp,θ ),
1 d
1
√ Ξθ (xi ), 1 ≤ i ≤ m −→ (ξi , 1 ≤ i ≤ m), θ → ∞
2π
log θ
where ξi , 1 ≤ i ≤ m are independent normal random variables.
Before giving the proof of Theorems 5.9 and 5.10, let us prove Theorems
5.6 and 5.7 with the help of the de-Poissonization technique.
Proof. We need only give the proof of the lower bound, since the upper
bound is similar. Let X be a Poisson random variable with mean θn+ . Then
we have EX = θn+ , V arX = θn+ and the following tail estimate
p
εn := P |X − θn+ | > n(log n)α
= O log−α n .
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 241
It follows by (5.61)
Qp,X (λ) = EPp,X (λ), λ∈P
and so for any event A ⊆ P
Qp,X (A) = EPp,X (A), A ⊆ P.
Note
EPp,X (A) = EPp,X (A)1(X<n) + EPp,X (A)1(X≥n) .
Trivially,
EPp,X (A)1(X<n) ≤ P (X < n) ≤ εn . (5.64)
In addition, set A = {λ ∈ P : ψλ (x) ≤ z}. Then using a similar argument
to that of Lemma 2.4 of Johansson (1998b), A is a monotonoic event under
(Pp,n , n ≥ 1), namely Pp,n+1 (A) ≤ Pp,n (A). Hence it follows
EPp,X (A)1(X≥n) ≤ Pp,n (A). (5.65)
Combining (5.64) and (5.65) together implies the lower bound.
Proof of Theorem 5.6. Set
√ √
nx nx
+
xn = p , x−
n =p
θn+ θn−
where θn± are as in Lemma 5.19. Trivially, x+ −
n , xn → x. Also, since 0 <
α < 1, then it follows
p √
θn± − n
√ → 0, n → ∞.
log n
Note
√ √ p p p
ψλ ( n x) − n ω(x) ψλ ( θn± x±n)− θn± ω(x±
n) log θn±
√ = p · √
log n log θn± log n
p √
θn± ω(x± ) − nω(x)
+ √n (5.66)
log n
and as n → ∞
p √ p
θn± ω(x±
n)− n ω(x) log θn±
√ → 0, √ → 1. (5.67)
log n log n
On the other hand, by Theorem 5.9, under P, Qp,θn±
p p
ψλ ( θn± x±n)− θn± ω(x±
n) d
−→ N 0, %2 (x) .
p
1 ±
2π log θn
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 242
Hence it follows from (5.66) and (5.67) that under P, Qp,θn±
√ √
ψλ ( n x) − n ω(x) d
−→ N 0, %2 (x) .
1
√
2π
log n
Taking Lemma 5.19 into account, we now conclude the proof.
To prove Theorem 5.9, we will again apply the Costin-Lebowitz-
Soshnikov theorem for determinantal point processes, see Theorem 3.7. To
do this, we need the following lemma due to Borodin, Okounkov and Ol-
shanski (2000), in which they proved the Tracy-Widom law for the largest
parts. Set for λ = (λ1 , λ2 , · · · , λl ) ∈ P
We will postpone the proof to Section 5.5. Now we are ready to give
Proof of Theorem 5.9. Fix 0 < x < 2. It suffices to show that for any
z∈R
%(x) p
Qp,θ Ξθ (x) ≤ log θ z → Φ(z), θ → ∞, (5.71)
2π
where Φ denotes the standard normal distribution function. Equivalently,
it suffices to show that for any z ∈ R
√ √
Qp,θ ψλ ( θ x) − d θ xe ≤ aθ → Φ(z),
where
√ %(x) p
aθ := aθ (x, z) = θ(ω(x) − x) + log θ z. (5.72)
2π
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 243
and
∞
X ∞
X
V arp,θ (Nk ) = Kθ (i, i) − Kθ (i, j)2
i=k i,j=k
∞ k−1
X X
= Kθ (i, j)2
i=k j=−∞
∞ k−1 √
X X θ √ √ √ √ 2
= Ji (2 θ)Jj+1 (2 θ) − Ji+1 (2 θ)Jj (2 θ) .
j=−∞
i−j
i=k
To figure out these infinite sums, we need some precise asymptotic be-
haviours of Bessel functions in the whole real line. They behave rather
differently in three regions so that we must take more care in treating the
sums over two critical values. The lengthy computation will appear in the
forthcoming paper, see Bogachev and Su (2015).
