20.07.2013 Views

Estimation optimale du gradient du semi-groupe de la chaleur sur le ...

Estimation optimale du gradient du semi-groupe de la chaleur sur le ...

Estimation optimale du gradient du semi-groupe de la chaleur sur le ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Journal of Functional Analysis 236 (2006) 369–394<br />

www.elsevier.com/locate/jfa<br />

<strong>Estimation</strong> <strong>optima<strong>le</strong></strong> <strong>du</strong> <strong>gradient</strong> <strong>du</strong> <strong>semi</strong>-<strong>groupe</strong><br />

<strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> <strong>le</strong> <strong>groupe</strong> <strong>de</strong> Heisenberg<br />

Hong-Quan Li ∗<br />

SFB 611, Institut für Angewandte Mathematik, Universität Bonn, Poppelsdorfer Al<strong>le</strong>e 82, D-53115 Bonn, Germany<br />

Reçu <strong>le</strong> 6 juil<strong>le</strong>t 2005 ; accepté <strong>le</strong> 21 février 2006<br />

Disponib<strong>le</strong> <strong>sur</strong> Internet <strong>le</strong> 18 avril 2006<br />

Communiqué par G. Pisier<br />

Résumé<br />

En utilisant l’inégalité <strong>de</strong> Poincaré et <strong>la</strong> formu<strong>le</strong> <strong>de</strong> représentation, on montre que <strong>sur</strong> <strong>le</strong> <strong>groupe</strong> <strong>de</strong> Heisenberg<br />

<strong>de</strong> dimension réel<strong>le</strong> 3, H1 , il existe une constante C>0 tel<strong>le</strong> que :<br />

<br />

∇e t f (g) Ce t |∇f | (g), ∀g ∈ H 1 ,t>0, f∈ C ∞ o<br />

H 1 .<br />

Ce résultat répond par l’affirmation à <strong>la</strong> question ouverte <strong>de</strong> [B.K. Driver, T. Melcher, Hypoelliptic heat<br />

kernel inequalities on the Heisenberg group, J. Funct. Anal. 221 (2005) 340–365]. Aussi, <strong>le</strong> résultat principal<br />

<strong>de</strong> [B.K. Driver, T. Melcher, Hypoelliptic heat kernel inequalities on the Heisenberg group, J. Funct. Anal.<br />

221 (2005) 340–365] peut être considéré comme une conséquence immédiate <strong>de</strong> l’inégalité précé<strong>de</strong>nte.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Mots-clés : Groupe <strong>de</strong> Heisenberg ; Semi-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> ; Inégalité <strong>de</strong> Poincaré ; Formu<strong>le</strong> <strong>de</strong> représentation ;<br />

Noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> ; Structure sous-riemannienne<br />

1. Intro<strong>du</strong>ction<br />

Sur R n , notons ∇ <strong>le</strong> <strong>gradient</strong>, et e t (t>0) <strong>le</strong> <strong>semi</strong>-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong>. On voit que ∇ et<br />

e t commutent, i.e.<br />

* Nouvel<strong>le</strong> adresse : Department of Mathematics, Fudan University, 220 Handan Road, Shanghai 200433, Peop<strong>le</strong>’s<br />

Republic of China.<br />

Adresse e-mail : li-hq@wiener.iam.uni-bonn.<strong>de</strong>.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.02.016


370 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

∇e t f(g)= e t (∇f )(g), ∀g ∈ R n ,t>0, f∈ C ∞ o<br />

R n .<br />

En général, cette propriété ne reste plus va<strong>la</strong>b<strong>le</strong> dans <strong>le</strong> cadre <strong>de</strong>s variétés riemanniennes complètes,<br />

(M, gM). Cependant, si <strong>la</strong> courbure <strong>de</strong> Ricci <strong>sur</strong> M, Ric(M), est minorée par k (k ∈ R),<br />

alors D. Bakry a montré dans [4] que :<br />

<br />

∇e t f(g) e −kt e t |∇f | (g), ∀g ∈ M, t > 0, f∈ C ∞ o (M). (1.1)<br />

Lorsque <strong>la</strong> courbure <strong>de</strong> Ricci <strong>sur</strong> M est minorée, M est stochastiquement complète, i.e.<br />

et1 = 1 pour tout t>0. Une conséquence immédiate <strong>de</strong> (1.1) est que pour tout 1 w 0, f∈ Co (M), (1.2)<br />

ou bien, plus faib<strong>le</strong>ment,<br />

<br />

∇e t f <br />

L∞ e −kt ∇f L∞, ∀t >0, f∈ C∞ o (M). (1.3)<br />

Les estimations <strong>de</strong> type (1.2) ou bien sous une forme affaiblie (ou plus forte) ont beaucoup<br />

d’applications, par exemp<strong>le</strong>, dans l’étu<strong>de</strong> <strong>de</strong>s inégalités <strong>de</strong> Poincaré et <strong>de</strong> Sobo<strong>le</strong>v logarithmique<br />

pour <strong>le</strong> <strong>semi</strong>-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong>, dans l’étu<strong>de</strong> <strong>du</strong> critère Γ2 et aussi dans l’étu<strong>de</strong> <strong>de</strong>s transformées<br />

<strong>de</strong> Riesz, voir par exemp<strong>le</strong> [1–3,7] ainsi que <strong>le</strong>urs références.<br />

Récemment, von Renesse et Sturm ont montré que dans <strong>le</strong> cadre <strong>de</strong>s variétés riemanniennes<br />

complètes, l’estimation (1.3) avec k ∈ R implique que Ric(M) k, voir [24]. Pour d’autres<br />

résultats équiva<strong>le</strong>nts à l’estimation (1.1), voir par exemp<strong>le</strong> [1,24] ainsi que <strong>le</strong>urs références.<br />

Et dans [17] (voir aussi [18]), on a montré que <strong>le</strong>s estimations (1.3) ou bien (1.2) ne restent<br />

plus va<strong>la</strong>b<strong>le</strong>s <strong>sur</strong> <strong>le</strong>s variétés coniques. Plus précisement, <strong>sur</strong> <strong>le</strong> cône C(Sϑ) avec ϑ>2π, qui est<br />

une variété riemannienne à courbure sectionnel<strong>le</strong> nul<strong>le</strong> mais pas complète, il existe une fonction<br />

lisse à support compact définie <strong>sur</strong> C(Sϑ), fo, tel<strong>le</strong> que ∇etfoL∞ =+∞pour tout t>0.<br />

Dans [9], Driver et Melcher ont étudié <strong>le</strong>s estimations <strong>de</strong> type (1.2) dans <strong>le</strong> cadre <strong>du</strong> <strong>groupe</strong><br />

<strong>de</strong> Heisenberg <strong>de</strong> dimension réel<strong>le</strong> 3, H1 , qui peut être considéré comme une variété munie d’un<br />

Lap<strong>la</strong>cien dégénéré.<br />

Rappelons, voir par exemp<strong>le</strong> [12, p. 98], que H1 = R2 × R est un <strong>groupe</strong> <strong>de</strong> Lie stratifié pour<br />

<strong>la</strong> loi<br />

(x1,x2,t)· (x ′ 1 ,x′ 2 ,t′ ) = x1 + x ′ 1 ,x2 + x ′ 2 ,t1 + t ′ 1 + 2(x′ 1 x2 − x1x ′ 2 ) .<br />

Et <strong>le</strong> sous-Lap<strong>la</strong>cien <strong>sur</strong> H 1 s’écrit comme<br />

= X 2 1 + X2 2 ,<br />

où <strong>le</strong>s <strong>de</strong>ux champs <strong>de</strong> vecteurs invariants à gauche <strong>sur</strong> H 1 ,X1 et X2, sont définis par<br />

X1 = ∂ ∂<br />

+ 2x2<br />

∂x1 ∂t , X2 = ∂ ∂<br />

− 2x1<br />

∂x2 ∂t .<br />

Notons ∇=(X1, X2) <strong>le</strong> <strong>gradient</strong>, et e t (t>0) <strong>le</strong> <strong>semi</strong>-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> H 1 .


H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 371<br />

Par une métho<strong>de</strong> probabiliste, Driver et Melcher ont montré que pour tout 1 1) tel<strong>le</strong> que :<br />

<br />

∇e t f <br />

(g)<br />

t<br />

Cw e |∇f | w (g) 1/w 1 ∞ 1<br />

, ∀g ∈ H ,t>0, f∈ Co H . (1.4)<br />

Ils ont montré aussi que <strong>le</strong>ur métho<strong>de</strong> ne marche pas pour <strong>le</strong> cas où w = 1, qui reste ouvert.<br />

De plus, Melcher a donné dans [23] quelques famil<strong>le</strong>s <strong>de</strong>s fonctions <strong>sur</strong> <strong>le</strong>squel<strong>le</strong>s l’estimation<br />

précé<strong>de</strong>nte avec w = 1 est satisfaite.<br />

On remarque aussi que Lust-Piquard et Vil<strong>la</strong>ni ont montré dans [21] <strong>le</strong>s estimations (1.4)<br />

(avec 1 0, f∈ C ∞ o<br />

H 1 .<br />

Par <strong>le</strong> Théorème 1.1, on obtient immédiatement l’inégalité <strong>de</strong> Sobo<strong>le</strong>v logarithmique pour <strong>le</strong><br />

<strong>semi</strong>-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> comme suit :<br />

Corol<strong>la</strong>ire 1.2. Soit C1 > 1 comme dans <strong>le</strong> Théorème 1.1. Alors, pour tout t>0 et toute f ∈<br />

C ∞ o (H1 ),ona:<br />

et aussi<br />

e t f 2 log f 2 (g) − e t f 2 (g) log e t f 2 (g) C 2 1 tet |∇f | 2 (g), ∀g ∈ H 1 ,<br />

e t f 2 log f 2 (g) − e t f 2 (g) log e t f 2 (g) C 2 1t |∇etf 2 | 2 (g)<br />

etf 2 , ∀g ∈ H<br />

(g)<br />

1 .<br />

Pour montrer ce corol<strong>la</strong>ire, il suffit <strong>de</strong> modifier un peu <strong>la</strong> preuve <strong>de</strong> <strong>la</strong> partie (i) ⇒ (ii) <strong>du</strong><br />

Théorème 5.4.7 <strong>de</strong> [1, pp. 85–86].<br />

L’idée <strong>de</strong> <strong>la</strong> démonstration <strong>du</strong> Théorème 1.1 est d’utiliser l’inégalité <strong>de</strong> Poincaré et <strong>la</strong> formu<strong>le</strong><br />

<strong>de</strong> représentation qu’on rappel<strong>le</strong> dans <strong>la</strong> Section 2 (<strong>le</strong>s <strong>le</strong>cteurs attentifs trouveront qu’en fait, <strong>la</strong><br />

propriété <strong>du</strong> doub<strong>le</strong>ment <strong>du</strong> volume loca<strong>le</strong>ment et <strong>la</strong> formu<strong>le</strong> <strong>de</strong> représentation loca<strong>le</strong>ment sont<br />

suffisantes). Cette idée provient d’une observation dans <strong>le</strong> cadre <strong>de</strong>s espaces euclidiens et c’est<br />

très faci<strong>le</strong> à comprendre. Cependant, <strong>la</strong> démonstration <strong>du</strong> Théorème 1.1 est re<strong>la</strong>tivement diffici<strong>le</strong> :<br />

on doit utiliser <strong>la</strong> structure sous-riemannienne <strong>de</strong> H 1 (i.e. l’expression explicite <strong>de</strong> <strong>la</strong> distance <strong>de</strong><br />

Carnot–Carathéodory et <strong>de</strong> <strong>la</strong> géodésique) ainsi que <strong>le</strong>s estimations <strong>optima<strong>le</strong></strong>s <strong>du</strong> noyau <strong>de</strong> <strong>la</strong><br />

<strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son <strong>gradient</strong>, qu’on rappel<strong>le</strong> ou donne dans <strong>la</strong> Section 3.<br />

1.1. Quelques remarques <strong>sur</strong> notre métho<strong>de</strong><br />

1. En général, dans <strong>le</strong> cadre <strong>de</strong>s variétés riemanniennes complètes à courbure <strong>de</strong> Ricci minorée,<br />

<strong>le</strong>s estimations (1.2) n’offrent <strong>de</strong>s informations précises que pour t → 0 + , et el<strong>le</strong>s <strong>de</strong>viennent


372 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

très faib<strong>le</strong>s lorsque t →+∞. Dans certains cas, <strong>la</strong> métho<strong>de</strong> <strong>de</strong> cet artic<strong>le</strong> nous permet d’obtenir<br />

<strong>de</strong>s estimations <strong>optima<strong>le</strong></strong>s <strong>de</strong> type (1.2) lorsque t →+∞.<br />

Pour mieux comprendre <strong>le</strong>s avantages <strong>de</strong> notre métho<strong>de</strong> (qui consiste à utiliser directement<br />

<strong>de</strong>s estimations <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son <strong>gradient</strong>, plutôt que <strong>la</strong> formu<strong>le</strong> <strong>de</strong> Bochner-<br />

Lichnerowicz, comme [4]), considérons une variété riemannienne compacte, sans bord et <strong>de</strong><br />

dimension n, N.<br />

Notons pt <strong>le</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> N, λ1 > 0 <strong>la</strong> première va<strong>le</strong>ur propre non nul<strong>le</strong> <strong>du</strong> Lap<strong>la</strong>cien<br />

et |N| son volume. Alors pour t ≫ 1, on a d’une part :<br />

<br />

∇e t f(g) <br />

<br />

= <br />

∇ <br />

<br />

<br />

N<br />

N<br />

pt(g, g ′ <br />

) f(g ′ ) − 1<br />

|N|<br />

<br />

N<br />

N<br />

<br />

f(g∗)dμ(g∗) dμ(g ′ <br />

<br />

) <br />

<br />

<br />

∇pt(g, g ′ ) <br />

<br />

· <br />

f(g′ ) − 1<br />

<br />

<br />

<br />

f(g∗)dμ(g∗) <br />

|N|<br />

dμ(g′ )<br />

C1e −λ1t<br />

<br />

<br />

<br />

f(g′ ) − 1<br />

<br />

<br />

<br />

f(g∗)dμ(g∗) <br />

|N|<br />

dμ(g′ )<br />

N<br />

N<br />

par <strong>le</strong> fait que ∇pt(g, g ′ ) C1e−λ1t pour tout t ≫ 1etg,g ′ ∈ N<br />

C2e −λ1t<br />

<br />

diam(N) |∇f |(g ′ )dμ(g ′ )<br />

N<br />

par l’inégalité <strong>de</strong> Poincaré<br />

Ce −λ1t<br />

<br />

diam(N)|N| pt(g, g ′ )|∇f |(g ′ )dμ(g ′ )<br />

N<br />

par <strong>le</strong> fait que pt(g, g ′ ) 2 −1 |N| −1 pour tout t ≫ 1etg,g ′ ∈ N<br />

Ce −λ1t diam(N)|N| e t |∇f | w (g) 1/w ,<br />

pour tout g ∈ N, f ∈ C ∞ (N) et 1 we<br />

∇υ1Lw −1/w 2<br />

,<br />

|N|<br />

puisque pt(g, g ′ ) 2|N| −1 pour tout t ≫ 1etg,g ′ ∈ N.<br />

Autrement dit, dans <strong>le</strong> cadre <strong>de</strong>s variétés riemanniennes compactes, sans bord, on a<br />

sup<br />

g∈N,f∈C ∞ (N)<br />

|∇e t f(g)|<br />

[e t (|∇f | w )(g)] 1/w ∼ e−λ1t .


H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 373<br />

Et on remarque que lorsque t →+∞, <strong>le</strong> facteur e−λ1t est décroissant beaucoup plus vite que<br />

e− max(0,ko)t avec ko = sup{k ∈ R; Ric(N) k}, en rappe<strong>la</strong>nt que <strong>le</strong> théorème <strong>de</strong> Lichnerowicz<br />

nous dit que λ1 n<br />

n−1ko pour ko > 0.<br />

2. Dans [21,23], <strong>le</strong>s estimations (1.4) avec 1


374 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

où<br />

<br />

f(g ′ ) − fr(g) C<br />

<br />

B(g,r)<br />

<br />

∇f(g∗) <br />

d(g ′ ,g∗)<br />

|B(g ′ ,d(g ′ ,g∗))| dg∗, ∀g ′ ∈ B(g,r), (2.2)<br />

fr(g) = B(g,r) −1<br />

<br />

B(g,r)<br />

f(g ′ )dg ′ .<br />

On remarque aussi que <strong>sur</strong> H 1 , l’inégalité <strong>de</strong> Poincaré (2.1) et <strong>la</strong> formu<strong>le</strong> <strong>de</strong> représentation<br />

(2.2) sont équiva<strong>le</strong>ntes, voir [19] (ou bien voir [10,11] pour <strong>de</strong>s résultats un peu faib<strong>le</strong>s mais<br />

suffisants pour montrer notre résultat).<br />

3. Quelques résultats <strong>sur</strong> H 1<br />

3.1. Rappel <strong>sur</strong> <strong>la</strong> structure sous-riemannienne <strong>de</strong> H 1<br />

On rappel<strong>le</strong> dans cette partie l’expression explicite <strong>de</strong> <strong>la</strong> distance <strong>de</strong> Carnot–Carathéodory et<br />

<strong>de</strong> <strong>la</strong> géodésique minima<strong>le</strong> entre l’origine o et (x, t) ∈ H 1 . Presque tous <strong>le</strong>s résultats <strong>de</strong> cette<br />

partie proviennent <strong>de</strong> [5, pp. 635–644] et [12].<br />

Dans <strong>la</strong> suite, pour simplifier <strong>le</strong>s notations, on suppose que<br />

Posons<br />

Alors<br />

x = (x1,x2) = (0, 0).<br />

2ϕ − sin 2ϕ<br />

μ(ϕ) =<br />

2sin2 : ]−π,π[→R. (3.1)<br />

ϕ<br />

ψ(x,ϕ)= x,μ(ϕ)x 2 : (x, ϕ) ∈ R 2 ×]−π,π[; x = (0, 0) → (x, t) ∈ H 1 ; x = (0, 0) <br />

est un C 1 -difféomorphisme.<br />

Notons μ −1 et ψ −1 <strong>la</strong> fonction réciproque <strong>de</strong> μ et <strong>de</strong> ψ respectivement, alors<br />

d 2 2 θ<br />

(x, t) = x<br />

sin θ<br />

2<br />

avec θ = μ −1<br />

<br />

t<br />

x2 <br />

(voir (1.40) et (1.30) <strong>de</strong> [5], et on a posé a = τ = 1 et remp<strong>la</strong>cé 2θ par θ dans (1.30)), et <strong>la</strong><br />

géodésique minima<strong>le</strong> unique entre l’origine et (x, t),<br />

γ(s)= x1(s), x2(s), t(s) : [0, 1]→H 1 ,<br />

s’écrit comme (il suffit <strong>de</strong> poser a = τ = 1 et <strong>de</strong> remp<strong>la</strong>cer 2θ par θ dans <strong>le</strong> Theorem 1.21 <strong>de</strong> [5])<br />

(3.2)<br />

(3.3)


x(s) =<br />

H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 375<br />

<br />

x1(s)<br />

=<br />

x2(s)<br />

sin sθ<br />

<br />

cos(s − 1)θ sin(s − 1)θ<br />

×<br />

sin θ − sin(s − 1)θ cos(s − 1)θ<br />

x1<br />

x2<br />

<br />

, (3.4)<br />

θ<br />

t(s)= t − (1 − s)<br />

sin2 θ x2 + 1 sin 2θ − sin 2sθ<br />

2 sin2 x<br />

θ<br />

2<br />

= 2sθ − sin 2sθ<br />

2sin2 x<br />

θ<br />

2 2sθ − sin 2sθ<br />

=<br />

2sin2 <br />

x(s) 2 ,<br />

sθ<br />

autrement dit, on a<br />

(3.5)<br />

μ −1<br />

<br />

t(s)<br />

x(s)2 <br />

= sμ −1<br />

<br />

t<br />

x2 <br />

, ∀0 s 1. (3.6)<br />

3.2. <strong>Estimation</strong>s <strong>optima<strong>le</strong></strong>s <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son <strong>gradient</strong><br />

Par [12,14] ou [20], on a l’expression explicite <strong>de</strong> ph comme suit :<br />

ph(x, t) =<br />

1<br />

2(4πh) 2<br />

<br />

R<br />

<br />

λ 2<br />

exp tı −x coth λ<br />

4h<br />

λ<br />

dλ. (3.7)<br />

sinh λ<br />

Pour montrer <strong>le</strong> Théorème 1.1, on aura besoin <strong>de</strong>s estimations <strong>optima<strong>le</strong></strong>s <strong>de</strong> p(x,t) et <strong>de</strong><br />

|∇p(x,t)| comme suit :<br />

Lemme 3.1. Il existe une constante L1 > 1 tel<strong>le</strong> que pour tout (x, t) ∈ H 1 ,ona:<br />

L −1<br />

1 e−d2 (x,t)/4 1 +xd(x,t) −1/2 p(x,t) L1e −d2 (x,t)/4 1 +xd(x,t) −1/2 . (3.8)<br />

Lemme 3.2. Il existe une constante L2 > 1 tel<strong>le</strong> que :<br />

<br />

∇p(x,t) L2d(x,t)p(x,t), ∀(x, t) ∈ H 1 . (3.9)<br />

Preuve <strong>du</strong> Lemme 3.1. L’estimation (3.8) est une conséquence immédiate <strong>de</strong> l’expression explicite<br />

<strong>de</strong> <strong>la</strong> distance <strong>de</strong> Carnot–Carathéodory (3.3) et <strong>de</strong>s Theorems 1.1, 1.3 et Remark 1.5 <strong>de</strong><br />

[13] (voir aussi [5,12] pour <strong>de</strong>s résultats partiels). ✷<br />

Preuve <strong>du</strong> Lemme 3.2. Observons que<br />

et<br />

X1p(x,t) =<br />

1<br />

4(4π) 2<br />

<br />

<br />

∇p(x,t) 2 =|X1p| 2 (x, t) +|X2p| 2 (x, t),<br />

R<br />

X2p(x,t) =− 1<br />

4(4π) 2<br />

<br />

λ<br />

2<br />

λ(−x1 coth λ + x2ı)exp tı −x coth λ<br />

4<br />

λ<br />

sinh λ dλ,<br />

<br />

R<br />

<br />

λ<br />

2<br />

λ(x2 coth λ + x1ı)exp tı −x coth λ<br />

4<br />

λ<br />

sinh λ dλ.


