• Aucun résultat trouvé

A log-Sobolev type inequality for free entropy of two projections

N/A
N/A
Protected

Academic year: 2022

Partager "A log-Sobolev type inequality for free entropy of two projections"

Copied!
11
0
0

Texte intégral

(1)

www.imstat.org/aihp 2009, Vol. 45, No. 1, 239–249

DOI: 10.1214/08-AIHP164

© Association des Publications de l’Institut Henri Poincaré, 2009

A log-Sobolev type inequality for free entropy of two projections

Fumio Hiai

a,1,2

and Yoshimichi Ueda

b,1,3

aGraduate School of Information Sciences, Tohoku University, Aoba-ku, Sendai 980-8579, Japan bGraduate School of Mathematics, Kyushu University, Fukuoka 810-8560, Japan Received 16 November 2006; revised 8 December 2007; accepted 2 January 2008

Abstract. We prove a kind of logarithmic Sobolev inequality claiming that the mutual free Fisher information dominates the microstate free entropy adapted to projections in the case of two projections.

Résumé. Nous prouvons un genre d’inégalité de Sobolev logarithmique qui montre que l’information de Fisher libre domine l’entropie de micro-états libre adaptée aux projections dans le cas de deux projections.

MSC:Primary 46L54; secondary 94A17; 60E15

Keywords:Logarithmic Sobolev inequality; Free entropy; Mutual free Fisher information

Introduction

The aim of this paper is to add a new result to Voiculescu’s liberation theory initiated in [20]. The mutual free Fisher informationϕand the mutual free informationiwere introduced there for subalgebras unlike those quantities in [19] for random variables. More precisely, what we will prove is the inequality

χproj(p, q)ϕ

Cp+C(1p):Cq+C(1q)

(0.1) for projectionsp, qin a tracialW-probability space under a natural assumption. Here,χprojdenotes the microstate definition of free entropy adapted to projections, which was proposed by Voiculescu [20], Section 14.2, and naturally appeared as the rate function in the large deviation principle for a pair of random projection matrices in [8]. One should remember that the original microstate free entropyχ is meaningless for projections asχ always takes−∞

for them. Our subsequent paper [10] will provide several basic properties ofχproj similar to those of the originalχ developed in [16–18].

At least to the best of our knowledge, no further work on the liberation theory has been made after the appearance of [20]. Our present result may add a new insight to what Voiculescu discussed in [20], Sections 14.1 and 14.2, though the situation we deal with is very restricted. In fact, our proof suggests that the inequality (0.1) should be regarded as a kind of logarithmic Sobolev inequality, whose original form is an inequality giving an upper bound of the (relative) entropy by the (relative) Fisher information (or Dirichlet form) up to a constant (see, e.g. [15], Section 9.2) and whose free analogs were obtained in [3,9,12]. From this point of view our result suggests a possible relation between−χproj andiat least for pairs of projections.

1Supported in part by Japan Society for the Promotion of Science, Japan–Hungary Joint Project.

2Supported in part by Grant-in-Aid for Scientific Research (B)17340043.

3Supported in part by Grant-in-Aid for Young Scientists (B)17740096.

(2)

1. Preliminaries

1.1. Free entropy for projections

ForN ∈Nlet U(N )be the N×N unitary group. LetG(N, k)denote the set of allN×N orthogonal projection matrices of rankk, that is, G(N, k)is identified with the Grassmannian manifold consisting ofk-dimensional sub- spaces inCN, and it is also identified with the homogeneous space U(N )/(U(k)⊕U(N−k)). The unitarily invariant probability measureγG(N ,k)onG(N, k)corresponds to the measure on U(N )/(U(k)⊕U(N−k))induced from the Haar probability measure on U(N ).

Let(p1, . . . , pn)be ann-tuple of projections in a tracialW-probability space(M, τ )withαi:=τ (pi), 1in.

Following Voiculescu’s proposal in [20], Section 14.2, we define thefree entropyχproj(p1, . . . , pn)of (p1, . . . , pn) as follows: Choosek(N, i)∈ {0,1, . . . , N}for eachN ∈Nand 1≤inin such a way that k(N, i)/Nαi as N → ∞ for 1≤in. For each m∈N and ε >0, Γproj(p1, . . . , pn;k(N,1), . . . , k(N, n);N, m, ε) denotes the set of(P1, . . . , Pn)n

i=1G(N, k(N, i))satisfying|trN(Pi1· · ·Pir)τ (pi1· · ·pir)|< ε for all 1≤i1, . . . , irn, 1≤rm, where trN stands for the normalized trace on theN×N matrices. Thenχproj(p1, . . . , pn)is defined to be

m∈Ninf, ε>0lim sup

N→∞

1 N2log

n

i=1

γG(N ,k(N ,i))

Γ

p1, . . . , pn;k(N,1), . . . , k(N, n);N, m, ε ,

which is known to be independent of the choices ofk(N, i) withk(N, i)/Nαi, 1≤in (see [10], Proposi- tion 1.1).