Having (5.75), it is natural to ask how fast it converges. This was first
studied by Fulman. In fact, in a series of papers, he developed a Stein
method and martingale approach to the study of the Plancherel measure. In
particular, Fulman (2005, 2006) obtained a speed of n−s for any 0 < s < 1/2
and conjectured the correct speed is n−1/2 . Following this, Shao and Su
(2006) confirmed the conjecture to get the optimal rate. The main result
reads as follows.
Theorem 5.11.
Pp,n λ ∈ Pn : Wn (λ) ≤ x − Φ(x) = O(n−1/2 ).
sup
−∞<x<∞
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 245
The basic strategy of the proof is to construct a Wn0 (λ) such that
0
Wn (λ), Wn (λ) is an exchangeable pair and to apply the Stein method
to Wn (λ), Wn0 (λ) . Let us begin with a Bratelli graph, namely an oriented
graded graph G = (V, E). Here the vertex set V = P = ∪∞ n=0 Pn and there
is an oriented edge from λ ∈ Pn to Λ ∈ Pn+1 if Λ can be obtained from λ
by adding one square, denoted by λ % Λ, see Figure 5.4 below.
Let us compute the quotient Hλ /HΛ . Assume that the new square is located
in the rth row and sth column of the diagram Λ. Since the squares outside
the rth row or sth column have equal hook lengths in the diagrams λ and
Λ, we have by the hook formula,
Y hri (λ) r−1
s−1 Y hjs (λ)
Hλ
= ,
HΛ h (λ) + 1 j=1 hjs (λ) + 1
i=1 ri
Then it follows
s−1 k−1
Y xk − ym
Y hri (λ)
=
i=1
hri (λ) + 1 m=1 xk − xm
and
r−1 q+1
Y hjs (λ) Y xk − ym−1
= .
j=1
hjs (λ) + 1 xk − xm
m=k+1
Thus we rewrite
k−1
Y xk − ym q+1
Hλ Y xk − ym−1
= =: ak .
HΛ x − xm
m=1 k
xk − xm
m=k+1
It remains to check
q+1
X
ak = 1. (5.76)
k=1
To do this, note that these numbers coincides with the coefficients of the
partial fraction expansion
Qq q+1
i=1 (u − yi ) ak
X
Qq+1 = .
i=1 (u − xi )
u − xk
k=1
Multiplying both sides by u and letting u → ∞ yields (5.76).
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 247
(n) (n)
To construct an exchange pair, we introduce a Markov chain X0 , X1 ,
(n)
· · · , Xk , · · · with state space Pn and transition probability
(n) (n)
pn (λ, µ) := P X1 = µX0 = λ
dµ
= ]P (λ, µ), (5.77)
(n + 1)dλ
where P (λ, µ) = τ ∈ Pn−1 , τ % λ, τ % µ .
Proof. Note the following formula (see the note following the proof of
Lemma 3.6 in Fulman (2005))
1 X
]P (λ, µ) = χµ (π)χλ (π) r1 (π) + 1 , (5.79)
n!
π∈Sn
where r1 (π) is the number of fixed points in π. Hence it follows from (5.77)
X
pn (λ, µ)
µ∈Pn
X dµ 1 X
= χµ (π)χλ (π) r1 (π) + 1
(n + 1)dλ n!
µ∈Pn π∈Sn
1 X 1 X
= dµ χµ (π) χλ (π) r1 (π) + 1 . (5.80)
(n + 1)dλ n!
π∈Sn µ∈Pn
By (2.11),
π = 1n ,
1 X 1,
dµ χµ (π) = (5.81)
n! 0, π 6= 1n .
µ∈Pn
Proof. By definition,
(n) (n)
EWn0 (λ) = EWn X1 X0 = λ
X
= Wn (µ)pn (λ, µ)
µ∈Pn
n−1 X
χµ 1n−2 , 2 ]P (λ, µ).