376 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

Soient<br />

Alors<br />

<br />

W1 =<br />

R<br />

R<br />

<br />

λ<br />

2<br />

λ exp tı −x coth λ<br />

4<br />

λ<br />

sinh λ dλ,<br />

<br />

λ<br />

2<br />

W2 = cosh λ exp tı −x coth λ<br />

4<br />

2 λ<br />

dλ.<br />

sinh λ<br />

<br />

∇p(x,t) <br />

1<br />

<br />

4(4π) 2<br />

<br />

|x1|+|x2| |W1|+|W2| .<br />

Comme |x1|+|x2| 2x 2d(x,t) (voir (3.3)), par (3.8), pour terminer <strong>la</strong> preuve <strong>de</strong> (3.9),<br />

il nous reste à montrer qu’il existe une constante C>0 tel<strong>le</strong> que pour tout (x = (0, 0), t) ∈ H 1 ,<br />

ona:<br />

|W1| Ce − d2 (x,t) −1/2, 4 1 +xd(x,t) (3.10)<br />

|W2| C d(x,t)<br />

x e− d2 (x,t) −1/2. 4 1 +xd(x,t) (3.11)<br />

En comparant l’expression intégra<strong>le</strong> <strong>de</strong> W1 <br />

(resp. W2) avec cel<strong>le</strong> <strong>de</strong> p (resp. p∗(x =<br />

x2 1 + x2 2 + x2 3 + x2 4 ,t) = p∗1(x1,x2,x3,x4,t), <strong>le</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> en temps h = 1<strong>sur</strong><strong>le</strong><br />

<strong>groupe</strong> <strong>de</strong> Heisenberg <strong>de</strong> dimension réel<strong>le</strong> 5, H2 ), on trouve qu’il y a dans l’intégra<strong>le</strong> qu’une différence<br />

d’un facteur, <strong>la</strong> fonction analytique <strong>sur</strong> <strong>le</strong> p<strong>la</strong>n comp<strong>le</strong>xe, (2π) −2λ (resp. (2π) −3 cosh λ).<br />

On observe aussi que l’estimation W1 est analogue à cel<strong>le</strong> <strong>de</strong> p obtenue dans <strong>le</strong> Theorem 2.17 <strong>de</strong><br />

[5] et que l’estimation W2 est plus faib<strong>le</strong> que cel<strong>le</strong> <strong>de</strong> p∗1 obtenue dans <strong>le</strong> Theorem 4.6 <strong>de</strong> [5].<br />

Donc, il suffit <strong>de</strong> répéter mot par mot <strong>la</strong> preuve <strong>du</strong> Theorem 2.17 (resp. Theorem 4.6 avec n = 2)<br />

<strong>de</strong> [5] pour montrer (3.10) (resp. (3.11)). ✷<br />

4. Preuve <strong>du</strong> Théorème 1.1<br />

Dans <strong>la</strong> suite, V(r)=|B(o,r)| (r >0). On rappel<strong>le</strong> que<br />

<br />

B(g,r) = V(r)∼ r 4 , ∀g ∈ H 1 ,r>0. (4.1)<br />

Par l’invariance à gauche <strong>de</strong> ∇ et <strong>de</strong> ph ainsi que <strong>la</strong> di<strong>la</strong>tation <strong>sur</strong> H 1 , on remarque que pour<br />

montrer <strong>le</strong> Théorème 1.1, il suffit <strong>de</strong> montrer qu’il existe une constante C>1 tel<strong>le</strong> que (voir<br />

aussi Lemma 2.3 et Proposition 2.6 <strong>de</strong> [9] pour l’explication détaillée) :<br />

<br />

∇e f (o) Ce |∇f | (o), ∀f ∈ C ∞ o<br />

. (4.2)


H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 377<br />

Dans toute <strong>la</strong> suite, pour simplifier <strong>le</strong>s notations, on pose<br />

1<br />

fo =<br />

|B(o,1)|<br />

<br />

B(o,1)<br />

f(g)dg, A∞ = e 1000 (L1 + L2) 8 ,<br />

où L1,L2 > 1 proviennent <strong>de</strong> (3.8) et (3.9) respectivement.<br />

L’idée principa<strong>le</strong> <strong>de</strong> <strong>la</strong> preuve <strong>de</strong> (4.2). On explique brièvement l’idée principa<strong>le</strong> <strong>de</strong> <strong>la</strong> preuve<br />

<strong>de</strong> (4.2) :<br />

Comme H1 est stochastiquement complète, on a<br />

<br />

∇e f <br />

<br />

(o) = <br />

∇p g −1 <br />

<br />

f(g)−fo dg<br />

<br />

<br />

<br />

∇p g −1 · <br />

f(g)−fo dg.<br />

H 1<br />

On écrit formel<strong>le</strong>ment<br />

<br />

∇e f <br />

<br />

(o) ∇p g −1 · <br />

f(g)−fo dg<br />

H 1<br />

<br />

<br />

H 1<br />

<br />

=<br />

H 1<br />

<br />

∇p g −1 <br />

<br />

<br />

∇f(g ′ ) <br />

<br />

H 1<br />

H 1<br />

H 1<br />

<br />

C p(g ′ ) ∇f(g ′ ) dg ′ .<br />

H 1<br />

<br />

∇f(g ′ ) K(g,g ′ )dg ′<br />

<br />

dg<br />

<br />

∇p g −1 K(g,g ′ )dg<br />

Autrement dit, pour démontrer l’estimation (4.2), il suffit <strong>de</strong> trouver une fonction K définie<br />

<strong>sur</strong> H 1 × H 1 qui satisfait <strong>le</strong>s <strong>de</strong>ux conditions suivantes :<br />

<br />

f(g)− fo<br />

<br />

H 1<br />

<br />

<br />

<br />

H 1<br />

<br />

∇f(g ′ ) K(g,g ′ )dg ′ , ∀g ∈ H 1 ,f∈ C ∞ o<br />

<br />

dg ′<br />

H 1 , (4.3)<br />

<br />

∇p g −1 K(g,g ′ )dg Cp(g ′ ), ∀g ′ ∈ H 1 . ✷ (4.4)<br />

La condition (4.3) nous con<strong>du</strong>it naturel<strong>le</strong>ment à utiliser l’inégalité <strong>de</strong> Poincaré (loca<strong>le</strong>ment)<br />

et <strong>la</strong> formu<strong>le</strong> <strong>de</strong> représentation (loca<strong>le</strong>ment). La difficulté se trouve à remplir <strong>la</strong> condition (4.4) :<br />

on doit choisir K à partir <strong>de</strong> (4.3) <strong>le</strong> mieux possib<strong>le</strong>, et doit utiliser <strong>le</strong>s estimations <strong>optima<strong>le</strong></strong>s<br />

<strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son <strong>gradient</strong> ainsi que <strong>la</strong> structure sous-riemannienne <strong>de</strong> H 1 .On<br />

verra aussi que <strong>la</strong> difficulté pour contrô<strong>le</strong>r K(g,g ′ ) se trouve dans <strong>le</strong> cas où d(g,g ′ ) ≪ 1avec<br />

d(g) →+∞.


378 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

Preuve <strong>de</strong> (4.2). Soient<br />

on a<br />

Par <strong>le</strong> fait que<br />

Comme<br />

J1 =<br />

J2 =<br />

<br />

{g∈H 1 ;d(g)A∞}<br />

<br />

{g∈H 1 ;d(g)>A∞}<br />

d(g)p(g) · <br />

f(g)−fo dg,<br />

d(g)p(g) · <br />

f(g)−fo dg.<br />

<br />

∇p g −1 L2d g −1 p g −1 = L2d(g)p(g),<br />

e |∇f | <br />

(o) =<br />

<br />

∇e f (o) L2(J1 + J2).<br />

H 1<br />

p g −1 <br />

|∇f |(g) dg =<br />

H 1<br />

p(g)|∇f |(g) dg,<br />

pour démontrer (4.2), il nous reste à montrer qu’il existe une constante C>0, qui ne dépend pas<br />

<strong>de</strong> f , tel<strong>le</strong> que<br />

<br />

J1 + J2 C p(g)|∇f |(g) dg. ✷<br />

4.1. <strong>Estimation</strong> <strong>de</strong> J1<br />

On voit que<br />

<br />

f(g)− fo<br />

H 1<br />

<br />

<br />

f1(g) − fo<br />

+ f(g)−f1(g) .<br />

La formu<strong>le</strong> <strong>de</strong> représentation (voir (2.2)) nous dit que<br />

Par ail<strong>le</strong>urs,<br />

J11∗ = f(g)− f1(g) C1<br />

= C1<br />

<br />

B(g,1)<br />

<br />

B(g,1)<br />

<br />

∇f(g ′ ) <br />

d(g,g ′ )<br />

|B(g,d(g,g ′ ))| dg′<br />

<br />

∇f(g ′ ) <br />

d(g ′ ,g)<br />

|B(g ′ ,d(g ′ ,g))| dg′ , par (4.1).<br />

J12∗ = fo − f1(g) fo − fA∞+2(o) + f1(g) − fA∞+2(o) .


H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 379<br />

En utilisant l’inégalité <strong>de</strong> Poincaré (voir (2.1)), on a<br />

et <strong>de</strong> <strong>la</strong> même façon,<br />

<br />

fo − fA∞+2(o) =<br />

<br />

1 <br />

<br />

|B(o,1)| <br />

1<br />

<br />

|B(o,1)|<br />

C(A∞)<br />

<br />

B(o,1)<br />

<br />

B(o,A∞+2)<br />

<br />

B(o,A∞+2)<br />

<br />

f1(g) − fA∞+2(o) C(A∞)<br />

′<br />

f(g ) − fA∞+2(o) dg ′<br />

<br />

<br />

<br />

<br />

<br />

f(g ′ ) − fA∞+2(o) dg ′<br />

|∇f |(g ′ )dg ′ ,<br />

<br />

B(o,A∞+2)<br />

Par <strong>le</strong>s estimations supérieures <strong>de</strong> p, (3.8), on a donc<br />

J1 C ′ (A∞)<br />

C∗(A∞)<br />

C(A∞)<br />

C ∗ (A∞)<br />

<br />

B(o,A∞+1)<br />

<br />

B(o,A∞+2)<br />

<br />

B(o,A∞+2)<br />

<br />

B(o,A∞+2)<br />

<br />

B(g,1)<br />

<br />

∇f(g ′ ) <br />

d(g ′ ,g)<br />

|B(g ′ ,d(g ′ ,g))| dg′ +<br />

|∇f |(g ′ <br />

) 1 +<br />

|∇f |(g ′ )dg ′<br />

<br />

B(g ′ ,1)<br />

p(g ′ )|∇f |(g ′ )dg ′<br />

C ∗ <br />

(A∞) p(g ′ )|∇f |(g ′ )dg ′ .<br />

H 1<br />

4.2. <strong>Estimation</strong> <strong>de</strong> J2<br />

Puisque<br />

J2 =<br />

on peut supposer dans toute <strong>la</strong> suite<br />

<br />

par (4.1)<br />

{g=(x,t)∈H 1 ;d(g)>A∞,x=(0,0)}<br />

|∇f |(g ′ )dg ′ .<br />

d(g ′ ,g)<br />

|B(g ′ ,d(g ′ ,g))| dg<br />

<br />

par (3.8)<br />

<br />

B(o,A∞+2)<br />

dg ′<br />

d(g)p(g) <br />

f(g)−fo dg,<br />

g = (x, t) ∈ Σ = R 2 × R \ (0, 0,t); t ∈ R <br />

; d(g) > A∞ .<br />

|∇f |(g ′ )dg ′<br />

<br />

dg


380 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

Notons γ <strong>la</strong> géodésique minima<strong>le</strong> (unique) d’origine à g = (x, t), qui est définie par (3.4)<br />

et (3.5).<br />

Soit<br />

N(g)= 100d 3/2 (g) ln d(g) + 1,<br />

où [h] (h ∈ R) désigne <strong>la</strong> partie entière <strong>de</strong> h,etsoit<br />

g N(g) = γ 1 − N(g)d −7/2 (g) .<br />

On doit dire au <strong>le</strong>cteur que <strong>le</strong> choix <strong>de</strong> N(g) et {g(i) = γ(1− id−7/2 (g))}0id(g)N(g) défini<br />

dans <strong>la</strong> Section 4.2.2 n’est pas unique mais très délicat, ça dépend non seu<strong>le</strong>ment <strong>de</strong>s estimations<br />

<strong>optima<strong>le</strong></strong>s <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son <strong>gradient</strong> et aussi <strong>de</strong> <strong>la</strong> structure sous-riemannienne<br />

<strong>de</strong> H1 (plus précisment, on aura besoin <strong>de</strong>s résultats analogues aux Lemmes 4.2 et 4.3 <strong>de</strong> cet<br />

artic<strong>le</strong>).<br />

Alors<br />

<br />

f(g)−fo <br />

f2d−5/2 (g) g N(g) − fo<br />

+ <br />

f(g)−f2d−5/2 (g) g N(g) .<br />

En utilisant l’inégalité <strong>de</strong> Poincaré (voir (2.1)) et l’estimation <strong>du</strong> volume <strong>de</strong>s bou<strong>le</strong>s dans H1 (voir (4.1)), on a<br />

<br />

f2d−5/2 (g) g N(g) − fo<br />

fo − fd(g(N(g)))+4d−5/2 (g) (o) <br />

+ <br />

f2d−5/2 (g) g N(g) − fd(g(N(g)))+4d−5/2 (g) (o) <br />

Cd 11 <br />

<br />

(g)<br />

∇f(g ′ ) dg ′ .<br />

Donc,<br />

<br />

f(g)−fo 11<br />

Cd (g)<br />

Soient<br />

J21 =<br />

<br />

J22 =<br />

On constate que<br />

<br />

<br />

B o,d(g)−90<br />

{g∈H 1 ;d(g)>A∞}<br />

Σ<br />

ln d(g) <br />

d(g)<br />

B o,d(g)−90<br />

ln d(g) <br />

d(g)<br />

<br />

′<br />

∇f(g ) ′<br />

dg + f(g)−f2d−5/2 (g) g N(g) .<br />

d 12 <br />

(g)p(g)<br />

ln d(g)<br />

B(o,d(g)−90 d(g) )<br />

d(g)p(g) <br />

f(g)−f2d−5/2 (g) g N(g) dg.<br />

J2 C(J21 + J22).<br />

<br />

∇f(g ′ ) dg ′<br />

<br />

dg,


H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 381<br />

Pour achever <strong>la</strong> preuve <strong>du</strong> Théorème 1.1, il suffit <strong>de</strong> montrer qu’il existe une constante C>0<br />

tel<strong>le</strong> que pour toute f ∈ C ∞ o (H1 ),ona<br />

<br />

J21 + J22 C<br />

4.2.1. <strong>Estimation</strong> <strong>de</strong> J21<br />

Par un calcul simp<strong>le</strong>, on observe que<br />

H 1<br />

p(g) ∇f(g) dg.<br />

<br />

(g, g ′ ) ∈ H 1 × H 1 ; d(g) > A∞,d(g ′ <br />

ln d(g)<br />

) A∞,d(g ′ ) A∞ − 1 <br />

<br />

∪ (g, g ′ ) ∈ H 1 × H 1 ; d(g ′ )>A∞ − 1,d(g)>d(g ′ ) + 46 ln d(g′ )<br />

d(g ′ <br />

.<br />

)<br />

Donc, en utilisant <strong>le</strong>s estimations supérieures <strong>de</strong> p(g) (voir (3.8)), on a<br />

J21 C<br />

<br />

d(g ′ )A∞−1<br />

+ C<br />

<br />

d(g ′ )>A∞−1<br />

<br />

∇f(g ′ ) dg ′<br />

<br />

d(g)>A∞<br />

<br />

∇f(g ′ ) <br />

<br />

d 12 (g)e −d2 (g)/4 dg<br />

d(g)>d(g ′ )+46 ln d(g′ )<br />

d(g ′ )<br />

Par l’estimation <strong>du</strong> volume <strong>de</strong>s bou<strong>le</strong>s dans H 1 (voir (4.1)), on a<br />

<br />

d(g)>A∞<br />

et lorsque d(g ′ )>A∞ − 1, on a<br />

d 12 (g)e −d2 <br />

(g)/4<br />

dg dg ′ .<br />

d 12 (g)e −d2 (g)/4 14<br />

dg CA∞e −A2∞ /4 C 1 + A 2 −1/2e−(A∞−1) ∞<br />

2 /4<br />

,<br />

<br />

d(g)>d(g ′ )+46 ln d(g′ )<br />

d(g ′ )<br />

d 12 (g)e − d2 (g)<br />

4 dg Cd −1 (g ′ )e − d2 (g ′ )<br />

4 .<br />

Par conséquent, <strong>le</strong>s estimations inférieures <strong>de</strong> p (voir (3.8)) nous disent que<br />

<br />

J21 C<br />

H 1<br />

<br />

∇f(g ′ ) p(g ′ )dg ′ .


382 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

4.2.2. <strong>Estimation</strong> <strong>de</strong> J22<br />

Rappelons que<br />

A∞ = e 1000 (L1 + L2) 8 ,<br />

où <strong>le</strong>s <strong>de</strong>ux constantes L1,L2 > 1 proviennent <strong>de</strong> (3.8) et (3.9) respectivement.<br />

Rappelons aussi que<br />

où<br />

<br />

J22 =<br />

Σ<br />

d(g)p(g) <br />

f(g)−f2d−5/2 (g) g N(g) dg,<br />

Σ = R 2 × R \ (0, 0,t); t ∈ R <br />

; d(g) > A∞ ,<br />

N(g)= 100d 3/2 (g) ln d(g) + 1, g N(g) = γ 1 − N(g)d −7/2 (g) ,<br />

et γ désigne <strong>la</strong> géodésique minima<strong>le</strong> (unique) d’origine à g = (x, t), qui est définie par (3.4)–<br />

(3.6).<br />

La difficulté <strong>de</strong> <strong>la</strong> preuve <strong>du</strong> Théorème 1.1 se trouve à montrer qu’il existe une constante<br />

C>0 tel<strong>le</strong> que pour toute f ∈ C∞ o (H1 ),ona<br />

<br />

J22 C p(g) ∇f(g) dg. (4.5)<br />

Dans <strong>la</strong> suite, pour simplifier <strong>le</strong>s notations, on note<br />

Et on remarque que<br />

On estime maintenant<br />

Observons que<br />

N(g)−1 <br />

J22∗ <br />

H 1<br />

g = (x, t) ∈ Σ,<br />

s(i) = 1 − id −7/2 (g), pour 0 i d(g)N(g),<br />

g(i) = x(i),t(i) = x1(i), x2(i), t(i) = γ s(i) .<br />

d g(i),g(j) =|i − j|d −5/2 (g), ∀0 i, j d(g)N(g).<br />

i=0<br />

J22∗ = <br />

f(g)−f2d−5/2 (g) g N(g) .<br />

<br />

f g(i) − f g(i + 1) + f g N(g) <br />

− f2d−5/2 (g) g N(g) .<br />

Par <strong>la</strong> formu<strong>le</strong> <strong>de</strong> représentation (voir (2.2)), on a


et<br />

et<br />

Donc,<br />

H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 383<br />

<br />

f g N(g) − f2d−5/2 (g) g N(g) <br />

<br />

<br />

C<br />

∇f(g ′ ) <br />

d(g(N(g)),g ′ )<br />

|B(g(N(g)),d(g(N(g)),g ′ ))| dg′ ,<br />

B(g(N(g)),4d −5/2 (g))<br />

<br />

f g(i) − f g(i + 1) <br />

f g(i) <br />

− f2d−5/2 (g) g(i) + f g(i + 1) <br />

− f2d−5/2 (g) g(i) <br />

<br />

<br />

C<br />

∇f(g ′ ) <br />

d(g(i),g ′ )<br />

|B(g(i), d(g(i), g ′ ))| dg′<br />

<br />

J22 C<br />

B(g(i),4d −5/2 (g))<br />

+ C<br />

<br />

B(g(i+1),4d −5/2 (g))<br />

N(g) <br />

J22∗ C<br />

Σ<br />

<br />

i=0<br />

B(g(i),4d−5/2 (g))<br />

N(g) <br />

<br />

d(g)p(g)<br />

i=0<br />

<br />

′<br />

∇f(g ) <br />

d(g(i + 1), g ′ )<br />

|B(g(i + 1), d(g(i + 1), g ′ ))| dg′ .<br />

B(g(i),4d −5/2 (g))<br />

<br />

∇f(g ′ ) <br />

d(g(i),g ′ )<br />

|B(g(i), d(g(i), g ′ ))| dg′ , (4.6)<br />

<br />

∇f(g ′ ) <br />

d(g(i),g ′ )<br />

|B(g(i), d(g(i), g ′ ))| dg′<br />

<br />

dg.<br />

Avant <strong>de</strong> continuer <strong>la</strong> preuve <strong>de</strong> (4.5), on doit dire au <strong>le</strong>cteur que l’étape crucial pour démontrer<br />

(4.5) (et donc <strong>le</strong> Théorème 1.1) se trouve à obtenir l’estimation (4.6) et que <strong>le</strong> choix<br />

<strong>de</strong> {g(i)}0ig(N) et d −5/2 (g) dans l’estimation (4.6) n’est pas unique mais très délicat (il dépend<br />

<strong>de</strong>s estimations <strong>optima<strong>le</strong></strong>s <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son <strong>gradient</strong> ainsi que <strong>la</strong> structure<br />

sous-riemannienne <strong>de</strong> H 1 ).<br />

Notons χ <strong>la</strong> fonction caractéristique,<br />

Q = Q(g, g ′ N(g) <br />

) = d(g)<br />

i=0<br />

d(g ′ , g(i))<br />

|B(g(i),d(g ′ , g(i)))| χ<br />

<br />

(g, g ′ ) ∈ Σ × H 1 ; d g ′ ,g(i) <<br />

J ∗ 22 =<br />

<br />

Alors, en utilisant <strong>le</strong> théorème <strong>de</strong> Fubini, on a<br />

<br />

J22 C<br />

Σ<br />

p(g)Qdg.<br />

d(g ′ )>A∞−100(ln A∞)/A∞−5A −5/2<br />

∞<br />

<br />

∇f(g ′ ) J ∗ 22 dg′ .<br />

4<br />

d5/2 <br />

,<br />

(g)


384 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

Donc, pour montrer (4.5), il nous reste à montrer <strong>le</strong> résultat suivant :<br />

Proposition 4.1. Il existe une constante C>0 tel<strong>le</strong> qu’on a<br />

Preuve. Observons que<br />

J ∗ 22 Cp(g′ ), ∀d(g ′ ln A∞<br />

)>A∞ − 100 − 5A<br />

A∞<br />

−5/2<br />

∞ . (4.7)<br />

Q = 0 ⇒ d g ′ ,g(i) 4<br />

<<br />

d5/2 (g)<br />

pour un certain 0 i N(g)+ 1<br />

⇒<br />

d(g ′ ) + 4/d5/2 (g)<br />

1 − 200(ln d(g))/d2 d(g(i))<br />

d(g) =<br />

(g) s(i) d(g′ 4<br />

) −<br />

d5/2 (g)<br />

⇒ d(g′ )/d(g) + 4/d7/2 (g)<br />

1 − 200(ln d(g))/d2 (g) 1 d(g′ )<br />

d(g) −<br />

4<br />

d7/2 .<br />

(g)<br />

(4.8)<br />

En particulier, on a<br />

N(g ′ ) − 1 N(g) N(g ′ ) + 1, et<br />

En utilisant (4.1), on a donc<br />

Q 2d(g ′ N(g<br />

)<br />

′ )+1<br />

Par conséquent,<br />

Rappelons que<br />

i=0<br />

J ∗ 22 2d(g′ N(g<br />

)<br />

′ )+1<br />

2<br />

√ d(g) d(g<br />

5 ′ √<br />

5<br />

) d(g). (4.9)<br />

2<br />

d(g ′ , g(i))<br />

|B(g ′ ,d(g ′ , g(i)))| χ<br />

<br />

(g, g ′ ) ∈ Σ × H 1 ; d g ′ ,g(i) <<br />

i=0<br />

<br />

{g∈Σ;d(g ′ ,g(i))