In this paper we are concerned with the case of two projections. Let(p, q)be a pair of projections in(M, τ )with α:=τ (p)andβ:=τ (q). Set

E11:=pq, E10:=pq, E01:=pq, E00:=pq, E:=1(E00+E01+E10+E11)

andαij:=τ (Eij)fori, j=0,1. ThenEandEij are in the center ofN := {p, q}and(ENE, τ|ENE)is isomorphic toL((0,1), ν;M2(C)), theL-algebra ofM2(C)-valued functions, whereνis a measure on(0,1)withν((0,1))= 1−1

i,j=0αij. HereEpEandEqEcorrespond to t(0,1)→ 1 0

0 0

and t

t (1t )

t (1t ) 1−t

,

respectively, andτ|ENEis represented as τ (a)=

1 0

tr2

a(t ) dν(t )

foraENEcorresponding toa(·)L((0,1), ν;M2(C)). In this way, the mixed moments of(p, q)with respect to τ are determined by the data(ν,{αij}1i,j=0). Althoughνis not necessarily a probability measure, we denote(ν):=

1

0

1

0log|xy|dν(x)dν(y)in the same fashion as in [16]. Furthermore we set ρ:=min{α, β,1−α,1−β} =1

2

1− 1 i,j=0

αij

, (1.1)

C:=ρ2B

|αβ|

ρ ,|α+β−1|

ρ

(1.2) (meant zero ifρ=0), where the functionB(s, t )ins, t≥0 was given in [8], Proposition 2.1. The following expression was obtained in [8] as a consequence of the large deviation principle for an independent pair of random projection matrices.

(3)

Lemma 1.1 ([8], Theorem 3.2, Proposition 3.3). Ifα00α11=α01α10=0,then χproj(p, q)=1

4(ν)+α01+α10 2

1

0

logxdν(x)+α00+α11 2

1

0

log(1−x)dν(x)−C, and otherwiseχproj(p, q)= −∞.

Note that the conditionα00α11=α01α10=0 is equivalent to α11=max{α+β−1,0}, α10=max{αβ,0},

α00=max{1−αβ,0}, α01=max{βα,0}, (1.3)

and in this case,α01+α10= |αβ|andα00+α11= |α+β−1|. 1.2. Mutual free Fisher information

Let AandB be two unital ∗-subalgebras in(M, τ ), which are assumed to be algebraically free. LetAB and W(AB)denote the subalgebra and the von Neumann subalgebra, respectively, generated byAB. LetδA:B be the derivation fromA∨Binto theA∨B-bimodule(A∨B)(A∨B)uniquely determined byδA:B(a)=a11a foraAandδA:B(b)=0 forbB. If there is an elementξL1(W(AB))such thatτ (ξ x)=τ )(δA:B(x)) forxAB, thenξ is called theliberation gradientof(A,B)and denoted byj (A:B). Voiculescu [20] introduced themutual free Fisher informationofArelative toBbyϕ(A:B):= j (A:B)22( · 2stands for theL2-norm with respect toτ) ifj (A:B)exists inL2(W(AB)); otherwiseϕ(A:B):= +∞. See [20] for more about the mutual free Fisher information.

Let(p, q)be a pair of projections in(M, τ )and setA:=Cp+C(1p),B:=Cq+C(1q). Then the liberation gradientj (A:B)and the mutual free Fisher informationϕ(A:B)were computed in [20]. Here, recall that the Hilbert transform of a functionf withf (x)/(1+ |x|)L1(R,dx)is defined to be

(Hf )(x):=lim

ε0(Hεf )(x) with(Hεf )(x):=

|xt|

f (t ) xt dt

whenever the limit exists almost everywhere. We need the following a-bit-improved version of [20], Proposition 12.7, because the original version is not applicable, in particular, to the free case when τ (p)=τ (q)=1/2 due to the L3-assumption with respect to dx rather than1(0,1)(x)x(1x)dx.

Lemma 1.2. With the same notations as in Section 1.1assume that α00α11=α01α10=0, ν has the densityf :=

dν/dx∈L3((0,1), x(1−x)dx)and moreover 1

0

α01+α10

x +α11+α00 1−x

f (x)dx <+∞. (1.4)

DefineX:=pqp+(1p)(1q)(1p)and φ (x):=(Hf )(x)+α01+α10

xα00+α11

1−x , 0< x <1.

