More generally let X and Y be Banach spaces and let A: X 
ightarrow Y, B: Y 
ightarrow X be bounded linear operators. Then the following commutation formula is true:

(75.1)

frac{lambda}{lambda + AB} + A frac{1}{lambda + BA} B = mathbf{1}_Y

in the sense that of 0 
e -lambda in 
ho(BA) = resolvent set of BA, then -lambda in 
ho(AB) and

frac{1}{lambda}(mathbf{1}_Y - A frac{1}{lambda + BA} B)

is the resolvent of AB, and vice versa. In particular, we see that (74.1) is true for general bounded operators in Banach spaces. (75.1) is also true, suitably interpreted, for certain classes of unbounded operators (see Deift [D. M. Jour, 1978]).

We now prove (71.2): Let mu and f_i, g_j in L^2(dmu) be given, then

idotsint det (f_i(x_k))_{i,k = 1, cdots, N} det (g_j(x_k))_{j,k = 1, cdots, N} dmu(x_1) cdots dmu(x_N)

= sum_{sigma, 	au} sgn(sigma)sgn(	au) int f_{sigma(1)}(x_1) cdots f_{sigma(N)}(x_N) 	imes g_{	au(1)}(x_1) cdots g_{	au(N)}(x_N) dmu(x_1) cdots dmu(x_N)

= sum_{sigma, 	au} sgn(sigma)sgn(	au) [int f_{sigma(1)}(x_1) g_{	au(1)}(x_1) dmu(x_1)] cdots [int f_{sigma(N)}(x_N) g_{	au(N)}(x_N) dmu(x_N)]

= sum_{sigma, 	au} sgn(sigma circ 	au^{-1}) [int f_{sigma circ 	au^{-1}(	au(1))}(x_1) g_{	au(1)}(x_1) dmu(x_1)] cdots [int f_{sigma circ 	au^{-1}(	au(N))}(x_N) g_{	au(N)}(x_N) dmu(x_N)]

= sum_sigma sum_{sigma circ 	au^{-1}: 	au in S_N} sgn(sigma circ 	au^{-1}) [int f_{sigma circ 	au^{-1}(	au(1))}(x) g_{	au(1)}(x) dmu(x)] cdots [int f_{sigma circ 	au^{-1}(	au(N))}(x) g_{	au(N)}(x) dmu(x)]

= sum_sigma sum_{sigma circ 	au^{-1}: 	au in S_N} sgn(sigma circ 	au^{-1}) [int f_{sigma circ 	au^{-1}(1)}(x) g_{1}(x) dmu(x)] cdots [int f_{sigma circ 	au^{-1}(N)}(x) g_{N}(x) dmu(x)]

= sum_sigma sum_{	au^prime in S_N} sgn(	au^prime) [int f_{	au^prime(1)}(x) g_1(x) dmu(x)] cdots [int f_{	au^prime(N)}(x) g_{N}(x) dmu(x)]

= sum_sigma det (int f_j(x) g_k(x))_{j,k=1, cdots, N}

=N! det (int f_j(x) g_k(x) dmu(x))_{j,k=1, cdots, N}

as desired. Box

The particular case

dmu(x) = sum_{k=1}^M delta_{lambda_k}(x)

for a given set of numbers lambda_1, cdots, lambda_M, then the LHS of (71.2) becomes

sum_{k_1=1}^M cdots sum_{k_N=1}^M idotsint det egin{pmatrix} f_1(x_1) & cdots & f_1(x_N) \ vdots & ddots & vdots \ f_N(x_1) & cdots & f_N(x_N) end{pmatrix} det egin{pmatrix} g_1(x_1) & cdots & g_1(x_N) \ vdots & ddots & vdots \ g_N(x_1) & cdots & g_N(x_N) end{pmatrix} delta_{lambda_{k_1}}(x_1) cdots delta_{lambda_{k_N}}(x_N)