H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 385<br />

Lemme 4.2. Soit L1 > 1 comme dans (3.8) et soit (g = (x, t), g ′ ) ∈ Σ × H 1 satisfaisant<br />

alors<br />

d g ′ ,g(i) < 4d −5/2 (g) pour un certain 0 i N(g ′ ) + 1,<br />

p(g) 400L 2 1p(g′ )e −i/(5d3/2 (g ′ <br />

)) sin s(i)μ−1 (t/x2 )<br />

sin μ−1 (t/x2 1/2 . (4.11)<br />

)<br />

Lemme 4.3. Il existe une constante C>0 tel<strong>le</strong> que pour tout g ′ ∈ H 1 et tout 0 i N(g ′ ) + 1,<br />

on a<br />

<br />

{g=(x,t)∈Σ;d(g ′ ,g(i))< 4<br />

d 5/2 (g) }<br />

sin s(i)μ −1 (t/x 2 )<br />

sin μ −1 (t/x 2 )<br />

1/2<br />

d(g ′ , g(i))<br />

|B(g ′ ,d(g ′ , g(i)))| dg<br />

Cd −5/2 (g ′ ). (4.12)<br />

On admet <strong>le</strong>s <strong>de</strong>ux <strong>le</strong>mmes pour l’instant et on continue avec <strong>la</strong> démonstration <strong>de</strong> <strong>la</strong> Proposition<br />

4.1.<br />

Fin <strong>de</strong> <strong>la</strong> preuve <strong>de</strong> <strong>la</strong> Proposition 4.1. Par (4.10)–(4.12), on a<br />

J ∗ 22 Cd−3/2 (g ′ )p(g ′ N(g<br />

)<br />

′ )+1<br />

e − i 5 d−3/2 (g ′ <br />

) ′<br />

Cp(g ) 1 +<br />

D’où <strong>la</strong> Proposition 4.1. ✷<br />

i=0<br />

+∞<br />

0<br />

e −h/5 <br />

dh = 6Cp(g ′ ).<br />

Preuve <strong>du</strong> Lemme 4.2. Par <strong>le</strong>s estimations supérieures et inférieures <strong>de</strong> p (voir (3.8)), on a<br />

et<br />

p(g) = p g(i) p(g)<br />

p(g(i)) L21 pg(i) e d2 (g(i))−d2 (g)<br />

4<br />

Et par <strong>la</strong> définition <strong>de</strong> g(i) et (3.5), on a<br />

d g(i) d(g),<br />

1 +x(s(i))d(g(s(i)))<br />

1 +xd(g)<br />

<br />

x s(i) = sin s(i)μ−1 (t/x 2 )<br />

sin μ −1 (t/x 2 ) x,<br />

e d2 (g(i))−d 2 (g)<br />

4 = e − d2 (g)<br />

4 (1−s(i) 2 )


386 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

donc,<br />

p(g) 100L 2 1pg(i) e − i 5 d−3/2 (g ′ <br />

) sin s(i)μ−1 (t/x2 )<br />

sin μ−1 (t/x2 1/2 .<br />

)<br />

Et pour terminer <strong>la</strong> preuve <strong>du</strong> Lemme 4.2, il nous reste à montrer que<br />

En fait, on va montrer que<br />

p g(i) 4p(g ′ ).<br />

<br />

<br />

<br />

p(g(i))<br />

p(g ′ )<br />

<br />

<br />

− 1<br />

1. (4.13)<br />

Preuve <strong>de</strong> (4.13). On observe que<br />

<br />

<br />

<br />

p(g(i))<br />

p(g ′ <br />

<br />

− 1<br />

) = d(g′ , g(i))<br />

p(g ′ <br />

∇p(g∗)<br />

)<br />

avec g∗ ∈ B g ′ , 4d−5/2 (g) .<br />

Mais, par <strong>le</strong>s estimations <strong>de</strong> p et <strong>de</strong> |∇p| (voir (3.8) et (3.9)), on a<br />

1<br />

p(g ′ 1 d<br />

2 ′<br />

L1 1 + d (g ) 2 e<br />

) 2 (g ′ )<br />

4 2L1d(g ′ )e d2 (g ′ )<br />

4 ,<br />

<br />

∇p(g∗) <br />

′ −5/2 −<br />

L2d(g∗)p(g∗) L1L2 d(g ) + 4d (g) e (d(g′ )−4d−5/2 (g)) 2<br />

4<br />

4L1L2d(g ′ )e − d2 (g ′ )<br />

4 ,<br />

par (4.9) et <strong>le</strong> fait que d(g) > 1000.<br />

D’ail<strong>le</strong>urs, (4.9) et <strong>le</strong> fait que g ∈ Σ nous disent que<br />

On a donc (4.13). ✷<br />

d g ′ ,g(i) < 4d −5/2 (g) < 5d −2 (g)A −1/2<br />

∞<br />

< 8L 2 1L2 −1d −2 ′<br />

(g ).<br />

Preuve <strong>du</strong> Lemme 4.3. Par (4.9), on peut supposer que d(g ′ )> 1 2 A∞.<br />

Rappelons que<br />

et que<br />

2ϕ − sin 2ϕ<br />

μ(ϕ) =<br />

2sin2 : ]−π,π[→R,<br />

ϕ<br />

ψ(x,ϕ)= x,μ(ϕ)x 2 : (x, ϕ) ∈ R 2 ×]−π,π[; x = (0, 0) → (x, t) ∈ H 1 ; x = 0 <br />

est un C 1 -difféomorphisme. Notons Dψ(x,θ) <strong>la</strong> matrice jacobienne <strong>de</strong> ψ en (x, θ), ona


et<br />

Posons<br />

H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 387<br />

2sinθ − 2θ cos θ<br />

<strong>de</strong>t Dψ(x,θ) =<br />

sin3 x<br />

θ<br />

2 .<br />

<br />

Ωg ′ = g = (x, t) ∈ H 1 ; x = (0, 0), d(g) > 1<br />

2 d(g′ <br />

) .<br />

Pour 0 i N(g ′ ) + 1, définissons<br />

Θi(g) = g(i): Ωg ′ → H1 \ (0, 0,t)∈ H 1 ; t ∈ R ,<br />

Φi : ψ −1 (Ωg ′) → R 2 \ (0, 0) ×]−π,π[,<br />

(x, θ) = ψ −1 g = (x, t) ↦→ ψ −1 g(i) = x(i), s(i)θ ,<br />

où (voir <strong>la</strong> définition <strong>de</strong> s(i), (3.3) et (3.4))<br />

i<br />

s(i) = s(i)(g) = 1 −<br />

d7/2 (g)<br />

avec d(g) = θ<br />

<br />

t<br />

x, θ= μ−1<br />

sin θ x2 <br />

, (4.14)<br />

sin s(i)θ <br />

x1(i) = x1 cos<br />

sin θ<br />

s(i) − 1 θ + x2 sin s(i) − 1 θ , (4.15)<br />

sin s(i)θ <br />

x2(i) = −x1 sin<br />

sin θ<br />

s(i) − 1 θ + x2 cos s(i) − 1 θ . (4.16)<br />

Observons que Φi et Θi sont <strong>de</strong> c<strong>la</strong>sse C 1 , et que<br />

Θi = ψ ◦ Φi ◦ ψ −1 .<br />

Et on a besoin <strong>de</strong>s <strong>de</strong>ux résultats suivants dont <strong>la</strong> preuve sera donnée plus tard :<br />

(i) Φi est injective <strong>sur</strong> ψ−1 (Ωg ′) ;<br />

(ii) on a<br />

<br />

<strong>de</strong>t DΦi (x, θ) =<br />

sin s(i)θ<br />

sin θ<br />

2 1 + 5<br />

2 id−7/2 <br />

(g) . (4.17)<br />

Les <strong>de</strong>ux résultats précé<strong>de</strong>nts nous disent que Θi est aussi injective <strong>sur</strong> Ωg ′, et pour tout<br />

g = ψ(x,θ)∈ Ωg ′,ona<br />

<strong>de</strong>t DΘi(g) = <strong>de</strong>t Dψ x(i), s(i)θ · <strong>de</strong>t DΦi(x, θ) · <strong>de</strong>t Dψ(x,θ) −1 <br />

sin s(i)θ<br />

4<br />

sin θ − θ cos θ<br />

=<br />

sin θ sin3 −1 sin s(i)θ − s(i)θ cos s(i)θ<br />

θ<br />

sin3 <br />

1 +<br />

s(i)θ<br />

5 i<br />

2 d7/2 <br />

= 0,<br />

(g)


388 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

donc, en distinguant <strong>le</strong>s <strong>de</strong>ux cas, |θ|=μ −1 (|t|/x 2 ) π/2 etπ/2 < |θ|


⎛<br />

⎜<br />

⎝ −<br />

H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 389<br />

sin s(i)θ<br />

sin θ K1 + ∂s(i) dx1(i)<br />

∂x1 ds(i)<br />

sin s(i)θ<br />

sin θ K2 + ∂s(i) dx2(i)<br />

∂x1 ds(i)<br />

θ ∂s(i)<br />

∂x1<br />

sin s(i)θ<br />

sin θ K2 + ∂s(i) dx1(i)<br />

∂x2 ds(i)<br />

sin s(i)θ<br />

sin θ K1 + ∂s(i) dx2(i)<br />

∂x2 ds(i)<br />

θ ∂s(i)<br />

∂x2<br />

∂s(i) dx1(i) dx1(i)<br />

∂θ ds(i) + dθ<br />

∂s(i) dx2(i) dx2(i)<br />

∂θ ds(i) + dθ<br />

s(i) + θ ∂s(i)<br />

∂θ<br />

où <strong>le</strong> sens <strong>de</strong> dx1(i)/ds(i), dx1(i)/dθ, dx2(i)/ds(i), dx2(i)/dθ est c<strong>la</strong>ire.<br />

Soient<br />

R ∗ 1 =<br />

<br />

<br />

<br />

<br />

−<br />

R ∗ 2 =<br />

<br />

<br />

<br />

<br />

<br />

−<br />

<br />

<br />

sin s(i)θ<br />

sin θ K1 + ∂s(i) dx1(i)<br />

∂x1 ds(i)<br />

sin s(i)θ<br />

sin θ K2 + ∂s(i) dx2(i)<br />

∂x1 ds(i)<br />

sin s(i)θ<br />

sin θ K1<br />

sin s(i)θ<br />

sin θ K2<br />

∂s(i)<br />

∂x1<br />

sin s(i)θ<br />

sin θ K2 + ∂s(i) dx1(i)<br />

∂x2 ds(i)<br />

sin s(i)θ<br />

sin θ K1 + ∂s(i) dx2(i)<br />

∂x2 ds(i)<br />

sin s(i)θ<br />

sin θ K2<br />

sin s(i)θ<br />

sin θ K1<br />

∂s(i)<br />

∂x2<br />

R1 = s(i)R ∗ 1 , R2 = θR ∗ 2 .<br />

dx1(i)<br />

dθ<br />

dx2(i)<br />

dθ<br />

∂s(i)<br />

∂θ<br />

Par <strong>le</strong> fait que<br />

<br />

θ ∂s(i)<br />

,θ<br />

∂x1<br />

∂s(i)<br />

,s(i)+ θ<br />

∂x2<br />

∂s(i)<br />

<br />

=<br />

∂θ<br />

0, 0,s(i) <br />

∂s(i)<br />

+ θ ,<br />

∂x1<br />

∂s(i)<br />

,<br />

∂x2<br />

∂s(i)<br />

<br />

,<br />

∂θ<br />

on peut donc écrire<br />

où on a noté<br />

<br />

<br />

<br />

<br />

R∗∗ = <br />

−<br />

<br />

<br />

sin s(i)θ<br />

sin θ K1 + ∂s(i) dx1(i)<br />

∂x1 ds(i)<br />

sin s(i)θ<br />

sin θ K2 + ∂s(i) dx2(i)<br />

∂x1 ds(i)<br />

∂s(i)<br />

∂x1<br />

<strong>de</strong>t DΦi(x, θ) = s(i)R ∗ 1 + θR∗∗,<br />

sin s(i)θ<br />

sin θ K2 + ∂s(i) dx1(i)<br />

∂x2 ds(i)<br />

sin s(i)θ<br />

sin θ K1 + ∂s(i) dx2(i)<br />

∂x2 ds(i)<br />

∂s(i)<br />

∂x2<br />

<br />

<br />

<br />

<br />

<br />

,<br />

<br />

<br />

<br />

<br />

<br />

<br />

,<br />

∂s(i) dx1(i)<br />

∂θ ds(i)<br />

∂s(i) dx2(i)<br />

∂θ ds(i)<br />

∂s(i)<br />

∂θ<br />

⎞<br />

⎟<br />

⎠<br />

<br />

dx1(i)<br />

+ <br />

dθ <br />

dx2(i)<br />

+ <br />

dθ .<br />

<br />

<br />

Or, <strong>la</strong> première ligne <strong>de</strong> R∗∗ peut s’écrire comme<br />

<br />

sin s(i)θ sin s(i)θ dx1(i)<br />

K1, K2, +<br />

sin θ sin θ dθ<br />

dx1(i)<br />

<br />

∂s(i)<br />

,<br />

ds(i) ∂x1<br />

∂s(i)<br />

,<br />

∂x2<br />

∂s(i)<br />

<br />

,<br />

∂θ<br />

et <strong>la</strong> secon<strong>de</strong> ligne <strong>de</strong> R∗∗ peut s’écrire comme<br />

<br />

−<br />

sin s(i)θ<br />

sin θ<br />

on a R∗∗ = R∗ 2 , donc<br />

<br />

sin s(i)θ dx2(i)<br />

K2, K1, +<br />

sin θ dθ<br />

dx2(i)<br />

<br />

∂s(i)<br />

ds(i) ∂x1<br />

, ∂s(i)<br />

,<br />

∂x2<br />

∂s(i)<br />

<br />

,<br />

∂θ<br />

<strong>de</strong>t DΦi(x, θ) = R1 + R2. (4.18)


390 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

Calcul <strong>de</strong> R1. Puisque <strong>la</strong> première colonne <strong>de</strong> R∗ 1 peut s’écrire comme<br />

et <strong>la</strong> secon<strong>de</strong> colonne <strong>de</strong> R ∗ 1<br />

on a<br />

où on a noté<br />

on a<br />

sin s(i)θ<br />

sin θ<br />

<br />

K1<br />

+<br />

−K2<br />

∂s(i)<br />

∂x1<br />

peut s’écrire comme<br />

sin s(i)θ<br />

sin θ<br />

K2<br />

K1<br />

<br />

+ ∂s(i)<br />

∂x2<br />

<br />

dx1(i)/ds(i)<br />

,<br />

dx2(i)/ds(i)<br />

<br />

dx1(i)/ds(i)<br />

,<br />

dx2(i)/ds(i)<br />

R ∗ 1 =<br />

2 <br />

sin s(i)θ K1 K2 <br />

<br />

sin θ −K2 K1 <br />

+ sin s(i)θ<br />

<br />

∂s(i) <br />

<br />

sin θ ∂x2<br />

K1<br />

−K2<br />

<br />

dx1(i)/ds(i) <br />

<br />

∂s(i) <br />

dx2(i)/ds(i) + <br />

∂x1<br />

dx1(i)/ds(i)<br />

dx2(i)/ds(i)<br />

<br />

K2 <br />

<br />

K1 <br />

2 sin s(i)θ<br />

=<br />

+<br />

sin θ<br />

sin s(i)θ <br />

K1R<br />

sin θ<br />

∗ 11 + K2R ∗ <br />

12 ,<br />

R ∗ 11<br />

∂s(i) dx2(i)<br />

=<br />

∂x2 ds(i)<br />

∂s(i) dx1(i)<br />

+<br />

∂x1 ds(i) , R∗ 12<br />

∂s(i) dx1(i)<br />

=<br />

∂x2 ds(i)<br />

− ∂s(i)<br />

∂x1<br />

Il nous reste à calcu<strong>le</strong>r ∂s(i)/∂x1, ∂s(i)/∂x2, dx1(i)/ds(i) et dx2(i)/ds(i).<br />

Dans <strong>la</strong> suite, pour simplifier <strong>le</strong>s notations, on pose<br />

s ′ (i) =<br />

d 7<br />

s(i) =<br />

d(d(g)) 2 id−9/2 (g).<br />

On commence par calcu<strong>le</strong>r ∂s(i)/∂x1 et ∂s(i)/∂x2. Par <strong>le</strong> fait que (voir (3.3))<br />

d 2 <br />

θ<br />

2x<br />

2<br />

(g) =<br />

1 + x<br />

sin θ<br />

2 2 ,<br />

dx2(i)<br />

. (4.19)<br />

ds(i)<br />

∂s(i)<br />

= s<br />

∂x1<br />

′ 1 ∂d<br />

(i)<br />

2d(g)<br />

2 (g)<br />

= s<br />

∂x1<br />

′ (i) 1<br />

2 θ<br />

x1,<br />

d(g) sin θ<br />

(4.20)<br />

∂s(i)<br />

= s<br />

∂x2<br />

′ 1 ∂d<br />

(i)<br />

2d(g)<br />

2 (g)<br />

= s<br />

∂x2<br />

′ (i) 1<br />

2 θ<br />

x2.<br />

d(g) sin θ<br />

(4.21)<br />

On calcu<strong>le</strong> maintenant dx1(i)/ds(i) et dx2(i)/ds(i). Par (4.15), (4.16) et <strong>de</strong>s formu<strong>le</strong>s élémentaires<br />

liant <strong>le</strong>s fonctions trigonométriques, on peut écrire


Donc,<br />

H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 391<br />

x1(i) = 1 <br />

x1 sin θ + sin 2s(i) − 1 θ + x2 cos θ − cos 2s(i) − 1 θ ,<br />

2sinθ<br />

(4.22)<br />

x2(i) = 1 <br />

−x1 cos θ − cos 2s(i) − 1 θ + x2 sin θ + sin 2s(i) − 1 θ .<br />

2sinθ<br />

(4.23)<br />

dx1(i)<br />

ds(i)<br />

dx2(i)<br />

ds(i)<br />

θ <br />

= x1 cos<br />

sin θ<br />

2s(i) − 1 θ + x2 sin 2s(i) − 1 θ , (4.24)<br />

θ <br />

= −x1 sin<br />

sin θ<br />

2s(i) − 1 θ + x2 cos 2s(i) − 1 θ . (4.25)<br />

Par (4.19)–(4.21), (4.24) et (4.25), on constate d’abord que<br />

R ∗ 11 = s′ (i) 1<br />

<br />

θ<br />

d(g)<br />

sin θ<br />

R ∗ 12 = s′ (i) 1<br />

<br />

θ<br />

d(g) sin θ<br />

3<br />

x 2 cos 2s(i) − 1 θ,<br />

3<br />

x 2 sin 2s(i) − 1 θ,<br />

puis, en utilisant <strong>le</strong> fait que d 2 (g) = ( θ<br />

sin θ )2 x 2 (voir (3.3)), que<br />

Donc,<br />

R ∗ 11 = s′ (i)d(g) θ<br />

sin θ cos 2s(i) − 1 θ,<br />

R ∗ 12 = s′ (i)d(g) θ<br />

sin θ sin 2s(i) − 1 θ.<br />

K1R ∗ 11 + K2R ∗ 12<br />

= s ′ (i)d(g) θ <br />

cos s(i) − 1 θ cos 2s(i) − 1 θ + sin s(i) − 1 θ sin 2s(i) − 1 θ<br />

sin θ<br />

= s ′ (i)d(g) θ<br />

cos s(i)θ.<br />

sin θ<br />

Par conséquent,<br />

Calcul <strong>de</strong> R2. Soient<br />

<br />

sin s(i)θ<br />

R1 = s(i)<br />

sin θ<br />

2 sin s(i)θ<br />

+ s(i) s<br />

sin θ<br />

′ (i)d(g) θ<br />

cos s(i)θ. (4.26)<br />

sin θ<br />

R ∗ 21 =<br />

<br />

sin s(i)θ<br />

sin θ<br />

2 ∂s(i)<br />

∂θ ,


392 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

R ∗ 22<br />

<br />

<br />

sin s(i)θ dx1(i) ∂s(i) ∂s(i)<br />

= −K2 − K1 −<br />

sin θ dθ ∂x2 ∂x1<br />

dx2(i)<br />

<br />

<br />

∂s(i) ∂s(i)<br />

K1 − K2<br />

dθ ∂x2 ∂x1<br />

= sin s(i)θ<br />

<br />

dx2(i) ∂s(i)<br />

K2<br />

−<br />

sin θ dθ ∂x1<br />

dx1(i)<br />

<br />

∂s(i) dx1(i) ∂s(i)<br />

− K1<br />

+<br />

dθ ∂x2 dθ ∂x1<br />

dx2(i)<br />

<br />

∂s(i)<br />

.<br />

dθ ∂x2<br />

On développe suivant <strong>la</strong> troisième colonne <strong>de</strong> R ∗ 2 , et on constate que R∗ 2 = R∗ 21 + R∗ 22<br />

R2 = θR ∗ 21 + θR∗ 22 .<br />

En rappe<strong>la</strong>nt que d(g) = θ<br />

sin θ x,ona<br />

donc,<br />

∂s(i)<br />

∂θ = s′ (i) ∂d(g)<br />

∂θ = s′ (i)<br />

R ∗ 21 =<br />

<br />

sin s(i)θ<br />

sin θ<br />

sin θ − θ cos θ<br />

sin 2 θ<br />

2<br />

s ′ (i)d(g) 1<br />

θ<br />

x=s ′ (i)d(g) 1<br />

θ<br />

sin θ − θ cos θ<br />

.<br />

sin θ<br />

sin θ − θ cos θ<br />

,<br />

sin θ<br />

Pour calcu<strong>le</strong>r R∗ dx1(i)<br />

22 , on aura besoin <strong>de</strong> l’expression explicite <strong>de</strong> dθ et <strong>de</strong> dx2(i)<br />

dθ . Notons<br />

U1 = d sin θ + sin(2s(i) − 1)θ<br />

=<br />

dθ 2sinθ<br />

2s(i)sin θ cos(2s(i) − 1)θ − sin 2s(i)θ<br />

U2 = d cos θ − cos(2s(i) − 1)θ<br />

=<br />

dθ 2sinθ<br />

−1 + 2s(i)sin θ sin(2s(i) − 1)θ + cos 2s(i)θ<br />

Alors, par (4.22) et (4.23),<br />

dx1(i)<br />

dθ = x1U1 + x2U2,<br />

En utilisant (4.20) et (4.21), on constate que<br />

2sin 2 θ<br />

2sin 2 θ<br />

dx2(i)<br />

dθ =−x1U2 + x2U1.<br />

,<br />

.<br />

et donc<br />

R ∗ sin s(i)θ<br />

22 = s<br />

sin θ<br />

′ (i) 1<br />

2 θ<br />

x<br />

d(g) sin θ<br />

2 (−K2U2 − K1U1)<br />

sin s(i)θ<br />

=− s<br />

sin θ<br />

′ (i)d(g)(K1U1 + K2U2) par (3.3)<br />

sin s(i)θ<br />

=− s<br />

sin θ<br />

′ s(i)sin θ cos s(i)θ − cos θ sin s(i)θ<br />

(i)d(g)<br />

sin2 ,<br />

θ<br />

où on a utilisé trois fois <strong>le</strong>s formu<strong>le</strong>s élémentaires liant <strong>le</strong>s fonctions trigonométriques dans <strong>la</strong><br />