Then

j (A:B)= [q, p]φ (EXE)L2(M, τ ) and hence

ϕ(A:B)= 1

0

φ (x)2f (x)x(1x)dx <+∞.

(4)

The assumption (1.4) can be reduced whenα=βorα+β=1. In fact, (1.4) is nothing whenα=β=1/2; it means 1

0x1f (x)dx <+∞whenα+β=1 butα=β. All the assumptions of Lemma 1.2 are satisfied, in particular, when pandqare free (see [21], Example 3.6.7). Note ([20], Propositions 5.17 and 9.3.c) thatj (A:B)=0 (or equivalently ϕ(A:B)=0) if and only ifpandqare free.

In [20], Section 12, the support ofνwas assumed to be an infinite set to guarantee thatAandBare algebraically free. As long asν=0, that is automatically satisfied from the assumption ofνhaving the density. In the case where ν=0 so thatρ=0 by (1.1), it follows thatp∈ {0,1}orq∈ {0,1}; hence Lemma 1.2 trivially holds.

Proof of Lemma 1.2. Since1

0(x(1x))1/2dx <+∞, the weighted version of M. Riesz’s theorem for Hilbert transform ([11], Theorem 8) shows that there is a constantCw>0 depending only on the weight functionw(x):=

1(0,1)(x)x(1x)such that for every functiong(whoseH gcan be defined)

H gw,3Cwgw,3 (1.5)

with the weighted normgw,p:=(1

0 |g(x)|px(1x)dx)1/pfor 1≤p <∞, and moreoverHεgH gw,3→0 as ε0 whenevergw,3<+∞. In what follows we use the same symbols as in [20], Section 12, with small exception;

p, q,A,Bandx, x1, x2are used instead ofP , Q,A, B andt, t1, t2, respectively. By the facts mentioned above one easily has

1 0

1 0

x1n+1x2n+1

x1x2x1nx2n x1x2

dν(x1)dν(x2)

= −lim

ε0

|x1x2|

x1n−1x1(1x1) f (x2)

x1x2f (x1)+x2n−1x2(1x2) f (x1) x2x1f (x2)

dx1dx2

= −2 lim

ε0

1

0

xn−1(Hεf )(x)f (x)x(1x)dx

= −2 1

0

xn1(Hf )(x)f (x)x(1x)dx.

Hence the assertion of [20], Lemma 12.6, can be changed to τ )δB:A

(pq)n

= −1 2

1 0

xn1(Hf )(x)f (x)x(1x)dx +1−α

2 1

0

xn1(x−1)dν(x)+α00+α11 2

1

0

xn1dν(x) (1.6)

under the assumptions of Lemma 1.2. The rest of the proof goes along the same line as [20], Proposition 12.7, with

replacing [20], Lemma 12.6, by (1.6).

2. Main result

For simplicity we hereafter writeϕ(p:q)for the mutual free Fisher informationϕ(Cp+C(1p):Cq+C(1q)) (see Section 1.2).

Theorem 2.1. The inequalityχproj(p, q)ϕ(p:q)holds under the same assumptions as in Lemma1.2.

The main idea of the proof is a random matrix approximation procedure based on the large deviation shown in [8]. In fact, we apply Bakry and Emery’s logarithmic Sobolev inequality in [1] to random projection matrix pairs (or probability measures on the product of two Grassmannian manifolds) and pass to the scaling limit as the matrix size goes to∞. Thus our inequality can be regarded as a kind of free probability counterpart of the logarithmic Sobolev inequality.

(5)

In the proof, we have to examine Bakry and Emery’sΓ2-criterion, which involves the Ricci curvature tensor. We thus need to compute the Ricci curvature tensor Ric(G(N, k))ofG(N, k). Letu(N )be the Lie algebra of U(N )and regardh(N, k):=u(k)⊕u(N−k)as a Lie subalgebra ofu(N ). The tangent spaceTPG(N, k)at eachPG(N, k) can be identified withg(N, k):=h(N, k), the orthocomplement ofh(N, k)inu(N )with respect to the Riemannian metric X, Y :=Re TrN(XY), where TrN is the usual trace onN ×N matrices. Choose the following complete orthonormal system ofg(N, k):

Eij:= 1

√2(eijej i), Fij:=

√√−1

2 (eij+ej i) (2.1)

with 1≤ik, k+1≤jN. According to well-known facts on compact matrix groups and O’Neill’s formula (see [6], Proposition 3.17 and Theorem 3.61, for example), the Ricci curvature tensor ofG(N, k)with respect to the above-mentioned Riemannian metric is computed as follows:

Ric

G(N, k)

P(X, X)=

1ik, k+1jN

[X, Eij]2

H S+[X, Fij]2

H S

, X∈g(N, k).