= sum_{k_1=1}^M cdots sum_{k_N=1}^M det egin{pmatrix} f_1(lambda_{k_1}) & cdots & f_1(lambda_{k_N}) \ vdots & ddots & vdots \ f_N(lambda_{k_1}) & cdots & f_N(lambda_{k_N}) end{pmatrix} det egin{pmatrix} g_1(lambda_{k_1}) & cdots & g_1(lambda_{k_N}) \ vdots & ddots & vdots \ g_N(lambda_{k_1}) & cdots & g_N(lambda_{k_N}) end{pmatrix}

= N! sum_{1 le k_1 < k_2 < cdots < k_N le M} det egin{pmatrix} f_1(lambda_{k_1}) & cdots & f_1(lambda_{k_N}) \ vdots & ddots & vdots \ f_N(lambda_{k_1}) & cdots & f_N(lambda_{k_N}) end{pmatrix} det egin{pmatrix} g_1(lambda_{k_1}) & cdots & g_1(lambda_{k_N}) \ vdots & ddots & vdots \ g_N(lambda_{k_1}) & cdots & g_N(lambda_{k_N}) end{pmatrix}

On the other hand, the RHS of (71.2) gives

N! det (sum_{i=1}^M f_j(lambda_i) g_k(lambda_i))_{j,k=1, cdots, N}

Equating these expressions we see that

if

F_{N 	imes M} = (F_{ij}) = { f_i(lambda_j) } = egin{pmatrix} f_1(lambda_1) & cdots & f_1(lambda_M) \ vdots & ddots & vdots \ f_N(lambda_1) & cdots & f_N(lambda_M) end{pmatrix}

and

G_{M 	imes N} = (G_{ij}) = { g_j(lambda_i) } = egin{pmatrix} g_1(lambda_1) & cdots & g_N(lambda_1) \ vdots & ddots & vdots \ g_1(lambda_M) & cdots & g_N(lambda_M) end{pmatrix}

we obtain the classical Cauchy-Binet formula

(79.1)

det (FG) = sum_{1 le j_1 < cdots < j_N le M} det egin{pmatrix} F_{1j_1} & F_{1j_2} & cdots & F_{1j_N} \ vdots & vdots & ddots & vdots \ F_{Nj_1} & F_{Nj_2} & cdots & F_{Nj_N} end{pmatrix} det egin{pmatrix} G_{j_1 1} & G_{j_1 2} & cdots & G_{j_1 N} \ vdots & vdots & ddots & vdots \ G_{j_N 1} & G_{j_N 2} & cdots & G_{j_N N} end{pmatrix}

e.g. if F = egin{pmatrix} a_1 & a_2 & a_3 \ b_1 & b_2 & b_3 end{pmatrix}, G = egin{pmatrix} c_1 & d_1 \ c_2 & d_2 \ c_3 & d_3 end{pmatrix}

then

det egin{pmatrix} a_1 & a_2 & a_3 \ b_1 & b_2 & b_3 end{pmatrix} egin{pmatrix} c_1 & d_1 \ c_2 & d_2 \ c_3 & d_3 end{pmatrix}

= det egin{pmatrix} a_1 & a_2 \ b_1 & b_2 end{pmatrix} det egin{pmatrix} c_1 & d_1 \ c_2 & d_2 end{pmatrix} + det egin{pmatrix} a_1 & a_3 \ b_1 & b_3 end{pmatrix} det egin{pmatrix} c_1 & d_1 \ c_3 & d_3 end{pmatrix} + det egin{pmatrix} a_2 & a_3 \ b_2 & b_3 end{pmatrix} det egin{pmatrix} c_2 & d_2 \ c_3 & d_3 end{pmatrix}

Of course if N = M, (79.1) is just the familiar fact that

det FG = det F det G

and for N > M, both LHS and RHS are 0 (why?)

We will also need the following basic definition.

Let dmu(x) be a Borel measure on mathbb{R} with all moments finite,

(80.1)

int_{mathbb{R}} |x|^m dmu(x) < infty, m = 0, 1, 2, cdots.