<strong>de</strong>rnière égalité.<br />

On a donc<br />

R2 = θR ∗ 21 + θR∗ 22 = s′ <br />

sin s(i)θ<br />

(i)d(g)<br />

sin θ<br />

Par (4.18), (4.26) et (4.27), on obtient (4.17). ✷<br />

2 sin s(i)θ<br />

− θ<br />

sin θ<br />

<br />

s(i)cos s(i)θ<br />

. (4.27)<br />

sin θ


5. Quelques problèmes ouverts<br />

H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394 393<br />

Pour terminer cet artic<strong>le</strong>, on donne quelques questions intéressantes.<br />

1. J’espère que <strong>le</strong> <strong>le</strong>cteur peut suivre <strong>la</strong> métho<strong>de</strong> <strong>de</strong> <strong>la</strong> preuve <strong>du</strong> Théorème 1.1 pour obtenir un<br />

résultat analogue au Théorème 1.1 dans <strong>le</strong> cadre <strong>de</strong>s <strong>groupe</strong>s <strong>de</strong> Heisenberg <strong>de</strong> dimension<br />

réel<strong>le</strong> 5 et aussi dans <strong>le</strong> cadre <strong>de</strong>s cônes c<strong>la</strong>ssiques, C(Sϖ ) avec 0 0, f∈ C ∞ o (M).<br />

En particulier, on s’intéressait à obtenir <strong>de</strong>s estimations <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et <strong>de</strong> son<br />

<strong>gradient</strong> à partir d’une tel<strong>le</strong> inégalité.<br />

3. Sur <strong>le</strong>s variétés coniques, C(N), à cause <strong>de</strong> <strong>la</strong> singu<strong>la</strong>rité conique, on trouve qu’ils ont beaucoup<br />

<strong>de</strong> propriétés spécia<strong>le</strong>s par rapport aux variétés riemanniennes complètes à courbure<br />

<strong>de</strong> Ricci minorée, voir par exemp<strong>le</strong> [8,15–17] ou [18]. Lorsque <strong>la</strong> première va<strong>le</strong>ur propre<br />

non nul<strong>le</strong> <strong>du</strong> Lap<strong>la</strong>cien <strong>sur</strong> N, λ1 dim N, ce sera très intéressant <strong>de</strong> savoir pour quels<br />

1 w 0, f∈ Co .<br />

Remarquons que dans cette situation, <strong>la</strong> courbure <strong>de</strong> Ricci <strong>sur</strong> C(N) peut être non minorée.<br />

Et on rappel<strong>le</strong> aussi que pour λ1 < dim N, <strong>le</strong>s estimations précé<strong>de</strong>ntes ne sont pas va<strong>la</strong>b<strong>le</strong>s<br />

(voir [17] ou [18]).<br />

4. Dans l’étu<strong>de</strong> <strong>de</strong> <strong>la</strong> formu<strong>le</strong> <strong>de</strong> représentation, ce sera très intéressant <strong>de</strong> savoir si <strong>la</strong> condition<br />

technique (H2) dans [19], i.e. en supposant que <strong>le</strong> volume <strong>de</strong>s bou<strong>le</strong>s est au moins à<br />

croissance linéaire, est nécessaire. En tout cas, cette condition est faib<strong>le</strong>.<br />

Remerciements<br />

L’auteur est soutenu par <strong>le</strong> DFG (SFB 611). Je tiens à remercier T. Coulhon et L. Saloff-Coste<br />

pour <strong>de</strong>s discutions et <strong>de</strong> nombreuses suggestions, F. Lust-Piquard et M. Vil<strong>la</strong>ni pour m’avoir<br />

communiqué [21]. Je voudrais remercier aussi <strong>le</strong> référé pour m’avoir indiqué une faute dans <strong>la</strong><br />

version originaire <strong>de</strong> cet artic<strong>le</strong> et <strong>de</strong>s suggestions.<br />

Références<br />

[1] C. Ané, S. B<strong>la</strong>chère, D. Chafaï, P. Fougères, I. Gentil, F. Malrieu, C. Roberto, G. Scheffer, Sur <strong>le</strong>s inégalités <strong>de</strong><br />

Sobo<strong>le</strong>v logarithmiques, in : Panoramas et Synthèses, vol. 10, Soc. Math. France, Paris, 2000.<br />

[2] P. Auscher, T. Coulhon, X.T. Duong, S. Hofmann, Riesz transform on manifolds and heat kernel regu<strong>la</strong>rity, Ann.<br />

Sci. Éco<strong>le</strong> Norm. Sup. (4) 37 (2004) 911–957.<br />

[3] D. Bakry, Transformations <strong>de</strong> Riesz pour <strong>le</strong>s <strong>semi</strong>-<strong>groupe</strong>s symétriques. II. Étu<strong>de</strong> sous <strong>la</strong> condition Γ2 0, in :<br />

Séminaire <strong>de</strong> probabilités, XIX, 1983/84, in : Lecture Notes in Math., vol. 1123, Springer, Berlin, 1985, pp. 145–<br />

174.<br />

[4] D. Bakry, Un critère <strong>de</strong> non-explosion pour certaines diffusions <strong>sur</strong> une variété riemannienne complète, C. R. Acad.<br />

Sci. Paris Sér. I Math. 303 (1986) 23–26.


394 H.-Q. Li / Journal of Functional Analysis 236 (2006) 369–394<br />

[5] R. Beals, B. Gaveau, P.C. Greiner, Hamilton–Jacobi theory and the heat kernel on Heisenberg groups, J. Math. Pures<br />

Appl. (9) 79 (2000) 633–689.<br />

[6] J. Cheeger, M. Gromov, M. Taylor, Finite propagation speed, kernel estimates for functions of the Lap<strong>la</strong>ce operator,<br />

and the geometry of comp<strong>le</strong>te Riemannian manifolds, J. Differential Geom. 17 (1982) 15–53.<br />

[7] T. Coulhon, X.-T. Duong, Riesz transform and re<strong>la</strong>ted inequalities on noncompact Riemannian manifolds, Comm.<br />

Pure Appl. Math. 56 (2003) 1728–1751.<br />

[8] T. Coulhon, H.-Q. Li, <strong>Estimation</strong>s inférieures <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> <strong>le</strong>s variétés coniques et transformées <strong>de</strong><br />

Riesz, Arch. Math. 83 (2004) 229–242.<br />

[9] B.K. Driver, T. Melcher, Hypoelliptic heat kernel inequalities on the Heisenberg group, J. Funct. Anal. 221 (2005)<br />

340–365.<br />

[10] B. Franchi, G. Lu, R.L. Whee<strong>de</strong>n, Representation formu<strong>la</strong>s and weighted Poincaré inequalities for Hörman<strong>de</strong>r<br />

vector fields, Ann. Inst. Fourier (Grenob<strong>le</strong>) 45 (1995) 577–604.<br />

[11] B. Franchi, G. Lu, R.L. Whee<strong>de</strong>n, A re<strong>la</strong>tionship between Poincaré-type inequalities and representation formu<strong>la</strong>s in<br />

spaces of homogeneous type, Int. Math. Res. Not. (1996) 1–14.<br />

[12] B. Gaveau, Principe <strong>de</strong> moindre action, propagation <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> et estimées sous-elliptiques <strong>sur</strong> certains <strong>groupe</strong>s<br />

nilpotents, Acta Math. 139 (1977) 95–153.<br />

[13] H. Hueber, D. Mül<strong>le</strong>r, Asymptotics for some Green kernels on the Heisenberg group and the Martin boundary,<br />

Math. Ann. 283 (1989) 97–119.<br />

[14] A. Hu<strong>la</strong>nicki, The distribution of energy in the Brownian motion in the Gaussian field and analytic-hypoellipticity<br />

of certain subelliptic operators on the Heisenberg group, Studia Math. 56 (1976) 165–173.<br />

[15] H.-Q. Li, La transformation <strong>de</strong> Riesz <strong>sur</strong> <strong>le</strong>s variétés coniques, J. Funct. Anal. 168 (1999) 145–238.<br />

[16] H.-Q. Li, <strong>Estimation</strong>s <strong>du</strong> noyau <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> <strong>le</strong>s variétés coniques et ses applications, Bull. Sci. Math. 124<br />

(2000) 365–384.<br />

[17] H.-Q. Li, Sur <strong>la</strong> continuité <strong>de</strong> Höl<strong>de</strong>r <strong>du</strong> <strong>semi</strong>-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> <strong>le</strong>s variétés coniques, C. R. Math. Acad. Sci.<br />

Paris 337 (2003) 283–286.<br />

[18] H.-Q. Li, Sur <strong>la</strong> continuité <strong>de</strong> Höl<strong>de</strong>r <strong>du</strong> <strong>semi</strong>-<strong>groupe</strong> <strong>de</strong> <strong>la</strong> <strong>cha<strong>le</strong>ur</strong> <strong>sur</strong> <strong>le</strong>s variétés coniques, soumis.<br />

[19] G.Z. Lu, R.L. Whee<strong>de</strong>n, An optimal representation formu<strong>la</strong> for Carnot–Carathéodory vector fields, Bull. London<br />

Math. Soc. 30 (1998) 578–584.<br />

[20] F. Lust-Piquard, A simp<strong>le</strong>-min<strong>de</strong>d computation of heat kernels on Heisenberg groups, Colloq. Math. 97 (2003)<br />

233–249.<br />

[21] F. Lust-Piquard, M. Vil<strong>la</strong>ni, Heat inequalities on the Heisenberg group: An analytic proof of a result by Driver–<br />

Melcher, prépublication.<br />

[22] P. Maheux, L. Saloff-Coste, Analyse <strong>sur</strong> <strong>le</strong>s bou<strong>le</strong>s d’un opérateur sous-elliptique, Math. Ann. 303 (1995) 713–740.<br />

[23] T. Melcher, Hypoelliptic heat kernel inequalities on Lie groups, PhD thesis, 2004.<br />

[24] M.-K. von Renesse, K.-T. Sturm, Transport inequalities, <strong>gradient</strong> estimates, entropy and Ricci curvature, Comm.<br />

Pure Appl. Math. 58 (2005) 923–940.<br />

[25] N.T. Varopoulos, L. Saloff-Coste, T. Coulhon, Analysis and Geometry on Groups, Cambridge Tracts in Math.,<br />

vol. 100, Cambridge Univ. Press, Cambridge, 1992.


Journal of Functional Analysis 236 (2006) 395–408<br />

Large Fredholm trip<strong>le</strong>s ✩<br />

Dan Kucerovsky<br />

Department of Mathematics and Statistics, UNB-F, Fre<strong>de</strong>ricton, NB, Canada E3B 5A3<br />

Received 28 July 2005; accepted 15 November 2005<br />

Avai<strong>la</strong>b<strong>le</strong> online 24 April 2006<br />

Communicated by A<strong>la</strong>in Connes<br />

www.elsevier.com/locate/jfa<br />

Abstract<br />

We give several equiva<strong>le</strong>nt condition for Busby extensions of a given algebra to be absorbing, consi<strong>de</strong>rably<br />

improving our earlier results [G.A. Elliott, D. Kucerovsky, An abstract Brown–Doug<strong>la</strong>s–Fillmore<br />

absorption theorem, Pacific J. Math. 198 (2001) 385–409], and establish sufficient conditions for Fredholm<br />

trip<strong>le</strong>s to be absorbing in a suitab<strong>le</strong> sense. As an application of one of our criteria, we prove a multivariab<strong>le</strong><br />

Brown–Doug<strong>la</strong>s–Fillmore type theorem.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: K-Theory; C ∗ -Algebras<br />

1. Intro<strong>du</strong>ction and background<br />

We study the prob<strong>le</strong>m of simplifying the standard equiva<strong>le</strong>nce re<strong>la</strong>tion on KK-theory. The<br />

theorems obtained are applicab<strong>le</strong> to a c<strong>la</strong>ss of algebra that inclu<strong>de</strong>s many real rank zero algebras,<br />

purely infinite algebras (simp<strong>le</strong> or not), and many type I C ∗ -algebras. In hopes of making this<br />

paper somewhat self-contained, we shall inclu<strong>de</strong> background material on KK-theory [10,11].<br />

Since we make fundamental use of a purely <strong>la</strong>rge condition that was originally phrased in terms<br />

of the generalized Brown–Doug<strong>la</strong>s–Fillmore group of extensions, Ext(A, B), we shall initially<br />

work in this setting; however, we <strong>la</strong>ter consi<strong>de</strong>r the case of Fredholm trip<strong>le</strong>s, in particu<strong>la</strong>r the c<strong>la</strong>ss<br />

we term absorbing trip<strong>le</strong>s. The paper is organized as follows. In the first two sections we give<br />

<strong>de</strong>finitions and establish some equiva<strong>le</strong>nt forms for the purely <strong>la</strong>rge property. In the third section,<br />

we give <strong>le</strong>mmas and technical results of various sorts. In Section 5 we establish a topological<br />

✩ Supported by NSERC, un<strong>de</strong>r grant #228065-00.<br />

E-mail address: dkucerov@unb.ca.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2005.11.018


396 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

formu<strong>la</strong>tion of an absorption criterion that <strong>le</strong>ads to a multivariab<strong>le</strong> Brown–Doug<strong>la</strong>s–Fillmore<br />

theorem in Section 6. In Section 7 we apply the techniques we have built up to the case of<br />

Fredholm trip<strong>le</strong>s, giving a nice sufficient condition for a trip<strong>le</strong> to be absorbing.<br />

2. Preliminaries<br />

As pointed out by Kasparov [11], the Ext group can be <strong>de</strong>fined as equiva<strong>le</strong>nce c<strong>la</strong>sses of absorbing<br />

extensions un<strong>de</strong>r unitary equiva<strong>le</strong>nce by multiplier unitaries, but on the other hand, it<br />

is often convenient to work with lifted Busby maps [5]. Thus, we <strong>de</strong>fine an extension to be an<br />

injective comp<strong>le</strong>tely positive map into the multipliers which becomes a homomorphism when<br />

composed with the canonical map into the corona. The equiva<strong>le</strong>nce re<strong>la</strong>tion we consi<strong>de</strong>r is unitary<br />

equiva<strong>le</strong>nce by multiplier unitaries mo<strong>du</strong>lo the i<strong>de</strong>al. In or<strong>de</strong>r to effectively apply this mo<strong>de</strong>l<br />

of Kasparov’s group KK 1 to concrete prob<strong>le</strong>ms, one needs to be ab<strong>le</strong> to <strong>de</strong>ci<strong>de</strong> which extensions<br />

are absorbing. Before giving a re<strong>le</strong>vant criterion, for the rea<strong>de</strong>r’s convenience we recall several<br />

<strong>de</strong>finitions:<br />

Definition. Let B be a σ -unital, nuc<strong>le</strong>ar, and stab<strong>le</strong> C ∗ -algebra. Let A be a separab<strong>le</strong> C ∗ -algebra.<br />

(i) An extension τ : A → M(B) is said to be full if π ◦ τ : A → M(B)/B intersects no nontrivial<br />

i<strong>de</strong>al of the corona. Thus, π(τ(C ∗ (a))) ∩ I ={0} for all proper i<strong>de</strong>als I of the corona,<br />

and all positive a ∈ A.<br />

(ii) An extension τ : A → M(B) is said to be trivial if it is a homomorphism (rather than just a<br />

comp<strong>le</strong>tely positive map).<br />

(iii) An extension τ : A → M(B) is said to be essential if the map π ◦ τ has no kernel.<br />

(iv) An extension τ : A → M(B) is said to be unital if A is unital and τ maps the unit of A to<br />

the unit of the multipliers. Otherwise, the extension is said to be nonunital.<br />

(v) An extension τ is weakly nuc<strong>le</strong>ar if the maps a ↦→ bτ(a)b ∗ are nuc<strong>le</strong>ar for every b in B.For<br />

an extension to be weakly nuc<strong>le</strong>ar, it is sufficient that either A or B is a nuc<strong>le</strong>ar C ∗ -algebra.<br />

(vi) The BDF sum [4] of two extensions τ and φ is the extension v1τv ∗ 1 + v2φv ∗ 2 where v1 and<br />

v2 are the generators of some given copy of O2 in the multipliers (such a copy exists if B is<br />

stab<strong>le</strong>, and is unique up to unitary equiva<strong>le</strong>nce).<br />

(vii) A weakly nuc<strong>le</strong>ar extension π is said to be absorbing if the BDF sum of π with a weakly<br />

nuc<strong>le</strong>ar trivial extension φ is approximately unitarily equiva<strong>le</strong>nt to π, whenever φ and π are<br />

either both unital, or both nonunital.<br />

We notice that fullness, as <strong>de</strong>fined above, is an extremely strong condition. However, unital<br />

extensions τ : A → M(B) of simp<strong>le</strong> (unital) C ∗ -algebras A are necessarily full, as are absorbing<br />

extensions.<br />

3. C ∗ -algebraic absorption criteria<br />

Recalling [5] that an extension τ : A → M(B) can equiva<strong>le</strong>ntly be given as a (<strong>semi</strong>split)<br />

short exact sequence 0 → B → C → A → 0, we say that an extension is purely <strong>la</strong>rge if the<br />

extension algebra C has the property that, for every positive e<strong>le</strong>ment c ∈ C + , either c is in B, or<br />

else the hereditary subalgebra cBc contains a stab<strong>le</strong> subalgebra that generates B as an i<strong>de</strong>al. This<br />

<strong>de</strong>finition is motivated by the following theorem [6], which is, in fact, the first of our C ∗ -algebraic<br />

absorption criteria. The importance of the theorem is that it shows (in the spirit of in<strong>de</strong>x theory!)


D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408 397<br />

that a purely topological property, namely absorption, is equiva<strong>le</strong>nt to an algebraic property. The<br />

specific algebraic property in question is the somewhat technical-sounding purely <strong>la</strong>rge property<br />

that we have just <strong>de</strong>fined, but we will shortly come up with nicer criteria.<br />

Theorem 3.1. Let A and B be separab<strong>le</strong> C ∗ -algebras, with B stab<strong>le</strong>. Let 0 → B → C → A → 0<br />

be an essential (unital) extension with weakly nuc<strong>le</strong>ar splitting map (a comp<strong>le</strong>tely positive map<br />

s : A → C such that a ↦→ bs(a)b ∗ is nuc<strong>le</strong>ar for each b ∈ B). Then the following are equiva<strong>le</strong>nt:<br />

(i) The extension absorbs all trivial weakly nuc<strong>le</strong>ar (unital) extensions.<br />

(ii) The extension is purely <strong>la</strong>rge.<br />

(iii) The extension algebra has the approximation property that, for every c ∈ C + that is not zero<br />

in C/B, and every positive b in B, there is an e<strong>le</strong>ment r ∈ B making the norm of b − rcr ∗<br />

arbitrarily small. Moreover, the e<strong>le</strong>ment r can be assumed to be in the unit ball if b and c/B<br />

are of norm one.<br />

It is a corol<strong>la</strong>ry of the above theorem that an extension τ : A → M(B) is absorbing if and<br />

only if its restrictions to subalgebras C ∗ (a), with a ∈ A + , are. In the main application of the<br />

theorem, it is customary to assume that the absorbing extension and the absorbed extension are<br />

unital or nonunital together. This stems from the observation that the BDF sum of a unital and a<br />

nonunital extension is never unital, so that a unital extension therefore cannot absorb a nonunital<br />

extension. It is not hard to check that, although this is not important in BDF theory, a nonunital<br />

absorbing extension in fact absorbs a unital extension, and it is not even necessary for the<br />

absorbed extension to be essential (the absorbing extension must, however, be essential).<br />

Definition. We say that an algebra B is an absorbing algebra if every full weakly nuc<strong>le</strong>ar extension<br />

of B by a separab<strong>le</strong> algebra A is absorbing.<br />

Of course, K and O2 ⊗ K are the canonical examp<strong>le</strong>s of absorbing algebras [6]. In addition,<br />

many real rank zero algebras, all purely infinite stab<strong>le</strong> algebras (simp<strong>le</strong> or not) [21], and many<br />

type I C ∗ -algebras [17] are absorbing.<br />

We now prove a p<strong>le</strong>asantly geometrical if and only if criterion for absorption, phrased in terms<br />

of a weak form of Fredholmness for full projections. We also obtain several alternative forms of<br />

the criterion.<br />

Theorem 3.2. Let B be a stab<strong>le</strong> and separab<strong>le</strong> C ∗ -algebra. Then the following are equiva<strong>le</strong>nt:<br />

(i) B is an absorbing algebra.<br />

(ii) All full multiplier projections in M(B) have quasi-invertib<strong>le</strong> image in the corona.<br />

(iii) All full multiplier projections are properly infinite.<br />

(iv) All full corona projections are quasi-invertib<strong>le</strong>.<br />

(v) All full corona projections are properly infinite.<br />

(vi) All full multiplier projections majorize, in the corona, a corona projection equiva<strong>le</strong>nt to 1.<br />

In the statement of the above theorem, a projection is said to be full if the i<strong>de</strong>al it generates is<br />

the who<strong>le</strong> algebra. A projection P is said to be quasi-invertib<strong>le</strong> if there exists an e<strong>le</strong>ment x such<br />

that xPx ∗ = 1.