A simple direct computation shows that the above right-hand side isNX2H Sso that Ric

G(N, k)

=N I2k(Nk). (2.2)

Proof of Theorem 2.1. Letα,βand(ν,{αij}1i,j=0)be as in Section 1.1 for the given pair(p, q). Since the inequality trivially holds ifν=0, assumeν=0 and letν1:=ν(1)1ν, the normalization ofν. In addition to the assumptions of Lemma 1.2 we first assume the following (A) and (B):

(A) νis supported in[δ,1−δ]for a sufficiently smallδ >0 and it has the continuous density dν/dx.

(B) The functionQν1(x):=21

0 log|xy|dν1(y)is a well-definedC1-function on[0,1]. ChooseC1-functionsh0(x)andh1(x)on[0,1]such that

h0(x)

=logx (δ≤x≤1),

≥logx (0≤xδ), h1(x)

=log(1−x) (0≤x≤1−δ),

≥log(1−x) (1−δx≤1).

For eachN∈Nchoosek(N ), l(N )∈ {1, . . . , N−1}such thatk(N )/Nαandl(N )/NβasN→ ∞, and set n0(N ):=N−min

k(N ), l(N ) , n1(N ):=max

k(N )+l(N )N,0 , n(N ):=min

k(N ), l(N ), Nk(N ), Nl(N )

=Nn0(N )n1(N ).

Letting

ψN(x):=n(N )

N Qν1(x)+|k(N )l(N )|

N h0(x)+|k(N )+l(N )N|

N h1(x), 0≤x≤1,

we define a probability measure (regarded as a pair ofN×N random projection matrices)λψNN onG(N, k(N ))× G(N, l(N ))by

ψNN(P , Q):= 1 ZNψN exp

NTrN

ψN(P QP )

0N(P , Q) (2.3)

with the normalization constant ZNψN and the reference measure λ0N :=γG(N ,k(N ))γG(N ,l(N )). When (P , Q)G(N, k(N ))×G(N, l(N )) is distributed under λ0N, the joint eigenvalue distribution of P QP is known due to Ol’shanskij [13] (see also [5]). In the formulation of [8], Eq. (2.1), the eigenvalues ofP QP are

0, . . . ,0

n0(N )times

, 1, . . . ,1

n1(N )times

, x1, . . . , xn(N ) (2.4)

(6)

and the joint distribution of(x1, . . . , xn(N ))∈ [0,1]n(N )is d˜λ0N(x1, . . . , xn(N )):= 1

ZN0

n(N )

i=1

xi|k(N )l(N )|(1xi)|k(N )+l(N )N|

1i<jn(N )

(xixj)2

n(N )

i=1

dxi (2.5)

with the normalization constantZ0N. Hence it turns out that when(P , Q)G(N, k(N ))×G(N, l(N ))is distributed underλψNN, the eigenvalues ofP QP are listed as in (2.4) but the joint distribution of(x1, . . . , xn(N ))∈ [0,1]n(N ) is changed to

λ˜ψNN(x1, . . . , xn(N )):= 1 ZNψN exp

− ˜ψN(x1, . . . , xn(N ))

1i<jn(N )

(xixj)2

n(N )

i=1

dxi (2.6)

with the new normalization constantZNψN, where ψ˜N(x1, . . . , xn(N )):=

n(N )

i=1

n(N )Qν1(xi)+k(N )l(N )h0(xi)−logxi +k(N )+l(N )Nh1(xi)−log(1−xi)

.

Similarly to [8], Proposition 2.1 and [7], Section 5.5, we have:

(a) The limitC:=limN→∞N12logZNψN exists as well asC=limN→∞N12logZN0 (see (1.2)).

(b) When(x1, . . . , xn(N ))is distributed underλ˜ψNN, the empirical measuren(N )1 n(N )

i=1 δxi satisfies the large deviation principle in the scale 1/N2with the rate function

I (μ):= −ρ2(μ)+ρ2 1

0

F (x)dμ(x)+C forμM [0,1]

,

whereM([0,1])is the set of probability measures on[0,1],ρis given in (1.1) and F (x):=Qν1(x)+|αβ|

ρ

h0(x)−logx

+|α+β−1|

ρ

h1(x)−log(1−x)

for 0≤x≤1.

(c) ν1is a unique minimizer ofI withI (ν1)=0.

The last assertion follows from [14], Theorems I.1.3 and I.3.1, because by the construction ofh0andh1we get Qν1(x)

=F (x) ifx∈ [δ,1−δ] ⊃suppν1,

F (x) forx∈ [0,1].