Suppose that the support of dmu is infinite, i.e., dmu is not a finite linear combination of delta functions, then by the Gram-Schmidt procedure there exist unique monic polynomials

pi_k (x) = x^k + cdots, k ge 0

which are orthogonal with respect to dmu i.e.,

int_{mathbb{R}} pi_k(x) pi_j(x) dmu(x) = 0 for j 
e k.

For

gamma_k = (int_{mathbb{R}} pi_k^2(x) dmu(x))^{-frac{1}{2}} > 0, k=0, 1, 2, cdots,

set

p_k(x) = gamma_k pi_k(x), k ge 0.

The polynomials {p_k} are the orthonormal polynomials with respect to dmu, i.e.,

int_{mathbb{R}} p_k(x) p_j(x) dmu(x) = delta_{kj}, j,k=0, 1, 2, cdots

The p_ks are called the orthonormal polynomials (w.r.t. dmu) and the pi_ks are called the monic orthogonal polynomials (w.r.t. dmu).

Exercise: what happens if dmu has finite support?

As we will see, the p_ks and pi_ks play a central role in RMT.


We now show how to compute certain basic statistics for invariant unitary ensembles with probability distributions

P_N(M) dM = frac{e^{-tr Q(M)} dM}{Z_N}

where Q(x) 
ightarrow + infty sufficiently rapidly (see 84.x - 85.x below) as |x| 
ightarrow infty.

In particular, we will compute

<f> = int f(M) P_N(M) dM

for f(M) of the form

(82.0)

f(M) = det(I + g(M))

for bounded functions g: mathbb{R} 
ightarrow mathbb{R}.

From the Hermitized analog of (68.1) we have

(82.1)

<f> = c_N^prime int_{lambda_1 < cdots < lambda_N} prod_{i=1}^N (1 + g(lambda_i)) prod_{1 le i < j le N} (lambda_i - lambda_j)^2 prod_{i=1}^N omega(lambda_i) d^N lambda

= c_N^prime int_{lambda_1 < cdots < lambda_N} prod_{i=1}^N (1 + g(lambda_i)) V^2(lambda) prod_{i=1}^N omega(lambda_i) d^N lambda

where

(82.2)

omega(x) = e^{-Q(x)}

(82.3)

V(lambda) = VanderMonde = det(lambda_i^{j-1})_{1 le i, j le N}

and

(82.4)

c_N^prime = (int_{lambda_1 < cdots < lambda_N} V^2(lambda) prod_1^N omega(lambda_i) d^N lambda)^{-1}

As the integrands in (82.1) and (82.4) are invariant under all permutation sigma

(lambda_1, cdots, lambda_N) 
ightarrow (lambda_{sigma_1}, cdots, lambda_{sigma_N}),

we may rewrite (82.1) (82.4) in the following symmetrized form

(83.1)

<f> = c_N int_{mathbb{R}^N} prod_{i=1}^N (1+g(lambda_i)) V^2(lambda) prod_1^N omega(lambda_i) d^N lambda

where

(83.2)

c_N = c_N^prime (N!)^{-1}

Now define

(83.3)

f_i(x) = g_i(x) = x^{i-1}, 1 le i le N

then

V(lambda)^2 = det (f_i(lambda_j))_{1 le i, j le N} det (g_i(lambda_j))_{1 le i, j le N}

Hence we can apply the generalized Cauchy-Binet formula to (83.1) with measure dmu(x) = omega(x) (1 + g(x)) (the fact that dmu(x) may be a signed measure, as opposed to a positive measure, clearly does not affect the proof of (83.1)).

We obtain

<f> = c_N^prime det (int f_j(lambda) g_k(lambda) dmu(lambda))_{j,k=1}^N

where we note from (82.4) that c_N^prime depends only on N and omega, but not on g(x).

We have

(84.1)

int f_j(lambda) g_k(lambda) dmu(lambda) = int f_j(lambda) g_k(lambda) (1 + g(lambda)) omega(lambda) dlambda

= int lambda^{j+k-2} (1 + g(lambda)) omega(lambda) dlambda.