398 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

The above theorem will be further generalized in a future publication, [16], where we also<br />

apply it to the study of Rørdam’s group KL 1 (A, B). We notice that the theorem suggests that<br />

absorption can be characterized so<strong>le</strong>ly in terms of the image in the corona of an extension—this<br />

was, however, already pointed out in [6].<br />

The proof is <strong>de</strong>ferred to page 400, as we need to establish a few <strong>le</strong>mmas and propositions first.<br />

Remark 3.3. Since the canonical i<strong>de</strong>al is stab<strong>le</strong>, quasi-invertibility in the multipliers is equiva<strong>le</strong>nt<br />

to quasi-invertibility in the corona. To see this, suppose that for some corona e<strong>le</strong>ment c,wehave<br />

1 = rcr ′ for some r. Then, lifting c to any full e<strong>le</strong>ment ˜c in the multipliers, we have ˜r ˜c˜r ′ = 1 + b<br />

for some e<strong>le</strong>ment b of the i<strong>de</strong>al. But since the i<strong>de</strong>al is stab<strong>le</strong>, we can find an isometry v such that<br />

v ∗ bv is small in norm, implying that v ∗ ˜r ˜c˜r ′ v is invertib<strong>le</strong>. In fact, by the same type of argument,<br />

if an extension is full in the sense of Definition 2, then e<strong>le</strong>ments of the extension algebra that are<br />

not in the canonical i<strong>de</strong>al are in fact full in the multipliers.<br />

Remark 3.4. It follows from the theorem that an algebra B is absorbing if and only if every<br />

nonunital full extension of C by B is absorbing.<br />

We point out that an early version of part (ii) of this theorem motivated the corona factorization<br />

condition proposed in a preprint [13], and several conference talks, as part of a strategy for<br />

studying i<strong>de</strong>al-re<strong>la</strong>ted absorption:<br />

Definition. A stab<strong>le</strong> σ -unital C ∗ -algebra B has the corona factorization property if, for every<br />

positive c in M(B) + , either of the following equiva<strong>le</strong>nt conditions hold:<br />

(i) there is an r in M(BcB)/BcB such that rcr ∗ = 1inM(BcB)/BcB, and<br />

(ii) there is an r in M(BcB) such that rcr ∗ = 1inM(BcB).<br />

(The equiva<strong>le</strong>nce of (i) and (ii) follows from the fact that i<strong>de</strong>als in a stab<strong>le</strong> C ∗ -algebra are<br />

stab<strong>le</strong>, allowing a cut-down by a suitab<strong>le</strong> isometry as in the above remark.)<br />

The corona factorization condition refined by means of a spectral condition is used in [14] to<br />

study i<strong>de</strong>al-re<strong>la</strong>ted absorption.<br />

4. Lemmas and technical results<br />

We of course need to use several properties of purely <strong>la</strong>rge algebras in proving Theorem 3.2.<br />

The most important is our abstract Weyl–von Neumann type theorem:<br />

Lemma 4.1. [6] Let B be a separab<strong>le</strong> stab<strong>le</strong> C ∗ -algebra. Let C be a separab<strong>le</strong> subalgebra of<br />

M(B), containing B and 1M(B) and having the purely <strong>la</strong>rge property with respect to B. Let<br />

φ : C → M(B) be a comp<strong>le</strong>tely positive weakly nuc<strong>le</strong>ar map which is zero on B. Then there<br />

exists a sequence (vn) of isometries such that, for each c ∈ C, the expression φ(c) − v ∗ n cvn is<br />

in B, and goes to zero in norm as n →∞.<br />

Remark 4.2. In view of the condition on units, it is useful to note that the unitization of a purely<br />

<strong>la</strong>rge algebra is purely <strong>la</strong>rge. More precisely, if the image of C in the corona is not already unital,<br />

then one of the arguments in [6] shows that C + 1M(B) is purely <strong>la</strong>rge. The i<strong>de</strong>a is the following.<br />

Given that for all positive c ∈ C − B, the hereditary subalgebra cBc contains a stab<strong>le</strong> full (in B)


D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408 399<br />

subalgebra, we are to show the same for (1 + c)B(1 + c). Because the image of C in the corona is<br />

not unital, there exists c ′ ∈ C such that c ′ (1 + c) is in C but not in B, so that (1 + c)c ′ Bc ′ (1 + c)<br />

contains a full stab<strong>le</strong> subalgebra. But this then gives a full stab<strong>le</strong> subalgebra of the <strong>la</strong>rger algebra<br />

(1 + c)B(1 + c). This, in fact, is exactly how the nonunital version of the main result of [6] is<br />

<strong>de</strong>rived from the easier unital case.<br />

The second major theorem that we need is the Kasparov stabilization theorem [10], one of<br />

the fundamental tools of Kasparov’s KK-theory [11], which will here be used to “diagonalize” a<br />

singly generated hereditary subalgebra of a stab<strong>le</strong> algebra. One form of the Kasparov stabilization<br />

theorem is as follows.<br />

Theorem 4.3. [19] Let E be a Hilbert B-mo<strong>du</strong><strong>le</strong> that is countably generated in M(E). Ifwe<br />

<strong>de</strong>note the standard Hilbert B-mo<strong>du</strong><strong>le</strong> by HB, then E ⊕ HB is unitarily equiva<strong>le</strong>nt to HB.<br />

In the statement of the theorem, a Hilbert mo<strong>du</strong><strong>le</strong> E is said to be countably generated in<br />

M(E) if there is a sequence of multipliers (mi) ⊆ M(E) such that the e<strong>le</strong>ments {mib: b ∈ B}<br />

span a <strong>de</strong>nse submo<strong>du</strong><strong>le</strong> of E. This <strong>de</strong>finition generalizes Kasparov’s, since the generators mi do<br />

not need to be in the mo<strong>du</strong><strong>le</strong> E. It is likely that the assumption of σ -uniticity can be removed in<br />

the next theorem.<br />

Theorem 4.4. Let B be stab<strong>le</strong> and σ -unital. A hereditary subalgebra, ℓBℓ ∗ , generated by an<br />

e<strong>le</strong>ment ℓ of the multipliers M(B) is isomorphic to a hereditary subalgebra generated by a<br />

multiplier projection, P .Ifℓ is not contained in a proper i<strong>de</strong>al of the multipliers, then neither<br />

is P .<br />

Proof. The closed right i<strong>de</strong>al E := ℓB is, if we take B to act in the natural way from the right,<br />

a Hilbert B-mo<strong>du</strong><strong>le</strong> with inner pro<strong>du</strong>ct 〈a,b〉:=a∗b, and is countably generated (in the generalized<br />

sense <strong>de</strong>scribed above) with generators (ℓ1/n ) ∞ n=1 . Thus, by the above form of the Kasparov<br />

stabilization theorem (Theorem 4.3), there is a unitary U in L(E ⊕ HB, HB) imp<strong>le</strong>menting an<br />

isomorphism of E ⊕ HB and HB.<br />

Let P be the projection of E ⊕ HB onto the first factor, E. The projection T := UPU∗ ∈<br />

L(HB) has image isomorphic (by a unitary equiva<strong>le</strong>nce) to E, and thus by the <strong>de</strong>finition of the<br />

compact operators on a Hilbert mo<strong>du</strong><strong>le</strong>,<br />

K(T HB) ∼ = K(ℓB).<br />

Now, however, recalling the <strong>de</strong>finition of the compact operators on a Hilbert mo<strong>du</strong><strong>le</strong>, we see that<br />

K(ℓB) is generated by e<strong>le</strong>ments of the form ℓb1b ∗ 2 ℓ∗ , where the bi are in B. Hence (up to unitary<br />

equiva<strong>le</strong>nce), T K(HB)T ∼ = ℓBℓ ∗ . However, T is in L(HB) = M(B ⊗K), and K(HB) = B ⊗K,<br />

which thus comp<strong>le</strong>tes the proof of the first part of the <strong>le</strong>mma since B is stab<strong>le</strong>. For the fullness<br />

c<strong>la</strong>imed in the <strong>la</strong>st part of the theorem, notice that if we write L(E ⊕ HB, HB)P L(HB,E⊕ HB)<br />

as a formal two-by-two matrix, and consi<strong>de</strong>r the lower right-hand corner, we are to show that<br />

L(ℓB, HB)L(HB,ℓB)is <strong>de</strong>nse in L(HB). However,<br />

L(ℓB, HB)L(HB,ℓB)= L(B, HB)ℓ ∗ ℓL(HB,B)<br />

= L(B, HB)L(B)ℓ ∗ ℓL(B)L(HB,B)


400 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

and since ℓ ∗ ℓ is full in L(B) by hypothesis, the <strong>la</strong>st expression simplifies to L(B, HB)L(HB,B),<br />

which is equal to L(HB) if B is stab<strong>le</strong> and σ -unital. ✷<br />

Corol<strong>la</strong>ry 4.5. If D is a hereditary σ -unital subalgebra of the σ -unital stab<strong>le</strong> algebra B, then<br />

M(D) can be i<strong>de</strong>ntified with a corner insi<strong>de</strong> M(B).<br />

Last, we recall the <strong>de</strong>finition of properly infinite positive e<strong>le</strong>ments [12].<br />

Definition. A positive e<strong>le</strong>ment P in a C∗-algebra C is said to be properly infinite if P ⊕ P P<br />

in M2(C), where a b means that rnbr∗ n → a in norm for some sequence (rn).<br />

The main fact that we need about such e<strong>le</strong>ments is the following well-known result, <strong>du</strong>e to<br />

J. Cuntz in the simp<strong>le</strong> case, with generalizations by M. Rørdam (and possibly others).<br />

Lemma 4.6. A full properly infinite projection P has the property that xPx ∗ = 1 for some e<strong>le</strong>ment<br />

x.<br />

The <strong>le</strong>mma is established by noticing that if rnbr∗ n → 1 then rnbr∗ n is invertib<strong>le</strong> for some<br />

sufficiently <strong>la</strong>rge n.<br />

Proof of Theorem 3.2. We now show that (i) implies (ii). Thus, we assume that all weakly<br />

nuc<strong>le</strong>ar full extensions are absorbing, and prove quasi-invertibility. If P is a full multiplier projection,<br />

then if π(P) is equal to 1 in the corona, we can readily show, using Remark 3.3, that<br />

P is quasi-invertib<strong>le</strong> in the multipliers. We now have the case where P is not equal to 1 in<br />

the corona. The algebra C ∗ (1,P,B) can be regar<strong>de</strong>d as the extension algebra of some trivial<br />

full extension. Since this extension is absorbing by hypothesis (and is weakly nuc<strong>le</strong>ar because<br />

the quotient algebra is abelian, hence nuc<strong>le</strong>ar), the algebra C := C ∗ (1,P,B) is purely <strong>la</strong>rge.<br />

If Q = 1 is a multiplier projection equiva<strong>le</strong>nt to 1M(B), there is an obvious homomorphism<br />

δ : C ∗ (1, π(P )) → C ∗ (1,Q). We can thus regard δ as a homomorphism from C ∗ (1, π(P )) into<br />

the multipliers. The unital map δ ◦π : C → M(B) satisfies the hypotheses of Lemma 4.1, so there<br />

are isometries (vn) such that δ ◦ π(c) − v ∗ n cvn is in the canonical i<strong>de</strong>al B and, moreover, goes<br />

to zero pointwise as n →∞. In particu<strong>la</strong>r, Q − v ∗ n Pvn can be ma<strong>de</strong> arbitrarily small in norm.<br />

Since Q = WW ∗ for some multiplier isometry W , it follows that 1 − W ∗ v ∗ n PvnW is small in<br />

norm, implying that W ∗ v ∗ n PvnW is invertib<strong>le</strong>. This finishes the proof that (i) implies (ii).<br />

We now show that (ii) implies (i).<br />

By Theorem 3.1 it is sufficient to show that cBc contains a stab<strong>le</strong> full subalgebra for all positive<br />

full multiplier e<strong>le</strong>ments c ∈ M(B). By Theorem 4.4, it is sufficient to show this for the case<br />

where c is a full projection in the multipliers. By hypothesis, the projection c is quasi-invertib<strong>le</strong>,<br />

so that 1 = x ∗ cx, implying that cx is an isometry. But this isometry gives a homomorphism<br />

Ad(cx) : B → cBc, and the image of the homomorphism is the <strong>de</strong>sired stab<strong>le</strong> full subalgebra.<br />

The equiva<strong>le</strong>nce with the other conditions listed is straighforward, using Lemma 4.6 and Remark<br />

3.3. ✷<br />

5. An enveloping algebra criterion for absorption<br />

In this section we give a comp<strong>le</strong>tely different set of absorption criteria, applicab<strong>le</strong> to sing<strong>le</strong><br />

extensions rather than to the set of all full extensions. We shall find that every positive e<strong>le</strong>ment of


D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408 401<br />

the image of an absorbing extension is contained, up to unitary equiva<strong>le</strong>nce, in a canonical zerodimensional<br />

abelian subalgebra that we <strong>de</strong>note W . This has highly interesting consequences for<br />

the functional calculus, <strong>le</strong>ading us to a multivariab<strong>le</strong> functional calculus prob<strong>le</strong>m that we solve<br />

by means of a multivariab<strong>le</strong> BDF-type theorem.<br />

We shall need the following topological result, characterizing the Stone–Čech corona of the<br />

positive integers.<br />

Theorem 5.1. (Parovičenko [18,22]) A totally disconnected compact F-space without iso<strong>la</strong>ted<br />

points, having weight ℵ1, and such that every zero-set is a regu<strong>la</strong>r closed set, is homeomorphic<br />

to β(N)/N. Moreover, β(N)/N maps onto every compact space that has weight at most ℵ1.<br />

In this section, we specialize to stably unital separab<strong>le</strong> algebras. By Brown’s theorem [3],<br />

a separab<strong>le</strong> algebra is stably unital if and only if the stabilization contains a full projection. We<br />

may as well, therefore, re<strong>du</strong>ce to the case of B ⊗ K where B is a unital and separab<strong>le</strong> C ∗ -algebra.<br />

Define a map ¯δ : ℓ ∞ → M(B ⊗ K) by embedding ℓ ∞ in the natural way into 1 ⊗ M(K) ⊂<br />

M(B ⊗ K). C<strong>le</strong>arly this is a homomorphism, and composing with the natural quotient map<br />

π : M(B ⊗ K) → M(B ⊗ K)/B ⊗ K, we find that the kernel of π ◦ ¯δ is exactly the space of<br />

sequences converging to zero. Thus we have an injective homomorphism, <strong>de</strong>noted δ, from ℓ ∞ /c0<br />

to M(B ⊗ K)/B ⊗ K.<br />

Noticing that c0 can be thought of as C0(N), where the integers N have the usual discrete<br />

topology, we see that the space of characters of ℓ ∞ /c0 ∼ = M(C0(N))/C0(N) is β(N) \ N, where<br />

β(N) is the Stone–Čech compactification of the natural numbers. Hence, by the Gelfand theorem,<br />

δ can be i<strong>de</strong>ntified with a map from the highly nonseparab<strong>le</strong> algebra C(β(N) \ N) into M(B ⊗<br />

K)/B ⊗ K. The map ¯δ maps C(β(N)) into M(B).<br />

Let us henceforth <strong>de</strong>note the image of ¯δ by ¯W , and the image in the corona by W . Note,<br />

however, that we do not obtain an injective map of β(N) \ N into the multipliers. From the<br />

viewpoint of extension theory, we have an apparently nontrivial Busby map τ : β(N) \ N →<br />

M(B)/B. If we want an essential extension, there does not appear to be any way to make this<br />

extension trivial. (It is, in princip<strong>le</strong>, possib<strong>le</strong> that there is some other construction that does give<br />

a trivial essential extension τ : β(N) \ N → M(B)/B.)<br />

We point out that the usual lifting theorems that convert a Busby map into a comp<strong>le</strong>tely<br />

positive map into the multipliers all require separability of the source algebra. Therefore, it is<br />

difficult to show that we can represent τ by a comp<strong>le</strong>tely positive map into the multipliers un<strong>le</strong>ss<br />

we restrict to a separab<strong>le</strong> subalgebra of β(N) \ N. We return to the issue of prob<strong>le</strong>ms arising<br />

from nonseparability <strong>la</strong>ter. The next proposition studies the process of restriction to separab<strong>le</strong><br />

subalgebras. First a <strong>le</strong>mma (which is actually [7, Exercise 3.2.I]).<br />

Lemma 5.2. An unital abelian C ∗ -algebra has a countab<strong>le</strong> <strong>de</strong>nse subset if and only if the Gelfand<br />

spectrum has a countab<strong>le</strong> base as a topological space. The algebra has a <strong>de</strong>nse subset of cardinality<br />

at most ℵ1 if and only if the Gelfand spectrum has a base of cardinality at most ℵ1.<br />

Proof. Suppose that we are given a <strong>de</strong>nse subset D of C(X), where X is compact and Hausdorff.<br />

We are to show that there is a base of X of cardinality |D|. We c<strong>la</strong>im that the sets B(f ) := {x ∈ X:<br />

|f(x)| < 1}, with f ∈ D, form such a base. Given an open set O, choose a point x ∈ O. Since<br />

X is comp<strong>le</strong>tely regu<strong>la</strong>r [8], there is a function s : X →[0, 1] that is 0 at x and is 1 outsi<strong>de</strong> O.<br />

Approximate 2s within ɛ = 1/4byafunctionf ∈ D. It follows that B(f ) contains x and is itself<br />

contained in O. It is now c<strong>le</strong>ar that the B(f ) do form a base (of cardinality |D|).


402 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

For the converse, use comp<strong>le</strong>te regu<strong>la</strong>rity to find a base whose e<strong>le</strong>ments are co-zero sets (a cozero<br />

set is {x: f(x)>0} for some function f ) by shrinking the e<strong>le</strong>ments of the given base. This<br />

may lower the cardinality of the base, but cannot increase it. Thus, we have a family of functions<br />

supported on basis e<strong>le</strong>ments. The finite linear combinations with rational coefficients, and their<br />

finite pro<strong>du</strong>cts, form an algebra that is (by the Stone–Weierstrass theorem) <strong>de</strong>nse in C(X). Since<br />

we only allow multiplication by rationals plus the formation of finite pro<strong>du</strong>cts and sums, the<br />

cardinality of this algebra is boun<strong>de</strong>d by max{ℵ0, |B|}, where B is the base we started with. The<br />

bound on cardinality is obtained by use of Cantor’s result that |A × B| max{|A|, |B|} if one of<br />

the sets A or B is infinite. ✷<br />

In the next proposition, the restriction to injective Busby maps is necessary. Without injectivity,<br />

counter-examp<strong>le</strong>s can be found.<br />

Proposition 5.3. Given a unital abelian algebra A with a <strong>de</strong>nse subset of cardinality at most ℵ1,<br />

there is a commutative diagram<br />

C(β(N)) ∼ = ℓ ∞ (N)<br />

0<br />

<br />

B ⊗ K<br />

<br />

¯δ<br />

−→ M(B ⊗ K) → 0<br />

<br />

<br />

A → C(β(N) \ N) ∼ = ℓ∞ δ<br />

(N)/c0(N) −→ M(B ⊗ K)/B ⊗ K → 0<br />

The (Busby) map from A to M(B ⊗ K)/B ⊗ K <strong>de</strong>fined by this diagram is injective, purely <strong>la</strong>rge,<br />

and nuc<strong>le</strong>ar. It is trivial if and only if the Gelfand spectrum of A has a countab<strong>le</strong> <strong>de</strong>nse subset.<br />

Proof. By the Gelfand theorem and the previous proposition A = C(X) for some compact Hausdorff<br />

space X with a topological basis of cardinality ℵ1 or <strong>le</strong>ss.<br />

First, suppose the extension is trivial. There then is an injective unital homomorphism from A<br />

into the image of ¯δ in M(B ⊗K). Composing this injection with ¯δ −1 , we obtain a unital injection<br />

of A into C(β(N)), which by the Gelfand theorem corresponds to a mapping g of β(N) onto X.<br />

The inverse image g −1 (V ) of an open set V of X must intersect N since N is <strong>de</strong>nse in β(N),<br />

showing that g(N) is a countab<strong>le</strong> <strong>de</strong>nse subset of X.<br />

Conversely, if we assume that X has a countab<strong>le</strong> <strong>de</strong>nse subset, this countab<strong>le</strong> <strong>de</strong>nse subset<br />

gives us a continuous mapping h : N → X with <strong>de</strong>nse image in X. Using the fundamental property<br />

of Stone–Čech compactifications to extend h uniquely to a map β(h): β(N) → X, wesee<br />

that the image of β(h) is compact, hence closed, and <strong>de</strong>nse in X. Therefore the image is all of X,<br />

and the pullback of β(h) is thus an injective homomorphism of A into C(β(N)), which we then<br />

compose with ¯δ to obtain (the lifting of) a trivial Busby map. (The map β(h) is unique up to<br />

homeomorphism of β(N).)<br />

This comp<strong>le</strong>tes the proof for the separab<strong>le</strong> (or trivial) case. We now consi<strong>de</strong>r the nonseparab<strong>le</strong><br />

(or nontrivial) case. By the <strong>le</strong>mma and by Parovičenko’s theorem (Theorem 5.1), the space<br />

β(N)/N maps onto X, so we have a unital injection of A into C(β(N) \ N). (This map is in fact<br />

unique up to homeomorphism of β(N) \ N.) Composing with the map δ we have a injective map<br />

π


D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408 403<br />

τ into M(B)/B. Since A is commutative, the map is nuc<strong>le</strong>ar. Recall that the extension algebra<br />

C associated with the constructed Busby map, τ , is by <strong>de</strong>finition the pullback<br />

C −→ A<br />

τ<br />

<br />

<br />

π<br />

M(B ⊗ K) −→ M(B ⊗ K)/B ⊗ K<br />

Since τ has no kernel, we may as well regard C as being i<strong>de</strong>ntified with its image in M(B ⊗ K),<br />

and then by comparing with the above diagram, we see that the extension algebra is in fact a<br />

subalgebra of the image ¯W of ¯δ in M(B ⊗ K).<br />

We notice that the image algebra W is purely <strong>la</strong>rge: choosing some positive e<strong>le</strong>ment (λi) of<br />

ℓ ∞ that is not in c0,themap¯δ gives an e<strong>le</strong>ment of 1⊗B(H) that majorizes some nonzero multip<strong>le</strong><br />

of a projection P in the image of ¯δ. Since this projection can be taken to be nonzero on infinitely<br />

many λi, it is properly infinite. Thus, P is a properly infinite e<strong>le</strong>ment of M(B ⊗ K), and there<br />

exists x ∈ M(B ⊗ K) such that xPx ∗ = 1M(B⊗K). Since the given e<strong>le</strong>ment majorizes P up<br />

to a sca<strong>la</strong>r multip<strong>le</strong>, we can therefore find y ∈ M(B ⊗ K) such that y ¯δ((λi))y ∗ = 1. Therefore<br />

the hereditary subalgebra ¯δ((λi))(B ⊗ K)¯δ((λi)) contains a subalgebra that is stab<strong>le</strong> and full in<br />

B ⊗ K. Strictly speaking, we must also check the case of perturbation by e<strong>le</strong>ments of B ⊗ K,<br />

thus we should show a simi<strong>la</strong>r result for (¯δ((λi)) + b)(B ⊗ K)(¯δ((λi)) + b) where b ∈ B ⊗ K.<br />

But using stability we can easily find, as in Remark 3.3, a y ′ such that y ′ (¯δ((λi)) + b)y ′∗ =<br />

1M(B⊗K). ✷<br />

We now arrive at a very interesting absorption criterion. The criterion will be stated in terms<br />

of “copies of W ,” a term with several possib<strong>le</strong> interpretations, so <strong>le</strong>t us take a few moments to<br />

discuss this before continuing. A copy of W is simply the extension algebra of some (generally<br />

nontrivial) injective and purely <strong>la</strong>rge Busby map ℓ ∞ /c0 → M(B ⊗ K)/B ⊗ K. These extensions<br />

have the peculiar property, <strong>de</strong>scribed in Proposition 5.3, of having separab<strong>le</strong> subalgebras coming<br />

from trivial extensions, without being themselves trivial. We would naturally expect them to<br />

be absorbing, however, we have difficulty even lifting them to comp<strong>le</strong>tely positive maps into the<br />

multipliers, which would certainly be a necessary first step in showing this. Both the Choi–Effros<br />

and the Haagerup lifting theorems have separability of the source algebra as a hypothesis, and,<br />

moreover, the Elliott–Kucerovsky absorption theorem also requires separability. The work of<br />