Furthermore the above large deviation yields:

(d) The mean eigenvalue distributionλˆψNN:=

[0,1]n(N ) 1 n(N )

n(N )

i=1 δxidλ˜ψNN(x1, . . . , xn(N ))weakly converges toν1as N→ ∞.

Since the Riemannian manifoldG(N, k(N ))×G(N, l(N )) has the volume measureλ0N and its Ricci curvature tensor isN I2k(Nk)+2l(Nl)by (2.2), the classical logarithmic Sobolev inequality due to Bakry and Emery [1] implies that

S

λψNN, λ0N

≤ 1 2N

G(N ,k(N ))×G(N ,l(N ))

∇logdλψNN0N

2

H S

ψNN, (2.7)

(7)

where the left-hand side is the relative entropy of λψNN with respect to λ0N and the gradient

∇log(dλψNN/dλ0N)(P , Q)is considered ing(N, k(N ))⊕g(N, l(N ))via the natural identificationT(P ,Q)G(N, k(N ))× G(N, l(N ))=g(N, k(N ))⊕g(N, l(N )). By (2.5) and (2.6) notice that

ψNN

0N (P , Q)= 1 ZψNN exp

NTrN

ψN(P QP )

= Z0N ZNψN exp

N

n(N )

i=1

ψN(xi)

for (P , Q)G(N, k(N ))×G(N, l(N )) and for the eigenvalues(x1, . . . , xn(N ))of P QP exceptn0(N )zeros and n1(N )ones (see (2.4)). Hence we get

S

λψNN, λ0N

=

G(N ,k(N ))×G(N ,l(N ))

logdλψNN

0N (P , Q)ψNN(P , Q)

=logZN0 −logZψNNN n(N ) 1

0

ψN(x)dλˆψNN(x).

SinceψN(x)converges toρQν1(x)+ |α−β|h0(x)+ |α+β−1|h1(x)uniformly on[0,1], it follows from (a) and (d) above that

Nlim→∞

1 N2S

λψNN, λ0N

=CCρ 1

0

ρQν1(x)+ |αβ|h0(x)+ |α+β−1|h1(x)1(x).

Since (c) gives

−C= −ρ21)+ρ2 1

0

F (x)1(x)= −ρ21)+ρ2 1

0

Qν1(x)1(x), we have

Nlim→∞

1 N2S

λψNN, λ0N

= −χproj(p, q) (2.8)

thanks to Lemma 1.1 and (1.3) together withν1=(2ρ)1ν(see [8], Eq. (3.4)).

On the other hand, since∇log(dλψNN/dλ0N)(P , Q)= −N(TrNN(P QP ))), one can compute ∇logdλψNN

0N (P , Q) 2

H S

=4N2TrN

ψN (P QP )2

P QP (IP QP ) ,

whose proof will be given as Lemma 2.2 below for completeness. Therefore,

G(N ,k(N ))×G(N ,l(N ))

∇logdλψNN

0N (P , Q) 2

H S

ψNN(P , Q)

=4N2

[0,1]n(N ) n(N )

i=1

ψN (xi)2

xi(1xi)dλ˜ψNN(x1, . . . , xn(N ))

=4N2n(N ) 1

0

ψN (x)2

x(1x)dλˆψNN(x)

=4n(N ) 1

0

n(N )Qν1(x)+k(N )l(N )h0(x)+k(N )+l(N )−1h1(x)2

x(1x)λψNN(x),

(8)

and thus by (d) we have

Nlim→∞

1 2N3

G(N ,k(N ))×G(N ,l(N ))

∇logdλψNN0N

2

H S

ψNN

=2ρ 1

0

ρQν

1(x)+ |αβ|h0(x)+ |α+β−1|h1(x)2

x(1x)1(x)

= 1

0

ρQν

1(x)+|αβ|

x −|α+β−1|

1−x 2

x(1x)dν(x)

=ϕ(p:q) (2.9)

by Lemma 1.2, sinceν1=(2ρ)1ν so thatρQν

1(x)=(Hf )(x)withf :=dν/dx. Combining (2.7)–(2.9) yields the desired inequality under (A) and (B).