= √ (5.84)
(n + 1) 2dλ µ∈Pn
Lemma 5.25.
2 1 2(n − 1)(n − 2)2 χλ 1n−3 , 3
E Wn0 (λ) = 1− + ·
n n(n + 1) dλ
2 χ 1n−4 , 22
(n − 1)(n − 2)(n − 3) λ
+ · . (5.85)
2n(n + 1) dλ
Proof. Similarly to (5.84), it follows
X
E(Wn0 (λ))2 = Wn2 (µ)pn (λ, µ)
µ∈Pn
(n − 1)2 X χ2µ 1n−2 , 2
= ]P (λ, µ). (5.86)
2(n + 1)dλ dµ
µ∈Pn
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 249
where ](π) is the number of pairs of (σ, τ ) such that σ and τ come from
the same conjugacy class 1n−2 , 2 and σ ◦ τ = π. See also Lemma 3.4 of
Fulman (2005).
Note σ ◦ τ = π can assume only values in three distinct conjugacy
classes: 1n , 1n−3 , 3 , 1n−4 , 22 , and ](π), χλ (π) and r1 (π) are all class
functions. It is easy to see
n
] 1n =
2
n
] 1n−3 , 3 = 2(n − 2)
2
n n − 2
] 1n−4 , 22 =
.
2 2
In combination, we easily get the desired conclusion (5.85).
As a direct consequence, we obtain the following
Corollary 5.3.
1
V arp,n (Wn ) = V arp,n Wn0 = 1 − .
n
The last lemma we need is to control the difference between Wn (λ) and
Wn0 (λ).
where
1
A = 1− ,
n
2(n − 1)(n − 2)2 χλ (1n−3 , 3)
B= · ,
n(n + 1) dλ
(n − 1)(n − 2)(n − 3)2 χλ (1n−4 , 22 )
C= · ,
2n(n + 1) dλ
4 (n − 1)2 χ2 (1n−2 , 2)
D= −1 · λ 2 .
n+1 2 dλ
What we next need is to compute explicitly Ep,n (A + B + C + D)2 . We
record some data as follows.
Ep,n AB = Ep,n AC = 0;
(n − 1)2 4
Ep,n AD = 2
−1 ;
n n+1
(n − 1)2 (n − 2)3 (n − 3)2 χλ 1n−3 , 3 χλ 1n−4 , 22
Ep,n BC = Ep,n ·
n2 (n + 1)2 dλ dλ
2 3 2
(n − 1) (n − 2) (n − 3) 1 X
χλ 1n−3 , 3 χλ (1n−4 , 22 )
=
n2 (n + 1)2 n!
λ∈Pn
= 0;
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 252
(n − 1)3 (n − 2)2 (n − 3) χλ 1n−3 , 3 χ2λ 1n−2 , 2
Ep,n BD = − Ep,n ·
n(n + 1)2 dλ d2
n−3
2 λn−2
3 2
(n − 1) (n − 2) (n − 3) 1 X χλ 1 , 3 χλ 1 ,2
=− 2
n(n + 1) n! dλ
λ∈Pn
2
12(n − 1)(n − 2) (n − 3)
=− ;
n3 (n + 1)2
(n − 1)3 (n − 2)(n − 3)3 χλ 1n−4 , 22 χ2λ 1n−2 , 2
Ep,n CD = − Ep,n ·
4n(n + 1)2 dλ d2
n−4 2
2 λn−2
(n − 1)3 (n − 2)(n − 3)3 1 X χλ 1 , 2 χλ 1 ,2
=− 2
4n(n + 1) n! dλ
λ∈Pn
3
2(n − 1)(n − 2)(n − 3)
=− ;
n3 (n + 1)2
Lemma 5.28.
det L(xi , xj ) 2`×2`
Qp,θ (λ) = , (5.92)
det(1 + L)
where xi = ai , xi+` = −bi , 1 ≤ i ≤ `.
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 254
As a direct consequence,
X
eθ = eθ Qp,θ (λ)
λ∈P
X
= det L(X) = det(1 + L).