Let p_j(lambda), j ge 0 be the orthogonal polynomials with respect to the measure omega(lambda) dlambda i.e.,

int_{mathbb{R}} p_k(lambda) p_j(lambda) omega(lambda) dlambda = delta_{jk}, j, k ge 0.

(Note that omega(lambda) dlambda = e^{-Q(lambda)} dlambda does not have finite support; also by "sufficient decay" we are assuming, at least, that e^{-Q(lambda)} dlambda has finite moments; note that this implies

int (V(lambda))^2 prod_{i=1}^N omega(lambda_j) d^N lambda < infty

so that P_N(M) dM is a normalized probability measure with finite moments).

Now as we can add rows and columns without changing a determinant, we have (recall p_j(lambda) = gamma_j pi_j(lambda))

(85.1)

<f> = c_N^prime det (int pi_j(lambda) pi_k(lambda) (1 + g(lambda)) omega(lambda) dlambda)_{0 le j,k le N-1}

= frac{c_N^prime}{prod_{j=0}^{N-1} gamma_j^2} det (int p_j(lambda) p_k(lambda) (1 + g(lambda)) omega(lambda) dlambda)_{0 le j,k le N-1}

Set

(85.2)

phi_j(x) = p_j(x) omega(x)^{frac{1}{2}}, j ge 0

have int phi_j(x) phi_k(x) dx = delta_{jk}, jak ge 0

then (85.1) takes the form

(85.3)

<f> = frac{c_N^prime}{prod_{j=0}^{N-1} gamma_j^2} det (int phi_j(lambda) phi_k(lambda) (1 + g(lambda)) omega(lambda) dlambda)_{0 le j,k le N-1}

= c_N^{prime prime} det (delta_{j,k} + int g(lambda) phi_j(lambda) phi_k(lambda) dlambda)_{j,k=0}^{N-1}

where again c_N^{prime prime} does not depend on g(lambda). Setting g = 0, we have f(M) = 1 and so LHS of (85.3) = 1, but RHS = c_N^{prime prime} det (delta_{j,k}) = c_N^{prime prime}, so c_N^{prime prime} =1.

Thus

(86.1)

<f> = det (delta_{j,k} + int_{mathbb{R}} phi_j(x) phi_k(x) g(x) dx)_{j,k=0}^{N-1}

Now we are primarily interested in the situation where the size N of the matrices become large. We see from (86.1) that <f> is expressed in terms of determinants of larger and larger size. Limits of this kind are generally very difficult to control. Fortunately, we can use (71.1) det (mathbf{1}_X + AB) = det (mathbf{1}_Y + BA) to reduce (86.1) to the (Fredholm) determinant on a fixed space (see below). Such limits are, generally speaking, easier to control; in particular if K_N is a trace class operator on a (fixed) Hilbert space H, and K_N 
ightarrow K is trace norm,

then

det (mathbf{1} + K_N) 
ightarrow det (mathbf{1} + K)

It is just a matter of continuity of the determinant in the trace norm.

Now let A: L^2(mathbb{R}) 
ightarrow mathbb{C}^N denote the bounded operator

(87.1)

(Ah)_j = int phi_j(x) g(x) h(x) dx, j=0, cdots, N-1, h in L^2(mathbb{R})

and let B: mathbb{C}^N 
ightarrow L^2(mathbb{R}) denote the bounded operator

(87.2)

(Ba)(x) = sum_{j=0}^{N-1} phi_j(x) a_j, a = (a_0, cdots, a_{N-1})^T in mathbb{C}^N

then AB maps mathbb{C}^N 
ightarrow mathbb{C}^N and for a in mathbb{C}^N

(87.3)

((AB)(a))_j = (A(sum_{k=0}^{N-1} phi_k(ullet) a_k))_j

= sum_{k=0}^{N-1} a_k (A phi_k)_j

= sum_{k=0}^{N-1} a_k (int phi_j(x) g(x) phi_k(x) dx)

and we see from (86.1) that

(88.1)

<f> = det (mathbf{1}_{mathbb{C}^N} + AB)