Next let us remove (A) and (B). First suppose that (A) is still satisfied but (B) is not. For eachε >0 choose a non-negativeC-functionψε supported in[−ε, ε]with

ψε(x)dx =1. Letfε :=fψε for f :=dν/dx and define dνε(x):=fε(x)dx; thenνε is a measure supported in a closed proper subinterval of(0,1)withνε((0,1))= 1−1

i,j=0αij wheneverεis small enough. Let(pε, qε)be a pair of projections in some(M, τ )corresponding to the representing dataε,{αij}1i,j=0). (Such a pair can be constructed via the GNS representation associated with the tracial state determined by ε,{αij}1i,j=0); see [8], Section 3.) Since (A) and (B) are satisfied forνε, we get

χproj(pε, qε)ϕ(pε:qε). Sincefεfw,3→0 andHfεHfw,3→0 asε0 (see the proof of Lemma 1.2 for the weighted norm · w,3), the Hölder inequality together with (1.5) implies that

1 0

(Hfε)(x)2

fε(x)x(1x)dx−→

1 0

(Hf )(x)2

f (x)x(1x)dx (2.10)

and hence limε0ϕ(pε:qε)=ϕ(p:q). Since(μ)is weakly upper semicontinuous inμ(see, e.g. [7], Proposi- tion 5.3.2), we also have−χproj(p, q)≤lim infε0(−χproj(pε, qε))so that−χproj(p, q)ϕ(p:q).

Finally suppose only the assumptions stated in Lemma 1.2. Forδ >0 set dνδ(s):=1−1

i,j=0αij

ν([δ,1−δ]) 1[δ,1δ](x)dν(x)

and let(pδ, qδ)be a pair of projections corresponding toδ,{αij}1i,j=0). Let us denote the density ofνδ byfδ; then it is immediate to see thatfδfw,3→0. To show thatϕ(pδ:qδ)ϕ(p:q)asδ0, it suffices to prove the following convergences asδ0:

1 0

(Hfδ)(x)2

fδ(x)x(1x)dx−→

1 0

(Hf )(x)2

f (x)x(1x)dx, (2.11)

1

0

(Hfδ)(x)x1fδ(x) x(1x)dx−→

1

0

(Hf )(x)x1f (x) x(1x)dx, (2.12)

1

0

(Hfδ)(x)(1x)1fδ(x) x(1x)dx−→

1

0

(Hf )(x)(1x)1f (x) x(1x)dx, (2.13) 1

0

x2fδ(x) x(1x)dx−→

1

0

x2f (x) x(1x)dx, (2.14)

1 0

(1x)2fδ(x) x(1x)dx−→

1 0

(1x)2f (x) x(1x)dx. (2.15)

Remark here that (2.12) and (2.14) are unnecessary whenα01+α10= |αβ| =0, and so are (2.13) and (2.15) when α11+α00= |α+β−1| =0. The convergence (2.11) is verified as (2.10) above. Also, (2.14) and (2.15) immediately

(9)

follow from the hypothesis (1.4). Since (2.12) and (2.13) are similarly shown, let us prove only the former here. Thus we should assumeα=β, and (1.4) means1

0 x1f (x)dx <+∞. By the Hölder inequality together with (1.5) one can estimate

(Hfδ)x1fδ(Hf )x1f

w,1Cwfδfw,3·x1fδ1/2

w,2· fδ1/2w,3 +Cwfw,3·x1|fδf|1/2

w,2· fδf1/2w,3. Note that

x1fδ1/22

w,21

0

x1fδ(x)dx−→

1 0

x1f (x)dx, x1|fδf|1/22

w,21

0

x1fδ(x)f (x)dx−→0 as δ0, since1

0x1f (x)dx <+∞. These apparently imply (2.12) thanks to fL3((0,1), x(1−x)dx) and fδfw,3−→0.

Moreover, since−logx < x1near 0 and−log(1−x) < (1x)1near 1, the hypothesis (1.4) implies that 1

0

(−logx)fδ(x)dx−→

1

0

(−logx)f (x)dx <+∞, 1

0

−log(1−x)

fδ(x)dx−→

1 0

−log(1−x)

f (x)dx <+∞

as δ0 (whenever those are needed) so that−χproj(p, q)≤lim infδ0(−χproj(pδ, qδ)). Hence the proof is com-

pleted.

Lemma 2.2. Let ψ be aC1-function on [0,1] and define Ψ (P , Q):=TrN(ψ (P QP )) for (P , Q)G(N, k)× G(N, l).Then

Ψ (P , Q)2

H S=4TrN

ψ(P QP )2

P QP (IP QP ) holds for every(P , Q)G(N, k)×G(N, l).

Proof. Write(Xr)2k(Nr=1k)for the orthonormal basis ofg(N, k)given in (2.1) and also(Ys)2l(Ns=1l)for that ofg(N, l).

For each(P , Q)=(U PN(k)U, V PN(l)V)inG(N, k)×G(N, l), a local normal coordinate at(P , Q)is given by the mapping

(X, Y )∈g(N, k)⊕g(N, l)→

U eXPN(k)e−XU, V eYPN(l)e−YV

G(N, k)×G(N, l).