X
where in the last two equations we used the properties of Fredholm deter-
minants. On the other hand,
Y X Y
Ep,θ 1 + g(x) = 1 + g(x) Qp,θ F(λ) = X
x∈F (λ) X⊆Z+ 12 x∈X
X X Y
= g(x)Qp,θ F(λ) = X
X⊆Z+ 12 Y :Y ⊆X x∈Y
X Y X
= g(x) Qp,θ F(λ) = X
Y ⊆Z+ 21 x∈Y X:Y ⊆X
X Y
= g(x)%(Y ). (5.94)
Y ⊆Z+ 12 x∈Y
Lemma 5.30.
ML (x, y) = M (x, y).
√
Proof. Fix x, y ∈ Z + 12 and set z = θ. Note M and L are a function
of z. We need to prove for all z ≥ 0,
M + M L − L = 0. (5.95)
March 5, 2015 15:59 9197-Random Matrices and Random Partitions ws-book9x6 page 256
X− = X ∩ Z≤0 − 12 . If X ⊆ X (λ) + 12 , then by (5.97 ), X+ ⊆ F(λ) and
there exists a finite subset S ⊆ Z≤0 − 21 \ X− such that S ⊆ F(λ). This
implies
1
Qp,θ X ⊆ X (λ) +
2
1
= Qp,θ ∃S ⊆ Z≤0 − \ X− : X+ ∪ S ⊆ F(λ) . (5.98)
2
By exclusion-inclusion principle, the right hand side of (5.98) becomes
X
(−1)|S| Qp,θ X+ ∪ S ⊆ F(λ)
S⊆ Z≤0 − 21 \X−
X
= (−1)|S| %(X+ ∪ S)
S⊆ Z≤0 − 21 \X−
X
(−1)|S| det M (X+ ∪ S) ,
= (5.99)
S⊆ Z≤0 − 21 \X−
Equivalently,
1 − J(k, k) = J(−k − 1, −k − 1), k ∈ Z≥0 . (5.100)
Note for any k ∈ Z,
∞
X (2k + m + 2)↑m θk+m+1
J(k, k) = (−1) ,
m=0
((k + m + 1)!)2 m!
where we use the symbol (x)↑m = x(x + 1) · · · (x + m − 1). We need to show
∞
X (2k + m + 2)↑m θk+m+1
1− (−1)m
m=0
((k + m + 1)!)2 m!
∞
X (−2k + l + 2)↑l θ−k+l
= (−1)l . (5.101)
(Γ(−k + l + 1))2 l!
nml=0
Examine the right hand side of (5.101). The terms with l = 0, 1, · · · , k − 1
vanish because then 1/Γ(−k + l + 1) = 0. The term with l = k is equal to 1.
Next the terms with l = k + 1, · · · , 2k vanish because for these values of l,
the expression (−2k+l)↑l vanishes. Finally, for l ≥ 2k+1, say l = 2k+1+m,
(−2k + l)↑l θ−k+l (m + 1)↑l θk+m+1
(−1)l = (−1)m+1
(Γ(−k + l + 1))2 l! ((k + m + 1)!)2 (2k + 1 + m)!
(2k + m + 2)↑m θk+m+1
= (−1)m+1 .
((k + m + 1)!)2 m!
Thus we have proved (5.100).
Proof of Lemma 5.20 Fix X = (x1 , x2 , · · · , xk ) ⊆ Z. Then according to
(5.27), (5.98) and (5.99), we have
ρk (x1 , · · · , xk ) = Qp,θ λ ∈ P : x1 , · · · , xk ∈ X (λ)
X
(−1)|S| det M (X+ ∪ S) . (5.102)
=
S⊆(Z≤0 − 12 )\X−
S⊆(Z≤0 − 21 )\X−
Bibliography
261
March 5, 2015 16:11 9197-Random Matrices and Random Partitions ws-book9x6 page 262
Bibliography 263
Bibliography 265
Index
267
February 2, 2015 10:5 9197-Random Matrices and Random Partitions ws-book9x6 page 268
Index 269
Index 271
Wasserstein distance, 21
weak topology, 233
wedge product, 81
weight filtration, 225, 235
weight grading, 224, 226
Weyl formula, 34, 53
Widom, 219
Wieand, 55
Wigner semicircle law, 14, 95, 113,
133, 142
Zeitouni, 28, 90