Hadwin [9] suggests that fundamental differences between the separab<strong>le</strong> and nonseparab<strong>le</strong> cases<br />

are to be expected. As a result, we must <strong>le</strong>ave open for the moment the quite interesting question<br />

of whether copies of W are unitarily equiva<strong>le</strong>nt by multiplier unitaries. In Theorem 6.1 we prove<br />

that this is true if we restrict to finitely generated subalgebras (within given copies of W ), and<br />

this is sufficient for most applications: speaking loosely, one could say that there is a canonical<br />

copy of W after restriction to separab<strong>le</strong> subalgebras.<br />

Corol<strong>la</strong>ry 5.4. Let A be a separab<strong>le</strong> unital C ∗ -algebra and B be the stabilization of a unital<br />

separab<strong>le</strong> C ∗ -algebra. Let τ : A → M(B)/B be a trivial full extension. Then, τ is absorbing if<br />

and only if for each positive a ∈ A, the algebra τ(C ∗ (a)) is contained in a copy of W.<br />

Proof. It follows from the basic absorption criterion (Theorem 3.1) that if τ : A → M(B)/B is<br />

absorbing, then any restriction of τ to a subalgebra of A is absorbing. Thus, the restriction of τ to<br />

a subalgebra of the form C ∗ (a), with a being some positive nonzero e<strong>le</strong>ment, is absorbing. Being


404 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

still trivial it is thus unitarily equiva<strong>le</strong>nt to the trivial extension obtained from Proposition 5.3,<br />

and we have proven the “only if” part of the theorem.<br />

We prove the converse direction next, but in greater generality. ✷<br />

Theorem 5.5. Let A be a separab<strong>le</strong> unital C ∗ -algebra and B be the stabilization of a unital<br />

separab<strong>le</strong> C ∗ -algebra. Let τ : A → M(B)/B be a full extension, not necessarily trivial.<br />

Then, τ is absorbing if for each positive a ∈ A, the algebra τ(C ∗ (a)) is contained in a copy<br />

of W . In this case, the restricted extensions τ(C ∗ (a)) are necessarily trivial.<br />

Proof. Let us <strong>de</strong>note the restricted extensions by τa. It follows from Theorem 3.1 that τ is<br />

absorbing if and only if all the τa are.<br />

We first prove that each τa is, in fact, trivial. Choosing some particu<strong>la</strong>r τa, the image algebra<br />

D of τa is by hypothesis contained in W ∼ = ℓ ∞ /c0. Since the spectrum of a is some subset of<br />

[0, 1], the algebra D is singly generated. Letting g be the generator, we lift it to ¯g in ℓ ∞ .The<br />

spectrum of ¯g may be <strong>la</strong>rger than that of g by some countab<strong>le</strong> set, but we can find a function<br />

f : (0, 1]→(0, 1] such that f is the i<strong>de</strong>ntity map when restricted to the spectrum of g but maps<br />

the spectrum of ¯g to the spectrum of g. Applying this function, we thus have a lifting, ¯g, that<br />

is isospectral to g, and then Gelfand’s theorem gives us a splitting of τa. Now,τa is trivial,<br />

with a splitting map s : D → M(B) whose image is in a copy of ¯W. The absorption property<br />

is a property of the image of an extension in the corona. Therefore, τa is absorbing because the<br />

image is contained in a copy of W , or more exactly, is contained in the image of a copy of the<br />

extension of Proposition 5.3. ✷<br />

We remark that it is quite possib<strong>le</strong> for the restrictions of an extension, as in the above theorem,<br />

to be trivial without the who<strong>le</strong> extension being trivial. In such a case, we have local splitting<br />

homomorphisms that do not assemb<strong>le</strong> to give a globally <strong>de</strong>fined splitting homomorphism.<br />

Finally, one may inquire if the triviality condition can be dropped from Corol<strong>la</strong>ry 5.4.<br />

Corol<strong>la</strong>ry 5.6. Let A be a separab<strong>le</strong> unital C ∗ -algebra and B be the stabilization of a unital separab<strong>le</strong><br />

bootstrap c<strong>la</strong>ss C ∗ -algebra with K1(B) ={0}. Let τ : A → M(B)/B be a full extension.<br />

Then, τ is absorbing if and only if for each positive a ∈ A, the algebra τ(C ∗ (a)) is contained in<br />

a copy of W.<br />

Proof. Notice the spectrum of a is a closed subset of [0, 1]. It follows that K1(C ∗ (a)) ={0},<br />

and that K0(C ∗ (a)) is a free abelian group. Thus, by the UCT and the properties of countably<br />

generated abelian groups, we find that KK 1 (C ∗ (a), B) is zero, and therefore, the restrictions of<br />

τ are—if absorbing—always trivial extensions. We now have the c<strong>la</strong>imed result by combining<br />

the two previous results. ✷<br />

In summary, we have shown that:<br />

(i) For trivial extensions τ , absorption is equiva<strong>le</strong>nt to τ(C ∗ (a)) being contained (up to unitary<br />

equiva<strong>le</strong>nce) in W .<br />

(ii) For not necessarily trivial extensions, the above condition is still sufficient for absorption,<br />

but may not be necessary un<strong>le</strong>ss the K1-group of the canonical i<strong>de</strong>al is zero.


D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408 405<br />

It would, of course, be possib<strong>le</strong> to rep<strong>la</strong>ce the K-theoretical condition on the algebra,<br />

K1(B) ={0}, by a <strong>le</strong>ss restrictive but more technical KK-theoretic condition on τ .<br />

6. A multivariab<strong>le</strong> Brown–Doug<strong>la</strong>s–Fillmore theorem<br />

The absorption criteria of the previous section may seem at first g<strong>la</strong>nce to be of <strong>la</strong>rgely theoretical<br />

interest, but in fact they can be used to construct an exten<strong>de</strong>d functional calculus for<br />

suitab<strong>le</strong> e<strong>le</strong>ments of the corona, by using properties of zero-dimensional topological spaces to<br />

find e<strong>le</strong>ments of the enveloping algebra W having interesting properties. This exten<strong>de</strong>d functional<br />

calculus has been partially explored in [15]. As a consequence, the following question arises naturally:<br />

given (commuting) e<strong>le</strong>ments d1,...,dn from one copy of W , and e<strong>le</strong>ments v1,...,vn<br />

from another copy of W, when is C ∗ (d1,...,dn) unitarily equiva<strong>le</strong>nt to C ∗ (v1,...,vn)?<br />

We now proceed to resolve this question, which may be thought of in terms of the multivariab<strong>le</strong><br />

functional calculus: characterize unitary equiva<strong>le</strong>nce c<strong>la</strong>sses of n-tup<strong>le</strong>s from two different<br />

abelian subalgebras. We note that in some cases, W may in<strong>de</strong>ed be a masa.<br />

Thus, suppose that we are given two such n-tup<strong>le</strong>s, (d1,...,dn) and (v1,...,vn). The algebra<br />

D1 generated by the di is, by Gelfand’s theorem, isomorphic to some C(X) but is not usually<br />

singly generated. Since D1 is contained in some copy of W , and the Gelfand spectrum of W<br />

is totally disconnected, we can approximate each di by a finite sum of projections from W .<br />

Approximating di by an e<strong>le</strong>ment of this form, we have sin with sin − di < 1/n. Thesetof<br />

projections (pjin) used to form the sin is a countab<strong>le</strong> family, and the algebra D1 lies in the<br />

algebra generated by these projections. Re<strong>la</strong>bling the countab<strong>le</strong> family as (qm), we notice that by<br />

the Stone–Weierstrass theorem, the e<strong>le</strong>ment<br />

g :=<br />

∞<br />

1<br />

2qm − 1<br />

3 m<br />

generates the algebra C ∗ (pjin). (This argument is based on an i<strong>de</strong>a of von Neumann’s. Compare<br />

[20] and [1].) Denoting the generator by g, we notice that it has countab<strong>le</strong> spectrum. Thus, we<br />

may lift to ¯g in ¯W and apply a function to ¯g to in<strong>sur</strong>e isospectrality with g. It follows that<br />

there is, by Gelfand’s theorem, a splitting map s from C ∗ (pjin) to ¯W ⊂ M(B). Restricting the<br />

splitting map to D1, we see that there is a trivial (absorbing) extension τ : C(X) → M(B)/B<br />

with image exactly D1, and simi<strong>la</strong>rly for the D2 generated by the vi. We now have a theorem of<br />

Brown–Doug<strong>la</strong>s–Fillmore type, since two trivial absorbing extensions are unitarily equiva<strong>le</strong>nt if<br />

and only if the image algebras are isomorphic.<br />

Theorem 6.1. If (ui) and (di) are two tup<strong>le</strong>s of e<strong>le</strong>ments of two distinct copies of W , then the<br />

two abelian C ∗ -subalgebras of the corona, C ∗ (u1,...,un) and C ∗ (d1,...,dm), are unitarily<br />

equiva<strong>le</strong>nt by a multiplier unitary if and only if the Gelfand spectra of the two algebras are<br />

isomorphic.<br />

7. Fredholm trip<strong>le</strong>s<br />

For the rea<strong>de</strong>r’s convenience we first summarize the basic facts about the Fredholm mo<strong>du</strong><strong>le</strong><br />

picture of KK-theory. Since there are many equiva<strong>le</strong>nt ways to <strong>de</strong>fine KK 1 (A, B) in terms of<br />

Fredholm mo<strong>du</strong><strong>le</strong>s, and no one canonical choice, we follow the <strong>de</strong>finition in [2, Section 17.6.4],<br />

involving approximate (ungra<strong>de</strong>d) projections.


406 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

Definition. Let B be a σ -unital stab<strong>le</strong> C ∗ -algebra, and <strong>le</strong>t A be a separab<strong>le</strong> C ∗ -algebra. A stabilized<br />

ungra<strong>de</strong>d Fredholm cyc<strong>le</strong> in KK 1 (A, B) is a trip<strong>le</strong> (M(B), φ, F ) where φ is a homomorphism<br />

from A to M(B), and F ∈ M(B) is such that:<br />

(i) (F ∗ F − F)φ(a)∈ B;<br />

(ii) [F,φ(a)]∈B; and<br />

(iii) φ(a)(F − F ∗ ) ∈ B.<br />

A cyc<strong>le</strong> is said to be trivial if the above three expressions are actually zero.<br />

We can summarize the above <strong>de</strong>finition by saying that Fredholm trip<strong>le</strong>s are specified by a homomorphism<br />

φ and an operator that is an approximate projection: more specifically, an e<strong>le</strong>ment<br />

of Iφ that becomes a projection in Iφ/Jφ, where<br />

Iφ := m ∈ M(B): φ(a),m ∈ B for all a ∈ A ,<br />

Jφ := m ∈ M(B): φ(a)m∈ B for all a ∈ A .<br />

Sometimes, Fredholm trip<strong>le</strong>s in KK 1 (A, B) are specified by self-adjoint approximate unitaries<br />

instead, but this is equiva<strong>le</strong>nt since self-adjoint unitaries map to projections un<strong>de</strong>r the map<br />

u ↦→ 1 + u/2. For more information on the various pictures of KK-theory, and the interesting<br />

transformations that re<strong>la</strong>te one to the other, see [2].<br />

In some applications, cyc<strong>le</strong>s arise which are not stabilized, meaning that M(B) is rep<strong>la</strong>ced<br />

by the C ∗ -algebra of adjointab<strong>le</strong> operators on some given Hilbert B-mo<strong>du</strong><strong>le</strong>, but it can be shown<br />

that un<strong>de</strong>r an appropriate <strong>de</strong>finition of equiva<strong>le</strong>nce unstabilized cyc<strong>le</strong>s are neverthe<strong>le</strong>ss always<br />

equiva<strong>le</strong>nt to stabilized cyc<strong>le</strong>s.<br />

There are a number of apparently quite different equiva<strong>le</strong>nce re<strong>la</strong>tions on KK 1 -cyc<strong>le</strong>s, and<br />

the one most re<strong>le</strong>vant to our situation is given by “compact” perturbation, addition of <strong>de</strong>generate<br />

cyc<strong>le</strong>s, and unitary equiva<strong>le</strong>nce by multiplier unitaries. (By “compact” perturbation is meant<br />

perturbation of the operator F by e<strong>le</strong>ments of Jφ.)<br />

Definition. We say that a Fredholm trip<strong>le</strong> (M(B), φ, F ) is unital if and only φ(1) = 1 + b and<br />

F = 1 + b ′ for some b,b ′ ∈ B.<br />

C<strong>le</strong>arly this is an extremely strong property. It is, however, of interest in that it p<strong>la</strong>ys a ro<strong>le</strong> in<br />

the <strong>de</strong>finition of the absorption property for a Fredholm trip<strong>le</strong>.<br />

Definition. A unital Fredholm trip<strong>le</strong> F := (M(B), φ, F ) is absorbing if F ⊕ T is unitarily<br />

equiva<strong>le</strong>nt to F for all unital trivial cyc<strong>le</strong>s T . A nonunital Fredholm cyc<strong>le</strong> F is absorbing if it<br />

has this property for all trivial cyc<strong>le</strong>s T .<br />

One observes that there is a natural isomorphism<br />

KK 1 (A, B) → Ext(A, B),<br />

(φ, F ) ↦→ π(FφF ∗ )


D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408 407<br />

and that the equiva<strong>le</strong>nce re<strong>la</strong>tion on Ext exactly corresponds to the stabilized cp equiva<strong>le</strong>nce<br />

re<strong>la</strong>tion in KK 1 , meaning equiva<strong>le</strong>nce mo<strong>du</strong>lo perturbation by Jφ, unitary equiva<strong>le</strong>nce, and addition<br />

of stabilized <strong>de</strong>generate cyc<strong>le</strong>s [2, Section 17.6.4] which may be taken to be unital if the<br />

original cyc<strong>le</strong> is unital [11]. Addition of stabilized <strong>de</strong>generate cyc<strong>le</strong>s corresponds to the addition<br />

of trivial extensions in Ext, so we see that for absorbing cyc<strong>le</strong>s (φ1,F1) and (φ2,F2), wehave<br />

that (φ1,F1) ∼cp (φ2,F2) in KK 1 if and only if π(F1φ1F ∗ 1 ) ∼u π(F2φ2F ∗ 2 ).<br />

We now give an examp<strong>le</strong> showing that a cyc<strong>le</strong> can be absorbing even if (1 − F)φ(·)(1 − F)<br />

is small.<br />

Examp<strong>le</strong> 7.1. Let φ be the Kasparov GNS representation of A on B: thus, φ : A → M(B ⊗ K)<br />

is given by 1 ⊗ π where π is the usual universal representation of A on B(H) ∼ = M(K). Then it<br />

follows from Kasparov’s results [11] that (M(B), φ, 1) ∈ KK 1 (A, B) is an absorbing trip<strong>le</strong>. The<br />

comp<strong>le</strong>ment (M(B), φ, 0) is zero. We can e<strong>la</strong>borate this examp<strong>le</strong> somewhat, by <strong>le</strong>tting F = e11<br />

in M2(M(B ⊗ K)), and then <strong>le</strong>tting φ be the direct sum of Kasparov’s GNS representation in<br />

the e11 corner and any arbitrary homomorphism in the e22 corner.<br />

It would be <strong>de</strong>sirab<strong>le</strong>, of course, to give e<strong>le</strong>gant necessary and sufficient condition for a<br />

specific trip<strong>le</strong> to be absorbing. This is a hard prob<strong>le</strong>m, however, we point out one very pretty<br />

sufficient condition. An operator F ∈ M(B) is said to be quasi-Fredholm if it is quasi-invertib<strong>le</strong><br />

in the corona, or equiva<strong>le</strong>ntly, if there are x,y ∈ M(B) such that xFy = 1 + b for some b ∈ B.<br />

Theorem 7.2. Let B be stab<strong>le</strong> and separab<strong>le</strong>. A Fredholm trip<strong>le</strong> (M(B), φ, F ) ∈ KK 1 (A, B) is<br />

absorbing if F φ(a)F ∗ is a quasi-Fredholm operator whenever it is positive.<br />

Proof. Let C be the extension algebra of the associated Busby map a ↦→ F φ(a)F ∗ . A positive<br />

e<strong>le</strong>ment of C not in B is of the form F φ(a)F ∗ + b for some a ∈ A and b ∈ B, where F is the<br />

given essential projection from the Fredholm trip<strong>le</strong>.<br />

The hypothesis implies that there is r1 and r2 in the corona such that d := F φ(a)F ∗ satisfies<br />

r1dr2 = 1Q. But then, 1 r1d 2 r ∗ 1 r2r ∗ 2 , so that there is an r′ with r ′ d 2 r ′∗ = 1. Lifting r ′ to<br />

˜r ′ in the multipliers, we have that the original e<strong>le</strong>ment d + b satisfies ˜r ′ (d + b) 2 ˜r ′∗ = 1 + b0<br />

for some b0 in the canonical i<strong>de</strong>al, and we can cut down by a isometry obtained from stability,<br />

obtaining v ∗ ˜r ′ (d + b) 2 ˜r ′∗ v = 1 + v ∗ b0v with the norm of v ∗ b0v small. Thus, this expression is<br />

invertib<strong>le</strong>, and r ′′ (d + b) 2 r ′′∗ = 1forsomer ′′ in the multipliers. We now have an isometry V :=<br />

(d + b)r ′′∗ that imp<strong>le</strong>ments an equiva<strong>le</strong>nce of VBV ∗ ⊂ (d + b)B(d + b) and B, showing that<br />

(d + b)B(d + b) contains a stab<strong>le</strong> (full) hereditary subalgebra. Thus, the absorption criterion of<br />

Theorem 3.1 holds. Actually, the conclusion of that theorem implicitly has two cases, according<br />

to whether the given extension is unital or not, however, in both cases the hypothesis just involves<br />

existence of stab<strong>le</strong> full hereditary subalgebras as shown above. ✷<br />

References<br />

[1] Akemann, Newberger, Physical states on a C ∗ -algebra, Proc. Amer. Math. Soc. 40 (1973) 500.<br />

[2] B. B<strong>la</strong>ckadar, K-Theory for Operator Algebras, second ed., Cambridge Univ. Press, 1998.<br />

[3] L.G. Brown, Stab<strong>le</strong> isomorphism of hereditary subalgebras of C ∗ -algebras, Pacific J. Math. 71 (1977) 335–348.<br />

[4] L.G. Brown, R.G. Doug<strong>la</strong>s, P.A. Fillmore, Unitary equiva<strong>le</strong>nce mo<strong>du</strong>lo the compact operators and extensions of<br />

C ∗ -algebras, in: Proc. Conf. on Operator Theory, Amer. Math. Soc., in: Lecture Notes in Math., vol. 345, Springer,<br />

1973, pp. 58–128.<br />

[5] R.C. Busby, Doub<strong>le</strong> Centralizers and extensions of C ∗ -algebras, Trans. Amer. Math. Soc. 132 (1968) 79–99.


408 D. Kucerovsky / Journal of Functional Analysis 236 (2006) 395–408<br />

[6] G.A. Elliott, D. Kucerovsky, An abstract Brown–Doug<strong>la</strong>s–Fillmore absorption theorem, Pacific J. Math. 198 (2001)<br />

385–409.<br />

[7] R. Engelking, Intro<strong>du</strong>ction to General Topology, second ed., Sigma Ser. Pure Math., vol. 6, Hel<strong>de</strong>rmann, 1989.<br />

[8] L. Gillman, M. Jerison, Rings of Continuous Functions, van Nostrand, 1960.<br />

[9] D. Hadwin, Comp<strong>le</strong>tely positive maps and approximate equiva<strong>le</strong>nce, Indiana Univ. Math. J. 36 (1987) 211–228.<br />

[10] G.G. Kasparov, Hilbert C ∗ -mo<strong>du</strong><strong>le</strong>s: Theorems of Stinespring and Voicu<strong>le</strong>scu, J. Operator Theory 4 (1980) 133–<br />

150.<br />

[11] G.G. Kasparov, The operator K-functor and extension of C ∗ -algebras, Tr. Math. USSR Izv. 16 (1981) 513–636.<br />

[12] E. Kirchberg, M. Rørdam, Nonsimp<strong>le</strong> purely infinite C ∗ -algebras, preprint, 1999.<br />

[13] D. Kucerovsky, The nonsimp<strong>le</strong> absorption theorem, preprint, 2000.<br />

[14] D. Kucerovsky, Extensions contained in i<strong>de</strong>als, Trans. Amer. Math. Soc. (E<strong>le</strong>ctronic, posted August 25, 2003).<br />

[15] D. Kucerovsky, Commutative subalgebras of the corona, Proc. Amer. Math. Soc. 132 (10) (2004) 3027–3034.<br />

[16] D. Kucerovsky, P.W. Ng, The corona factorization property and approximate unitary equiva<strong>le</strong>nce, Houston J. Math.,<br />

in press.<br />

[17] D. Kucerovsky, P.W. Ng, Decomposition rank and absorbing extensions of type I algebras, J. Funct. Anal. 221 (1)<br />

(2005) 25–36.<br />

[18] I.I. Parovičenko, On a universal bicompactum of weight ℵ, Soviet Math. 4 (1963) 592–595.<br />

[19] I. Raeburn, S.J. Thompson, Countably generated Hilbert mo<strong>du</strong><strong>le</strong>s, the Kasparov stabilization theorem, and frames<br />

with Hilbert mo<strong>du</strong><strong>le</strong>s, Proc. Amer. Math. Soc. 131 (2003) 1557–1564.<br />

[20] Rickart, General Theory of Banach Algebras, Van Nostrand, Princeton, 1960.<br />

[21] M. Rørdam, On stab<strong>le</strong> C ∗ -algebras, preprint.<br />

[22] R.G. Woods, Co-absolutes of remain<strong>de</strong>rs of Stone–Čech compactifications, Pacific J. Math. 37 (1971) 545–560.<br />

Further reading<br />

[23] I.D. Berg, An extension of the Weyl–von Neumann theorem to normal operators, Trans. Amer. Math. Soc. 160<br />

(1971) 365–371.<br />

[24] M.D. Choi, E.G. Effros, The comp<strong>le</strong>tely positive lifting prob<strong>le</strong>m for C ∗ -algebras, Ann. of Math. (2) 104 (1976)<br />

585–609.<br />

[25] P.J. Cohen, Factorization in group algebras, Duke Math. J. 26 (1959) 199–205.<br />

[26] J. Cuntz, N. Higson, Kuiper’s theorem for Hilbert mo<strong>du</strong><strong>le</strong>s, in: Contemp. Math., vol. 62, Amer. Math. Soc., 1987,<br />

pp. 429–434.<br />

[27] N. Higson, A characterization of KK-theory, Pacific J. Math. 126 (1987) 253–276.<br />

[28] J. Hjelmborg, M. Rørdam, On stability of C ∗ -algebras, J. Funct. Anal. 155 (1998) 153–170.<br />

[29] D. Kucerovsky, Kasparov pro<strong>du</strong>cts in KK-theory and unboun<strong>de</strong>d operators with applications to in<strong>de</strong>x theory, thesis,<br />

University of Oxford (Magd.), 1995.<br />

[30] D. Kucerovsky, Extensions of C ∗ algebras with the corona factorization property, Int. J. Pure Appl. Math. 16 (2)<br />

(2004) 181–191.<br />

[31] J.A. Mingo, K-theory and multipliers of stab<strong>le</strong> C ∗ -algebras, Trans. Amer. Math. Soc. 299 (1987) 397–411.<br />

[32] G.K. Pe<strong>de</strong>rsen, C ∗ -Algebras and Their Automorphism Groups, Aca<strong>de</strong>mic Press, 1979.<br />

[33] M. Rørdam, A simp<strong>le</strong> C ∗ -algebra with a finite and an infinite projection, preprint, 2002.<br />

[34] K. Thomsen, On absorbing extensions, Proc. Amer. Math. Soc. 129 (2001) 1409–1417.<br />

[35] D.V. Voicu<strong>le</strong>scu, A non-commutative Weyl–von Neumann theorem, Rev. Roumaine Math. Pures Appl. 21 (1976)<br />

97–113.<br />

[36] J. von Neumann, Charakterisierung <strong>de</strong>s Spektrums eines Integraloperators, Actualites Sci. In<strong>du</strong>st., vol. 229, Hermann,<br />

Paris, 1935.<br />

[37] R.C. Walker, The Stone–Čech Compactification, Springer, 1974.<br />

[38] N.E. Wegge-Olsen, K-Theory and C ∗ -Algebras, Oxford Univ. Press, Oxford, 1993.