By a direct computation using this coordinate, one can compute

Ψ (P , Q)=

r

UQPf(P QP )P UUPf(P QP )P QU, Xr Xr

+

s

VPf(P QP )P QVVQPf(P QP )P V , Ys Ys

so that

Ψ (P , Q)2

H S=2Pk(N )UPf(P QP )P QU

IPk(N )2

H S

+2Pl(N )VQPf(P QP )P V

IPl(N )2

H S

=4TrN

f(P QP )2

P QP (IP QP )

.

(10)

3. Remarks

3.1. Classical vs.free probabilistic mutual information

The classical mutual information of two random variables, sayX, Y, is usually formulated to be I (X;Y ):=S(μ(X,Y ), μXμY)=

X×Xlog dμ(X,Y )

d(μXμY)(x, y)(X,Y )(x, y),

whereμX, μY are the distribution measures ofX, Y on the phase spaceX andμ(X,Y )the joint distribution of(X, Y ) onX ×X. The mutual information is in turn written as I (X;Y )=H (X)+H (Y )H (X, Y ) as long as all the Shannon–Gibbs entropiesH (X),H (Y )andH (X, Y )are finite. This is nothing but what Voiculescu mentioned in [20] as an initial motivation of his introduction of the liberation theory. Let us now apply the definition ofI (X;Y )to our random matrix model(P (N ), Q(N ))of a given pair(p, q)of projections, and then the proof of Theorem 2.1 (see (2.8)) shows that

χproj(p, q)= lim

N→∞

1 N2S

λψNN, λ0N

= lim

N→∞

1 N2I

P (N );Q(N ) .

3.2. −χproj=ifor two projections

We simply writei(p:q)for thefree mutual informationi(Cp+C(1p):Cq+C(1q))introduced in [20], and its “heuristic definition” should be “χproj(p)+χproj(q)χproj(p, q)” on the analogy of classical theory. However the actual definition is completely different based on the so-called liberation process so that in view ofχproj(p)= χproj(q)=0 it may be particularly interesting to examine whetheri(p:q)coincides with−χproj(p, q)or not. In fact, the inequality in Theorem 2.1 is a kind of logarithmic Sobolev inequality and its right-hand side (the Dirichlet form part) is a “derivative” ofi(p:q), which also gives us a strong reason for that question. The equalityχproj(p, q)= i(p:q) is technically similar to χ (X)=χ(X)in the single variable case shown in [19], though the former is much more involved. (Here it should be noted thatχχholds in general [4].) At the moment we can give only a heuristic argument for the question. Lett,{αij(t )}1i,j=0)be the representing data of the liberation process(p(t ):=

u(t )pu(t ), q)started at(p, q)with a free unitary Brownian motion{u(t )}t0(see [2]) freely independent of(p, q).

Remark that theαij(t )’s are constant intdue to [20], Lemma 12.5. Assume thatνt satisfies the same properties as in Lemma 1.2 for everyt≥0. By [20], Corollary 5.7, one has

τ

p(t+ε)qp(t+ε)m

=τ

p(t )qp(t )m +

2 τ

Jt, p(t )

qp(t )qm1 +O

ε2

with the liberation gradientJt:=j (Cp(t )+C(1p(t )):Cq+C(1q)), and hence letting dνt(x)=ft(x)dx one can derive

∂tft(x)= −

∂x x(1x)ft(x)

(Hft)(x)+α01+α10

xα00+α11 1−x

(3.1) under the additional assumption thatft(x)is smooth in(t, x). The differential formuladtdχproj(p(t ), q)=ϕ(p(t ):q) shows up after a rather heuristic computation by using (3.1). Then one can show limt→∞χproj(p(t ), q)=0 by using Theorem 2.1 and [20], Proposition 10.11.c, so that the desired−χproj(p, q)=i(p:q)follows. The insufficient points in the derivation outlined above are: (i) we need to prove that allνt,t >0, automatically satisfy the same properties as in Lemma 1.2 wheneverν0is assumed to satisfy; (ii) we need to prove thatft(x)is smooth in(t, x); and finally (iii) we have to give a rigorous derivation ofdtdχproj(p(t ), q)=ϕ(p(t ):q)from (3.1).

Acknowledgment

The authors thank Professor Masaki Izumi for fruitful discussions. They also thank the referees for comments.

(11)

References

[1] D. Bakry and M. Emery. Diffusion hypercontractives.Séminaire Probabilités XIX177–206. Lecture Notes in Math.1123. Springer, Berlin, 1985. MR0889476

[2] P. Biane. Free Brownian motion, free stochastic calculus and random matrices. InFree Probability Theory1–19. D. V. Voiculescu (Ed.).