On the other hand, BA: L^2(mathbb{R}) 
ightarrow L^2(mathbb{R}) and for b in L^2(mathbb{R})

(BAh)(x) = (B ((int phi_j(y) g(y) h(y) dy)_{j=0}^{N-1}))(x)

= sum_{j=0}^{N-1} phi_j(x) int phi_j(y) g(y) h(y) dy

= int K(x, y) g(y) h(y) dy

where

(88.2)

K(x, y) = sum_{j=0}^{N-1} phi_j(x) phi_j(y)

is the so-called correlation kernel for the ensemble: K(x, y) plays a central role in RMT.

Thus

(88.3)

BA = K chi_g

where chi_g denotes multiplication by g(x) and K denotes the operator with kernel K i.e., K h(x) = int K(x, y) h(y) dy.

Assembling the above results we obtain the key formula

(89.1)

<f> = det (mathbf{1}_{L^2(mathbb{R})} + K chi_g)

Note that as A (and also B!) is finite rank, A is trace class and the Fredholm determinant in (89.1) is well-defined. There are analogous, but more complicated, formulae for eta = 1 and eta = 4 (see Ref (3)).

If Omega is a Borel set in mathbb{R} and

g(x) = - chi_Omega

then we see from (89.1) that the gap probability considered before is given by

(89.2)

Prob (no eigenvalues in Omega)

= <prod_{i=1}^N (1 - chi_Omega(lambda_i))>, chi_Omega = characteristic function of Omega

= det (mathbf{1}_{L^2(mathbb{R})} - K chi_Omega)

Note that if Omega = phi, then RHS = 1, i.e. Prob (no eigenvalues in phi) = 1, as it should be. On the other hand for Omega = mathbb{R}, Prob (no eigenvalues in mathbb{R}) = 0.

On the other hand, for any k = 0, 1, cdots, N-1

Kphi_k(x) = int sum_0^{N-1} phi_j(x) phi_j(y) phi_k(y) dy

= sum_0^{N-1} phi_j(x) delta_{jk}

= phi_k(x)

Thus 1 is an eigenvalue of K (of multiplicity N > 0) and so det(I - K) = 0, as it should be.

Now take Omega = (a, infty) for any a in mathbb{R}, then clearly

Prob (no eigenvalues in (a, infty)) = Prob ( lambda_{max} le a )

where lambda_{max} is the largest eigenvalues of M. Thus

(90.1)

Prob ( lambda_{max} le a ) = det (mathbf{1}_{L^2(mathbb{R})} - K chi_{(a, infty)})

This is a key formula in RMT for unitary ensembles.

Exercise

If Omega = Omega_1 cup cdots cup Omega_k is a union of disjoint sets in mathbb{R}, Omega_j cap Omega_k = phi for j 
e k, then we have computed

Prob (0 eigenvalues in Omega_1, ..., 0 eigenvalues in Omega_k) = Prob (no eigenvalues in Omega = cup_{j=1}^k Omega_j ).

Derive an explicit formula for

Prob ( n_1 eigenvalues in Omega_1, ..., n_k eigenvalues in Omega_k)

for any non-negative integers n_1, cdots, n_k. (see ref (3) pp 86-88).

We now show how to compute correlation functions for unitary ensembles (recall that 2-point correlation functions arose in the study of the non-trivial zeros of the Riemann zeta function in the first lecture).

If P_N(x) dx is the probability distribution for a system of N identical random particles

P_N(x_1, dots, x_N) symmetric in the x_is

then the n-point correlation function

R_n = R_n(x_1, dots, x_n), 1 le n le N

for P_N(x) dx is defined by

(92.1)

R_n(x_1, dots, x_n) = frac{N!}{(N - n)!} idotsint P_N(x_1, dots, x_n, x_{n+1}, dots, x_N) dx_{n+1} dots dx_N

Note that int R_n(x_1, dots, x_n) dx_1 dots dx_n 
e 1 and hence R_n is not a probability distribution (more later!).


推薦閱讀:
相关文章