Journal of Functional Analysis 236 (2006) 409–446<br />

www.elsevier.com/locate/jfa<br />

Second or<strong>de</strong>r initial boundary-value prob<strong>le</strong>ms of<br />

variational type<br />

Denis Serre<br />

Éco<strong>le</strong> Norma<strong>le</strong> Supérieure <strong>de</strong> Lyon, France 1<br />

Received 29 August 2005; accepted 26 February 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 18 April 2006<br />

Communicated by H. Brezis<br />

Abstract<br />

We consi<strong>de</strong>r linear hyperbolic boundary-value prob<strong>le</strong>ms for second or<strong>de</strong>r systems, which can be written<br />

in the variational form δL = 0, with<br />

<br />

|∂t<br />

L[u]:= u| 2 − W(x;∇xu) dx dt,<br />

F ↦→ W(x; F)being a quadratic form over Md×n(R). The domain of L is the homogeneous Sobo<strong>le</strong>v space<br />

H˙ 1 (Ω × Rt ) n , with Ω either a boun<strong>de</strong>d domain or a half-space of Rd . The boundary condition inherent<br />

to this prob<strong>le</strong>m is of Neumann type. Such prob<strong>le</strong>ms arise for instance in linearized e<strong>la</strong>sticity. When Ω is<br />

a half-space and W <strong>de</strong>pends only on F , we show that the strong well-posedness occurs if, and only if, the<br />

stored energy<br />

<br />

W(∇xu) dx<br />

Ω<br />

is convex and coercive over ˙<br />

H 1 (Ω) n . Here, the energy <strong>de</strong>nsity W does not need to be convex but only<br />

strictly rank-one convex. The “only if” part is the new result. A remarkab<strong>le</strong> fact is that the c<strong>la</strong>ssical characterization<br />

of well-posedness by the Lopatinskiĭ condition needs only to be satisfied at real frequency pairs<br />

(τ, η) with τ 0, instead of pairs with ℜτ 0. Even stronger is the fact that we need only to examine pairs<br />

(τ = 0,η), and prove that some Hermitian matrix H(η)is positive <strong>de</strong>finite. Another significant result is that<br />

E-mail address: <strong>de</strong>nis.serre@umpa.ens-lyon.fr.<br />

1 UMPA (UMR 5669 CNRS), ENS <strong>de</strong> Lyon, 46, allée d’Italie, F-69364 Lyon, ce<strong>de</strong>x 07, France. The research of the<br />

author was partially supported by the European IHP project “HYKE”, contract # HPRN-CT-2002-00282.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.02.020


410 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

every such well-posed prob<strong>le</strong>m admits a pair of <strong>sur</strong>face waves at every frequency η = 0. These waves often<br />

have finite energy, like the Ray<strong>le</strong>igh waves in e<strong>la</strong>sticity. When we vary the <strong>de</strong>nsity W so as to reach nonconvex<br />

stored energies, this pair bifurcates to yield a Hadamard instability. This instability may occur for<br />

some energy <strong>de</strong>nsities that are quasi-convex, contrary to the case of the pure Cauchy prob<strong>le</strong>m, as shown in<br />

several examp<strong>le</strong>s. At the bifurcation, the corresponding stationary boundary-value prob<strong>le</strong>m enters the c<strong>la</strong>ss<br />

of ill-posed prob<strong>le</strong>ms in the sense of Agmon, Douglis and Nirenberg. For boun<strong>de</strong>d domains and variab<strong>le</strong><br />

coefficients, we show that the strong well-posedness is equiva<strong>le</strong>nt to a Korn-like inequality for the stored<br />

energy.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Initial boundary-value prob<strong>le</strong>m; Lopatinskiĭ condition; Convex energy<br />

1. Intro<strong>du</strong>ction<br />

Let W be a quadratic form on the space Md×n(R) of d × n matrices with real entries. For<br />

the sake of simplicity, we assume in this intro<strong>du</strong>ction that the spatial domain is a half-space:<br />

Ω := Rd−1 × (0, +∞). Boun<strong>de</strong>d domains will be consi<strong>de</strong>red in Section 7. We consi<strong>de</strong>r the<br />

stored energy functional<br />

<br />

W[u]:= W(∇xu) dx<br />

Ω<br />

for a field u : Ω → R n . In the context of Calculus of Variation, this energy is associated to a<br />

second or<strong>de</strong>r differential operator P : H 1 (Ω) n → H −1 (Ω) n <strong>de</strong>fined by<br />

(Pu)j =−<br />

d<br />

α=1<br />

∂α<br />

∂W<br />

∂Fαj<br />

and to the Neumann-type boundary operator B, <strong>de</strong>fined by<br />

(Bu)j := ∂W<br />

∂Fdj<br />

<br />

(∇xu) , j = 1,...,n,<br />

(∇xu), j = 1,...,n.<br />

Recall that when Pu ∈ L 2 (Ω) n , then Bu is in H −1/2 (∂Ω) n .<br />

We are concerned in this paper with the evolution boundary-value prob<strong>le</strong>m (IBVP) associated<br />

with the Lagrangian<br />

where<br />

<br />

L[u]:=<br />

R<br />

<br />

Ek[u]:=<br />

is the kinetic energy. Such an IBVP has the form<br />

Ek[u]−W[u] dt,<br />

Ω<br />

1<br />

2 |∂tu| 2 dx


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 411<br />

∂ 2 t<br />

u + Pu = f (x ∈ Ω, t ∈ R), (1)<br />

Bu = g (x ∈ ∂Ω, t ∈ R) (2)<br />

with P and B as above.<br />

In practical situations, W is not quadratic and Ω is a general domain, say a boun<strong>de</strong>d one.<br />

However, it is well known from the works by Kreiss [10], Sakamoto [15] and Majda [12] that the<br />

well-posedness of a general IBVP is intimately re<strong>la</strong>ted to that of IBVPs with frozen coefficients<br />

in half-spaces, obtained by linearization about uniform states. This exp<strong>la</strong>ins the re<strong>le</strong>vance of<br />

prob<strong>le</strong>m (1), (2) where P and B are linear with constant coefficients and Ω is a half-space. Such<br />

a prob<strong>le</strong>m is a building block in the general theory. In general, passing from the constant coefficient<br />

case to the variab<strong>le</strong> coefficient case requires much involved tools, like pseudo-differential<br />

estimates, symbolic symmetrizers and so on. This passage will be much simp<strong>le</strong>r in the situation<br />

studied here, as shown in Section 7.<br />

We point out that the right-hand si<strong>de</strong> f,g have a variational origin, if we add integrals<br />

<br />

R<br />

<br />

Ω<br />

f · udxdt and<br />

<br />

<br />

R ∂Ω<br />

g · udydt<br />

to the Lagrangian. As suggested by these expressions, we <strong>de</strong>note by x the space variab<strong>le</strong> interior<br />

in Ω and by y the variab<strong>le</strong> along the boundary ∂Ω. Therefore x = (y, xd).<br />

Throughout this paper, we make the minimal assumption that the wave-like operator ∂2 t + P<br />

is hyperbolic. 2 This amounts to saying that W is strictly rank-one convex, which means that for<br />

every F in Md×n(R),<br />

(rk F = 1) ⇒ W(F)>0 . (3)<br />

If W is convex, then it is obviously rank-one convex. Rank-one convexity amounts to saying<br />

that the stored energy W is convex over H˙ 1 (Rd ) n , when the domain is the who<strong>le</strong> space Rd instead of a half-space. Let us recall that W is polyconvex3 if it is the sum of a convex quadratic<br />

form W0 and of a form that vanishes on the cone of rank-one matrices. The <strong>la</strong>tter forms are<br />

cal<strong>le</strong>d (quadratic) null-Lagrangians or null-forms. Such null-forms do not contribute to the stored<br />

energy if Ω = Rd , but contribute through boundary integrals when Ω is a general domain. In<br />

other words, a null-form does not contribute to the operator P, but contributes to the boundary<br />

operator B. Quadratic null-forms are linear combinations of 2 × 2 minors of F . Polyconvexity<br />

obviously implies rank-one convexity, and the converse is true if and only if min{d,n} 2(see<br />

[16,18]).<br />

When the stored energy W is convex and the boundary condition is homogeneous (that is<br />

g ≡ 0), then the IBVP admits a convex total energy<br />

E = Ek + W.<br />

2 Some specialists say strongly hyperbolic.<br />

3 The <strong>de</strong>finition of polyconvexity, <strong>du</strong>e to J. Ball, is more e<strong>la</strong>borated for general (non-quadratic) functions W .


412 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

When W is coercive on ˙<br />

H 1 (Ω), the IBVP has the abstract form<br />

dU<br />

dt + AU = ˜<br />

f, U := (∂tu, ∇xu),<br />

where A is a maximal 4 monotone operator over L 2 (Ω) (d+1)n , this space being endowed with<br />

the norm in<strong>du</strong>ced by E. Therefore the homogeneous initial boundary-value prob<strong>le</strong>m (IBVP) is<br />

well-posed, in the sense that −A generates a continuous <strong>semi</strong>-group of contractions with respect<br />

to the norm E. Actually, because of conservativity, it generates a group of E-isometries (St)t∈R.<br />

Given a in the domain of A, t ↦→ Sta =: U(t) is the unique strong solution of U ′ + AU = 0<br />

such that U(0) = a. Then St is exten<strong>de</strong>d to L 2 by <strong>de</strong>nsity, thanks to the contractivity. When f is<br />

non-zero, the homogeneous IBVP is solved through the Duhamel’s princip<strong>le</strong><br />

<br />

U(t)= StU(0) +<br />

This solution satisfies the well-known a priori estimate<br />

e −2γT ∇x,tu(T ) 2<br />

L 2 + γ<br />

C<br />

0<br />

T<br />

∇x,tu(0) 2<br />

L 2 + 1<br />

γ<br />

0<br />

t<br />

St−τ ˜ f(τ)dτ.<br />

e −2γt ∇x,tu(t) 2<br />

L 2 dt<br />

T<br />

0<br />

e −2γtf(t) 2 L2 <br />

dt , (4)<br />

where C is a finite number, in<strong>de</strong>pen<strong>de</strong>nt of either (u, f ), orγ>0 and T>0. We point out that<br />

this estimate shares the scaling invariance of the IBVP.<br />

Droping the convexity/coercivity assumption for W, we say that the homogeneous IBVP is<br />

strongly well-posed if it satisfies an estimate of the form (4). This is the same inequality than that<br />

consi<strong>de</strong>red by Kreiss or Sakamoto, except for the boundary terms, which must be absent in our<br />

homogeneous case. The main goal of this paper is to characterize these variational prob<strong>le</strong>ms that<br />

are strongly well-posed. We show that they are precisely those for which the stored energy W is<br />

coercive over H˙ 1 (Ω).<br />

Of course, this does not mean that W be convex over Md×n(R). Therefore there is a need of<br />

a practical tool in or<strong>de</strong>r to characterize the <strong>de</strong>nsities W that yield strongly well-posed IBVPs. In<br />

the context of general hyperbolic IBVPs, the appropriate concept is that of Lopatinskiĭ condition.<br />

This is an algebraic property, which must be checked at every non-zero frequency pair (τ, η),<br />

with ℜτ 0(theuniform Lopatinskiĭ condition) and η ∈ Rd−1 . The situation turns out to be<br />

much simp<strong>le</strong>r in our variational context: we show that it is sufficient to check the Lopatinskiĭ<br />

condition at real pairs (τ, η) ∈ Rd \{0, 0}. We find an even simp<strong>le</strong>r characterization in terms of<br />

the simp<strong>le</strong> mo<strong>de</strong>s of the stationary equation: if η ∈ Rd−1 , the equation<br />

P e iη·y v(xd) = 0<br />

4 There are several proofs of maximality. The simp<strong>le</strong>st one uses Lax–Milgram theorem.


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 413<br />

is a second-or<strong>de</strong>r ODE whose solution space has dimension 2n. Ifη = 0, the subspace of solutions<br />

that vanish at +∞ (stab<strong>le</strong> solutions) has dimension n and is characterized by a first-or<strong>de</strong>r<br />

ODE v ′ = P(0,η)v, where the zero refers to τ = 0 and P(0,η) is the “stab<strong>le</strong>” solution of a<br />

quadratic matrix equation. To P(0,η), we associate by an explicit formu<strong>la</strong> a Hermitian matrix<br />

H(η). Then the IBVP is strongly well-posed if, and only if, H(η) is positive <strong>de</strong>finite for<br />

every η = 0.<br />

When the boundary data g is non-zero, the situation is not so nice. On the one hand, the<br />

general IBVP does not fall into a <strong>semi</strong>-group framework. On the other hand, we usually ask for<br />

additional boundary terms in estimate (4), both in right-hand si<strong>de</strong> (where g enters) and in <strong>le</strong>fthand<br />

si<strong>de</strong> (where the trace of ∇x,tu is estimated). We show here that these strong estimates à <strong>la</strong><br />

Kreiss [10] and Sakamoto [15] never hold in our variational framework, <strong>du</strong>e to the presence of<br />

<strong>sur</strong>face waves.<br />

Let us consi<strong>de</strong>r for instance isotropic e<strong>la</strong>sticity, where d 3 and the energy is given by<br />

W(F)= λ<br />

F + F<br />

4<br />

T 2 + μ<br />

2 (Tr F)2 . (5)<br />

The parameters λ and μ must satisfy λ>0 and 2λ + μ>0 for strict rank-one convexity. If<br />

moreover, λ + μ>0, the IBVP admits <strong>sur</strong>face waves (cal<strong>le</strong>d Ray<strong>le</strong>igh waves in this case).<br />

As mentioned above, such waves are incompatib<strong>le</strong> with a strong well-posedness for the nonhomogeneous<br />

IBVP (1), (2). We shall see that the situation is even worse if λ + μ


414 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

2. Convex stored energies<br />

The ro<strong>le</strong> of convex stored energies is fundamental in the study of the homogeneous IBVP. As<br />

mentioned in the intro<strong>du</strong>ction, there is a ba<strong>la</strong>nce <strong>la</strong>w for the total energy:<br />

∂t<br />

<br />

1<br />

2 |∂tu| 2 <br />

+ W(∇xu) −<br />

d<br />

n<br />

∂α<br />

α=1 j=1<br />

<br />

∂W<br />

∂tuj<br />

∂Fαj<br />

<br />

(∇xu) = (∂tu) · f. (6)<br />

When W is coercive on ˙<br />

H 1 (Ω) n , the well-posedness is en<strong>sur</strong>ed by the Hil<strong>le</strong>–Yosida theorem and<br />

Duhamel’s formu<strong>la</strong>, as long as g ≡ 0. C<strong>le</strong>arly, coerciveness implies strict convexity, although the<br />

converse is not true.<br />

The addition of a null-form to W modifies the stored energy W in general, because the integral<br />

of the minor ∂<strong>du</strong>j ∂αuk − ∂<strong>du</strong>k∂αuj over Ω equals a boundary integral. Such an addition,<br />

although <strong>le</strong>aving the differential operator P unchanged, does modify the boundary operator B.<br />

For this reason, when <strong>de</strong>aling with a specific IBVP, we are only free to add a tangential null-form<br />

(TNF) to W . This terminology <strong>de</strong>signates the linear combinations of minors Fαj Fβk − FαkFβj<br />

when 1 α


The assumption tells us that the form<br />

D. Serre / Journal of Functional Analysis 236 (2006) 409–446 415<br />

Q(G) + 2φ(G)· Z +|Z| 2<br />

takes non-negative values when Z ∈ R l and G has rank one. This amounts to saying that the<br />

quadratic form<br />

Q(G) − φ(G) 2<br />

is non-negative over the cone of rank-one matrices of size (d − 1) × n. Since we have min{d −<br />

1,n} 2, this implies ([16, Theorem 2.3], see also [18]) that there exists a null-form Q0 over<br />

M(d−1)×n(R) such that Q+(G) := Q(G) −|φ(G)| 2 + Q0(G) is non-negative. Hence the form<br />

is non-negative. ✷<br />

W(F)+ Q0(G) = Q+(G) + φ(G)+ p(Y) 2<br />

Examp<strong>le</strong>. Let us consi<strong>de</strong>r the energy of an isotropic e<strong>la</strong>stic material, given by (5), which is<br />

convex if and only if λ 0 and 2λ + μd 0, a condition that <strong>de</strong>pends on the dimension. Rankone<br />

convexity holds if and only if λ 0 and 2λ + μ 0. It is easy to see that in this <strong>la</strong>tter case,<br />

there exists a null-form Q0 such that W + Q0 is convex; however, this fact is use<strong>le</strong>ss when the<br />

physical domain has a non-trivial boundary, as in our case. The meaningful statement is that<br />

when λ 0 and λ + μ 0 (an intermediate assumption if d 3)), W(G,Y) is non-negative<br />

when G has rank one, and therefore there exists a tangential null-form that can be ad<strong>de</strong>d to W to<br />

make it convex. For instance, in the extreme case μ =−λ (say that λ = 2), then<br />

W(F)= 1<br />

T<br />

F + F<br />

2<br />

2 − 2(Tr F) 2<br />

= (F12 − F21) 2 + (F23 + F32) 2 + (F13 + F31) 2 + (F33 − F11 − F22) 2<br />

+ 4(F12F21 − F11F22). (7)<br />

The <strong>la</strong>st term of the right-hand si<strong>de</strong> is a TNF, and the rest is non-negative.<br />

2.2. Convex stored energies in general<br />

For more general energy <strong>de</strong>nsities, it may happen that W is convex, even though W is not<br />

convexifiab<strong>le</strong> in the sense given above. To study the convexity of W in a systematic way, we<br />

perform a Fourier transform v = Fyu with respect to the tangential variab<strong>le</strong>s. This has the effect<br />

to <strong>de</strong>coup<strong>le</strong> the tangential frequencies. The following <strong>le</strong>mma is obvious, except for the notations:<br />

we extend W to Mn(C) as a sesquilinear form. Thus W keeps real values. Additionally, we use<br />

the same <strong>de</strong>composition F = (G, Y ) as above.


416 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

Lemma 2.1. The stored energy W is convex on H 1 (Ω) n if, and only if, the following onedimensional<br />

quadratic functional Iη is non-negative over H 1 (0, +∞) n , for every vector η ∈<br />

R d−1 :<br />

Notice that when v = Fyu, then<br />

<br />

Iη[v]:=<br />

+∞<br />

0<br />

W(iη⊗ v,v ′ )dxd.<br />

Fy(∇yu) = iη ⊗ v, Fy(∂<strong>du</strong>) = v ′ .<br />

We are thus <strong>le</strong>d to the study of the convexity over H 1 (0, +∞) n of functionals of the form<br />

I[v]:= 1<br />

2<br />

<br />

+∞<br />

0<br />

w(v,v ′ )dxd, (8)<br />

where w : C n × C n → R is a sesquilinear form. In addition, the <strong>de</strong>nsities w satisfy<br />

∃ɛ >0 s.t. w(v,iξv) ɛ |η| 2 + ξ 2 |v| 2 , (9)<br />

because of uniform rank-one convexity. Notice that if W1 and W2 differ only by a TNF, then<br />

w1 = w2.<br />

Let us <strong>de</strong>compose a general <strong>de</strong>nsity w as<br />

w(v,v ′ ) =〈Λv ′ ,v ′ 〉+2ℜ〈Av ′ ,v〉+〈Σv,v〉, (10)<br />

where 〈·,·〉 is the Hermitian pro<strong>du</strong>ct in C n and Λ and Σ are Hermitian positive <strong>de</strong>finite, because<br />

of (9). We point out that in our variational context, Λ, Σ = Ση and iA = iAη have real entries.<br />

Here, we give a sufficient condition for the convexity of W, which will be shown necessary<br />

in Proposition 3.2.<br />

Theorem 2.2. Let the functional I be <strong>de</strong>fined by (8), (10). Assume that there exists a non-negative<br />

Hermitian n × n matrix K, with the property that the (2n) × (2n) Hermitian matrix<br />

<br />

Σ A+ K<br />

S :=<br />

A∗ <br />

(11)<br />

+ K Λ<br />

be non-negative. Then I is convex over H 1 (0, +∞) n .<br />

If S is positive <strong>de</strong>finite, then I is coercive.<br />

Proof. Let wS be the sesquilinear form associated to S. Then<br />

I[v]= 1<br />

2<br />

+∞<br />

<br />

wS(v, v ′ ) − 2ℜ〈Kv ′ ,v〉 dxd.<br />

0


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 417<br />

Since K is Hermitian, we have 2ℜ〈Kv ′ ,v〉=〈Kv,v〉 ′ , whence<br />

I[v]= 1<br />

2<br />

<br />

+∞<br />

0<br />

wS(v, v ′ )dxd + 1<br />

<br />

Kv(0), v(0) .<br />

2<br />

When K and S are non-negative, the right-hand si<strong>de</strong> is non-negative. If, moreover, S is positive<br />

<strong>de</strong>finite, then I dominates the H 1 -norm. ✷<br />

3. Fourier–Lap<strong>la</strong>ce analysis<br />

The most efficient method6 to study the well-posedness of a constant coefficients IBVP in a<br />

half-space is to perform a Fourier transform in the tangential variab<strong>le</strong>s y and a Lap<strong>la</strong>ce transform<br />

in the time variab<strong>le</strong> (see [10,15]). With the notations of the previous section, this yields the ODE<br />

τ 2 v − Λv ′′ + Aη − A ∗ ′<br />

η v + Σηv = 0, (12)<br />

where we keep track of the <strong>de</strong>pen<strong>de</strong>ncy of the matrices upon the frequency η. We recall that Λ<br />

and Σ are symmetric matrices with real entries, and that Aη = iA(η) where A(η) ∈ Mn(R).<br />

The advantage of the Fourier transform in y is that both the IBVP and the estimate <strong>de</strong>coup<strong>le</strong>.<br />

Thus we are <strong>le</strong>d to the equiva<strong>le</strong>nt question whether a col<strong>le</strong>ction of IBVPs in 1 + 1 dimension is<br />

strongly well-posed, with a constant C in<strong>de</strong>pen<strong>de</strong>nt of η in the estimate. Each IBVP is obtained<br />

by rep<strong>la</strong>cing ∇y by iη in the differential operator P and the boundary operator B. The re<strong>du</strong>ced<br />

system is<br />

∂ 2 t u − Λu′′ + Aη − A ∗ ′<br />

η u + Σηu = fη,<br />

whi<strong>le</strong> the re<strong>du</strong>ced boundary condition is<br />

The nee<strong>de</strong>d estimate is<br />

ˆB(η)u := Λu ′ (0) + A ∗ ηu(0) = 0. (13)<br />

e −2γT (η ⊗ u, u ′ )(T ) 2<br />

L 2 + γ<br />

C<br />

T<br />

(η ⊗ u, u ′ )(0) 2<br />

L 2 + 1<br />

γ<br />

0<br />

e −2γt (η ⊗ u, u ′ )(t) 2<br />

L 2 dt<br />

T<br />

0<br />

e −2γtfη(t) 2 L2 <br />

dt . (14)<br />

The estimate above implies c<strong>la</strong>ssically that the prob<strong>le</strong>m (12), together with boundary condition<br />

Λv ′ (0) + A∗ ηv(0) = 0 and v(+∞) = 0 does not admit a non-trivial solution, when ℜτ >0.<br />

6 In the sense that it always gives the right result. On specific situations, other methods, from <strong>semi</strong>-group theory for<br />

instance, may be faster.