Fields Inst. Commun.12. Amer. Math. Soc. Providence, RI, 1997. MR1426833

[3] P. Biane. Logarithmic Sobolev inequalities, matrix models and free entropy.Acta Math. Sinica19(2003) 497–506. MR2014030

[4] P. Biane, M. Capitaine and A. Guionnet. Large deviation bounds for matrix Brownian motion. Invent. Math.152 (2003) 433–459.

MR1975007

[5] B. Collins. Product of random projections, Jacobi ensembles and universality problems arising from free probability.Probab. Theory Related Fields133(2005) 315–344. MR2198015

[6] S. Gallot, D. Hulin and J. Lafontaine.Riemannian Geometry, 2nd edition. Universitext, Springer, Berlin, 1990. MR1083149 [7] F. Hiai and D. Petz.The Semicircle Law, Free Random Variables and Entropy. Amer. Math. Soc., Providence, RI, 2000. MR1746976 [8] F. Hiai and D. Petz. Large deviations for functions of two random projection matrices.Acta Sci. Math. (Szeged)72(2006) 581–609.

MR2289756

[9] F. Hiai, D. Petz and Y. Ueda. Free logarithmic Sobolev inequality on the unit circle.Canad. Math. Bull.49(2006) 389–406. MR2252261 [10] F. Hiai and Y. Ueda. Notes on microstate free entropy of projections.Publ. Res. Inst. Math. Sci.44(2008), 49–89.

[11] R. Hunt, B. Muckenhoupt and R. Wheeden. Weighted norm inequalities for the conjugate function and Hilbert transform.Trans. Amer. Math.

Soc.176(1973) 227–251. MR0312139

[12] M. Ledoux. A (one-dimensional) free Brunn–Minkowski inequality.C. R. Math. Acad. Sci. Paris340(2005) 301–304. MR2121895 [13] G. I. Ol’shanskij. Unitary representations of infinite dimensional pairs(g, k)and the formalism of R. Howe. InRepresentation of Lie Groups

and Related Topics269–463. A. M. Vershik and D. P. Zhelobenko (Eds).Adv. Stud. Contemp. Math.7. Gordon and Breach, New York, 1990.

MR1104279

[14] E. B. Saff and V. Totik.Logarithmic Potentials with External Fields. Springer, Berlin, 1997. MR1485778 [15] C. Villani.Topics in Optimal Transportation. Amer. Math. Soc., Providence, RI, 2003. MR1964483

[16] D. Voiculescu. The analogues of entropy and of Fisher’s information measure in free probability theory, I.Comm. Math. Phys.155(1993) 71–92. MR1228526

[17] D. Voiculescu. The analogues of entropy and of Fisher’s information measure in free probability theory, II.Invent. Math.118(1994) 411–440.

MR1296352

[18] D. Voiculescu. The analogues of entropy and of Fisher’s information measure in free probability theory, IV: Maximum entropy and freeness.

InFree Probability Theory293–302. D. V. Voiculescu (Ed.). Fields Inst. Commun.12. Amer. Math. Soc., Providence, RI, 1997. MR1426847 [19] D. Voiculescu. The analogues of entropy and of Fisher’s information measure in free probability theory, V: Noncommutative Hilbert trans-

forms.Invent. Math.132(1998) 189–227. MR1618636

[20] D. Voiculescu. The analogue of entropy and of Fisher’s information measure in free probability theory VI: Liberation and mutual free information.Adv. Math.146(1999) 101–166. MR1711843

[21] D. V. Voiculescu, K. J. Dykema and A. Nica.Free Random Variables. Amer. Math. Soc., Providence, RI, 1992. MR1217253

Références

Documents relatifs

We study global existence, uniqueness and positivity of weak solutions of a class of reaction-diffusion systems of chemical kinetics type, under the assumptions of log- arithmic

In Section 3, we present the proof of Theorem 1.2, and as an application, we explain how to use the parabolic Kozono-Taniuchi inequality in order to prove the long time existence

While the Wasserstein distance is a bona fide distance between probability measures, in the context of the log-Sobolev inequality it seems more natural to work with lower bounds

Logarithmic Sobolev inequality, Entropy, Fisher Information, Transport Distance, Gaussian measures.. The first author was partially supported by NSF

– We present a finite dimensional version of the logarith- mic Sobolev inequality for heat kernel measures of non-negatively curved diffusion operators that contains and improves

Motivated by recent work of Au, C´ ebron, Dahlqvist, Gabriel, and Male, we study regularity properties of the distribution of a sum of two selfad- joint random variables in a

Free entropy dimension and the orthogonal free quantum groups..

In the general case, inequality proved in this article, we do not know if the modified logarithmic Sobolev inequality (2.9) implies transportation inequality (2.16).... Using Theorem