418 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

This is the so-cal<strong>le</strong>d Lopatinskiĭ condition. For such a solution v(xd), one could construct solutions<br />

uκ(x, t) := e κτt v(κxd) (κ >0),<br />

whose growth rate κℜτ turns out to be unboun<strong>de</strong>d, revealing a Hadamard instability.<br />

The estimate (14) actually gives a litt<strong>le</strong> bit more, because the boundary frequencies ℜτ = 0<br />

do p<strong>la</strong>y a ro<strong>le</strong> in the theory. At <strong>le</strong>ast, we easily see that if v satisfies (12) with τ = 0, together<br />

with Λv ′ (0) + A ∗ η v(0) = 0 and v(+∞) = 0, then u(xd,t):= tv(xd) is a solution of the re<strong>du</strong>ced<br />

IBVP with fη ≡ 0, u(0) ≡ 0 and u ′ (0) ≡ v. Then taking γ = 1/T and <strong>le</strong>tting T →+∞vio<strong>la</strong>tes<br />

(14). We conclu<strong>de</strong> that for the homogeneous IBVP to be strongly well-posed, one needs the<br />

Lopatinskiĭ property at τ = 0 too. This is a non-trivial fact, re<strong>la</strong>ted to the special form of the<br />

prob<strong>le</strong>ms studied here, since this kind of well-posedness does not need the uniform Lopatinskiĭ<br />

property. See Section 4.2 for a re<strong>la</strong>ted look at the point τ = 0.<br />

Remark that since Λ is positive <strong>de</strong>finite, the energy Iη is coercive for η = 0. By Hil<strong>le</strong>–Yosida,<br />

the corresponding re<strong>du</strong>ced IBVP is strongly well-posed. Therefore one needs only to consi<strong>de</strong>r<br />

η = 0 in the sequel.<br />

3.1. The stab<strong>le</strong> subspace<br />

Assuming that W was strictly rank-one convex, we have at <strong>le</strong>ast Λ>0n, thus (12) is genuinely<br />

of second-or<strong>de</strong>r and its solution space has dimension 2n. Actually ∂2 t + P is hyperbolic<br />

and therefore it is c<strong>la</strong>ssical that when η and τ vary in such a way that η ∈ Rd−1 and ℜτ >0, then<br />

the space of solutions of (12) splits into a stab<strong>le</strong> and an unstab<strong>le</strong> subspaces, of which the dimensions<br />

do not <strong>de</strong>pend on (η, τ). Stab<strong>le</strong> (respectively unstab<strong>le</strong>) means that the e<strong>le</strong>ments v of the<br />

corresponding subspace satisfy v(+∞) = 0 (respectively v(−∞) = 0). When η = 0 and τ = 1,<br />

the equation re<strong>du</strong>ces to v − Λv ′′ = 0 and therefore both subspaces have dimension n. Inthesequel,<br />

we <strong>de</strong>note by E(τ,η) the space of values taken by (v(x), v ′ (x)) when v runs over the space<br />

of stab<strong>le</strong> solutions. This is an n-dimensional subspace of C2n , cal<strong>le</strong>d the stab<strong>le</strong> subspace of (12).<br />

This analysis is valid more generally when (τ, η) runs over a connected set U, provi<strong>de</strong>d the<br />

central subspace remains trivial for every (τ, η) in U. This means that the equation<br />

<strong>de</strong>t τ 2 In + ξ 2 Λ + iξ Aη − A ∗ <br />

η + Ση = 0 (15)<br />

does not have a real root ξ. ThesetUthat we have in mind is the union of the right and <strong>le</strong>ft<br />

half-p<strong>la</strong>nes ℜτ = 0, with the set of pairs (iρ, η) <strong>de</strong>fined by ρ ∈ R and |ρ| 0, or (elliptic points of the frequency boundary) τ = iρ<br />

with ρ ∈ (−h(η), h(η)).<br />

We now prove that whenever such a τ is the square root of a real number, then E(τ,η) can be<br />

parametrized by the component v, in the sense that there exists a matrix P = P(τ,η)∈ Mn(C)<br />

such that v ′ = Pv on E. Since dim E(τ,η) = n, this implies that E(τ,η) is exactly the solution


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 419<br />

space of the ODE v ′ = Pv. In particu<strong>la</strong>r, P must be a stab<strong>le</strong> matrix in the sense that its spectrum<br />

has negative real parts. Additionally, P(τ,η)must solve the quadratic equation<br />

τ 2 In − ΛP 2 + Aη − A ∗ η P + Ση = 0n. (17)<br />

We begin with<br />

Lemma 3.1. Let τ be a <strong>la</strong>rge enough real number. Then Eq. (17) admits a unique solution P<br />

among the matrices of which the spectrum has a negative real part (stab<strong>le</strong> matrices).<br />

This solution has the following properties:<br />

(1) ΛP + A ∗ is Hermitian.<br />

(2) P is a (Λ, Σ + τ 2 In)-isometry, in the sense that<br />

P ∗ ΛP = Σ + τ 2 In. (18)<br />

Proof. Let ɛ <strong>de</strong>note 1/τ and <strong>le</strong>t us <strong>de</strong>fine Q := ɛP. Then the equation may be rewritten as<br />

ΛQ 2 = In + 2ɛ(A − A ∗ )Q + ɛ 2 Σ.<br />

In the limit when ɛ → 0, this equation re<strong>du</strong>ces to Q 2 ∞ = Λ−1 . Since Λ is Hermitian and nonsingu<strong>la</strong>r,<br />

there exists a polynomial T(X) with non-zero simp<strong>le</strong> roots, such that T(Λ −1 ) = 0.<br />

Any solution of the limit equation satisfies T(Q 2 ∞ ) = 0. The roots of Y ↦→ T(Y2 ) being simp<strong>le</strong>,<br />

Q∞ is diagonalizab<strong>le</strong>. This tells us that the solutions are of the form Q∞ = U(Λ) where U is<br />

a polynomial that satisfies U(λ) 2 = 1/λ for every eigenvalue λ of Λ. Since the spectrum of the<br />

Q∞ must be of non-positive real part, we need U(λ)=−λ −1/2 , that is Q∞ =−Λ −1/2 .<br />

We now apply the Implicit Function theorem to the non-linear map<br />

The Q-differential at (0,Q∞) is<br />

(ɛ, Q) ↦→ ΛQ 2 − In + 2ɛ(A− A ∗ )Q − ɛ 2 Σ,<br />

C × Mn(C) → Mn(C).<br />

M ↦→ Λ(Q∞M + MQ∞).<br />

Since Λ is non-singu<strong>la</strong>r and the sum of two eigenvalues of Q∞ may not vanish (it must be<br />

negative), this differential is non-singu<strong>la</strong>r. We thus obtain the existence and uniqueness part.<br />

We emphasize that the unique stab<strong>le</strong> solution P <strong>de</strong>pends analytically on τ 2 in a neighbourhood<br />

of infinity.<br />

Let us <strong>de</strong>fine now a matrix R := α + ΛP where α := 1<br />

2 (A∗ − A) is skew-Hermitian. Then the<br />

equation can be rewritten as<br />

(R + α)Λ −1 (R − α) = Σ + τ 2 In. (19)<br />

When τ is real, the right-hand si<strong>de</strong> is Hermitian. Thus, taking the Hermitian adjoint of (19), we<br />

obtain<br />

(R ∗ + α)Λ −1 (R ∗ − α) = Σ + τ 2 In.


420 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

This shows that R ∗ is an other solution of (19) and thus P1 := Λ −1 (P ∗ Λ − 2α) is an other<br />

solution of (17). Since P ∼−τΛ −1/2 ,wehaveP1 ∼−τΛ −1/2 , and therefore the spectrum of<br />

P1 has a negative real part for τ <strong>la</strong>rge enough. From the uniqueness part, we <strong>de</strong><strong>du</strong>ce that P1 = P ,<br />

which means R ∗ = R. Hence R is Hermitian.<br />

Knowing that R is Hermitian, we may recast (19) into (18). ✷<br />

Fixing η ∈ R (say η = 0 since the case η = 0 is rather trivial), we consi<strong>de</strong>r the maximal<br />

subinterval Jη = (α, +∞) of (−h(η) 2 , +∞) on which there is an analytical function z ↦→ P(z),<br />

such that P(z)is a stab<strong>le</strong> solution of (17) with τ 2 = z. Lemma 3.1 en<strong>sur</strong>es that Jη is non-void. By<br />

analyticity, the properties stated in Lemma 3.1 remain valid over Jη. In particu<strong>la</strong>r P(z) remains<br />

uniformly boun<strong>de</strong>d over this interval, because of (18).<br />

Differentiating (17) along Jη, wehave<br />

ΛP dP<br />

dz<br />

+ ΛdP<br />

dz P + (A∗ − A) dP<br />

dz<br />

= In.<br />

Because of Lemma 3.1, part (1), this equation may be written as a Lyapunov equation for the<br />

unknown ΛP ′ :<br />

P ∗ Λ dP<br />

dz<br />

+ ΛdP<br />

dz P = In. (20)<br />

Since P is a stab<strong>le</strong> matrix, this equation has a unique solution, given by<br />

This formu<strong>la</strong> shows the following monotonicity property.<br />

Λ dP<br />

dz =−<br />

+∞<br />

e xP∗<br />

e xP dx. (21)<br />

0<br />

Lemma 3.2. The map z ↦→ ΛP (z) + A∗ η is monotonous <strong>de</strong>creasing for the natural or<strong>de</strong>r of<br />

Hermitian matrices.<br />

Equation (21) may be viewed as an ODE for z ↦→ P(z), with domain the set of stab<strong>le</strong> matrices.<br />

Notice that every solution of (21) satisfies Eq. (17) for z = τ 2 and some constant matrices A<br />

and Σ, the <strong>la</strong>tter being Hermitian. To see this, differentiate P(z) ∗ ΛP (z) and <strong>de</strong><strong>du</strong>ce that Σ :=<br />

P(z) ∗ ΛP (z) − zIn is constant. Then<br />

ΛP − (zIn + Σ)P(z) −1 = ΛP (z) − P(z) ∗ Λ<br />

is skew-Hermitian and constant (differentiate again). Thus there exists an A such that the <strong>le</strong>fthand<br />

si<strong>de</strong> equals A − A ∗ . Notice that ΛP (z) + A ∗ is Hermitian.<br />

From the above remark and Lemma 3.1, it becomes c<strong>le</strong>ar that (Jη; z ↦→ P) is the maximal<br />

solution of the ODE (21) that is <strong>de</strong>fined up to +∞. Because of the bound given by (18), this<br />

solution remains uniformly boun<strong>de</strong>d and therefore Jη equals (−h(η) 2 , +∞).<br />

We summarize our results in the following statement.


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 421<br />

Theorem 3.1. Assume that W is strictly rank-one convex. When τ 2 ∈ (−h(η) 2 , +∞), the<br />

quadratic matrix equation (17) admits a unique solution P(τ,η) among stab<strong>le</strong> matrices, with<br />

the following properties:<br />

• ΛP + A∗ η is Hermitian, that is ΛP − P ∗Λ = Aη − A∗ η ;<br />

• P ∗ΛP = Σ + τ 2In; • τ 2 ↦→ ΛP (τ, η) + A∗ η is <strong>de</strong>creasing for the or<strong>de</strong>r of Hermitian matrices.<br />

Remark.<br />

• A simi<strong>la</strong>r analysis works for the unstab<strong>le</strong> solution Pu of (17). In particu<strong>la</strong>r ΛPu + A ∗ is<br />

Hermitian. However, there is no reason why the other solutions of (17) have this property.<br />

Likewise, only the stab<strong>le</strong> and the unstab<strong>le</strong> solutions satisfy (18).<br />

• When τ 2 ∈ (−h(η) 2 , +∞), the stab<strong>le</strong> space is given by the equation v ′ = P(τ,η)v.<br />

The above analysis also en<strong>sur</strong>es that the limit P0(η) := P(−h(η) 2 ) exists. We discuss in which<br />

way P0(η) is not a stab<strong>le</strong> matrix.<br />

Proposition 3.1. The matrix P0(η) admits a pure imaginary eigenvalue iξ0 (ξ0 ∈ R). The associated<br />

eigenvector X0 can be chosen in R n . Finally, the Hermitian form <strong>de</strong>fined by ΛP0(η) + A ∗<br />

vanishes at X0.<br />

Proof. Since P(z) is a stab<strong>le</strong> matrix for z>−h(η) 2 , the spectrum of P0(η) has a non-positive<br />

real part. Likewise, the first-or<strong>de</strong>r ODE v ′ = P0(η)v <strong>de</strong>fines an n-dimensional invariant subspace<br />

within the solution space of (12) where τ = ih(η).IfP0(η) was stab<strong>le</strong>, then ih(η) would be an<br />

interior point of the elliptic interval, which is false. Therefore P0(η) must have a pure imaginary<br />

eigenvalue, say iξ0, ξ0 ∈ R. LetX0be an associated eigenvector. From Eq. (17), and with Aη =<br />

iA(η),wehave<br />

2 T<br />

ξ0 Λ − ξ0 A(η) + A(η) + Ση − h(η) 2 <br />

In X0 = 0. (22)<br />

Since the matrix in this equality has real entries, we may choose X0 in Rn .<br />

Recall that h(η) 2 is the <strong>le</strong>ast among the numbers<br />

2<br />

λ− ξ Λ + iξ Aη − A ∗ <br />

η + Ση<br />

when ξ runs over R. In particu<strong>la</strong>r, we have that<br />

ξ ↦→ g(ξ) := X ∗ 2<br />

0 ξ Λ + iξ Aη − A ∗ <br />

η + Ση X0 − h(η) 2 |X0| 2<br />

is non-negative. Since (22) implies that g(ξ0) = 0, there follows that g ′ (ξ0) vanishes, which gives<br />

Hence we have<br />

proving the c<strong>la</strong>im. ✷<br />

ξ0X T 0 ΛX0 = X T 0 A(η)X0. (23)<br />

X ∗ 0 ΛP0(η) + A ∗ X0 = X T <br />

0 iξ0Λ − iA(η) T X0 = 0,


422 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

Remark. The continuity result in Proposition 3.1 is weaker than that obtained by Métivier [14]<br />

when the characteristics have constant multiplicities. Un<strong>de</strong>r the <strong>la</strong>tter assumption, the limit of<br />

E(τ,η) exists at (ih(η0), η0), without any restriction about (τ, η) ∈ U, whi<strong>le</strong> our result concerns<br />

only the limit along an interval. We recall that points like (ih(η0), η0) are g<strong>la</strong>ncing points.<br />

Proposition 3.1 tells that ΛP0(η) + A∗ η cannot be negative <strong>de</strong>finite. We shall see <strong>la</strong>ter on that<br />

it may be non-positive. In such a case, the i<strong>de</strong>ntity X∗ 0 (ΛP0(η) + A∗ η )X0 = 0 implies that X0 is<br />

an eigenvector:<br />

<br />

ΛP0(η) + A ∗ η X0 = 0. (24)<br />

Then Eqs. (24) and (17) yield<br />

In other words, we have<br />

3.2. The Lopatinskiĭ <strong>de</strong>terminant<br />

Ση − h(η) 2 In − ξ0A(η) X0 = 0. (25)<br />

Λ A ∗ η<br />

Aη Ση − h(η) 2 In<br />

iξ0X0<br />

The boundary operator becomes after Fourier–Lap<strong>la</strong>ce transformation<br />

X0<br />

ˆB(η)v := Λv ′ (0) + A ∗ η v(0).<br />

<br />

= 0. (26)<br />

For the IBVP to be C∞-well-posed, it is necessary and sufficient that whenever η ∈ Rd−1 and<br />

ℜτ >0, every stab<strong>le</strong> solution of (12) satisfying ˆB(η)v = 0 vanishes i<strong>de</strong>ntically (see [9]). This<br />

is the so-cal<strong>le</strong>d Lopatinskiĭ condition, which amounts to saying that ˆB(η): E(τ,η) → Cn is an<br />

isomorphism for all pairs (τ, η) as above. The Lopatinskiĭ condition at point (τ, η) can be written<br />

as (τ, η) = 0 where the Lopatinskiĭ <strong>de</strong>terminant is <strong>de</strong>fined by<br />

(τ, η) := <strong>de</strong>t ΛP (τ, η) + A ∗ η , (27)<br />

provi<strong>de</strong>d a stab<strong>le</strong> solution P(τ,η) (necessarily unique) of (17) exists. Because of Theorem 3.1,<br />

this holds true at <strong>le</strong>ast when τ 2 ∈ (−h(η) 2 , +∞).<br />

Remark. The Eu<strong>le</strong>r–Lagrange equations of the extremum prob<strong>le</strong>m studied in the next section<br />

consist of the ODE (12), together with the boundary condition<br />

Λv ′ (0) + A ∗ ηv(0) = 0, (28)<br />

that is ˆB(η)v(0) = 0. Therefore the existence of a non-trivial solution v0 given in Theorem 5.1<br />

implies that ( √ −β,η) = 0.<br />

From Theorem 3.1, we have the following remarkab<strong>le</strong> property.


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 423<br />

Theorem 3.2. With the <strong>de</strong>finition (27), the Lopatinskiĭ <strong>de</strong>terminant is a real analytic function of<br />

τ 2 over the interval (−h(η) 2 , +∞), continuous over [−h(η) 2 , +∞).<br />

Remark. The fact that the Lopatinskiĭ <strong>de</strong>terminant is a real-valued function along R + <strong>le</strong>aves<br />

the possibility to <strong>de</strong>fine a stability in<strong>de</strong>x as in [8], even when η = 0. A stability in<strong>de</strong>x arises for<br />

boundary <strong>la</strong>yers or shock <strong>la</strong>yers in viscous systems of conservation <strong>la</strong>ws. This in<strong>de</strong>x, as <strong>de</strong>fined<br />

by Gardner and Zumbrun, concerns the stability of a <strong>la</strong>yer U(xd) un<strong>de</strong>r initial disturbances of the<br />

form δu0(xd); it counts the changes of sign of the so-cal<strong>le</strong>d Evans function between the origin<br />

and +∞, using the fact that this function is real-valued when τ ∈ R + and η = 0. Thus it makes<br />

sense only for one-dimensional systems. It was shown in [8] (one-dimensional case) and in [19]<br />

(multi-dimensional case) that the Evans function is “tangent” to the Lopatinskiĭ <strong>de</strong>terminant at<br />

the origin. Thus there is a hope to generalize the stability in<strong>de</strong>x to every Fourier frequency, in the<br />

variational case.<br />

3.3. Continuation of the Lopatinskiĭ <strong>de</strong>terminant to the hyperbolic region<br />

We have shown in previous works [4,17] that the generalized stab<strong>le</strong> space E(iρ,η), <strong>de</strong>fined<br />

as the limit of E(iρ + ɛ,η) as ɛ → 0 + , has a basis ma<strong>de</strong> of “real” vectors when ρ is real and<br />

<strong>la</strong>rge enough (hyperbolic region of the frequency boundary). Since the works cited above <strong>de</strong>alt<br />

with first-or<strong>de</strong>r systems<br />

∂tu + <br />

A α ∂αu = 0,<br />

α<br />

we need to exp<strong>la</strong>in what a “real vectors” means in the context of second-or<strong>de</strong>r systems. Since we<br />

can go back to the first-or<strong>de</strong>r by choosing the new variab<strong>le</strong> (∂tu, ∇xu), a “real vector” must be<br />

un<strong>de</strong>rstood as a vector of the form<br />

v ′<br />

v<br />

<br />

=<br />

<br />

iμv<br />

v<br />

for some real number μ and real vector v ∈ Rn .<br />

The property of reality allowed us to <strong>de</strong>fine an (equiva<strong>le</strong>nt and not canonical) Lopatinskiĭ<br />

<strong>de</strong>terminant, in such a way that it was a real analytic function of ρ in this region. This property,<br />

which is true for every hyperbolic IBVP, applies to the c<strong>la</strong>ss of variational IBVPs. Remark that<br />

on the contrary, Theorem 3.2 holds true only for the c<strong>la</strong>ss of variational IBVPs.<br />

In the <strong>la</strong>tter c<strong>la</strong>ss, one may wan<strong>de</strong>r whether our canonical <strong>de</strong>finition (27) of the Lopatinskiĭ<br />

<strong>de</strong>terminant is real analytic in the hyperbolic region too, <strong>de</strong>fined by τ = iρ with<br />

ρ 2 > min<br />

ξ∈R λ+<br />

2 T<br />

ξ Λ − ξ A(η) + A(η) <br />

+ Ση . (29)<br />

We can see that it is true, up to a factor ( √ −1) n . As a matter of fact, in the hyperbolic region the<br />

matrix P(iρ,η)takes the form iM(ρ,η) where M is a solution of the equation<br />

ΛM 2 − A(η) + A(η) T M + Σ = ρ 2 In, (30)


424 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

with real entries. To see that this is true, we have to show that if ρ satisfies (29), then the 2n roots<br />

of the polynomial<br />

X ↦→ <strong>de</strong>t ρ 2 In − X 2 Λ + X A(η) + A(η) T <br />

− Ση<br />

are real. Actually, these are the eigenvalues of the matrix<br />

Conjugation of M0 by<br />

yields the matrix<br />

with<br />

M0 :=<br />

Λ −1 (A(η) + A(η) T ) Λ −1 (ρ 2 In − Ση)<br />

In<br />

Λ 1/2 −ξΛ 1/2<br />

0n<br />

ξΛ 1/2<br />

<br />

,<br />

<br />

S ξ−1 T<br />

M(ξ) :=<br />

ξIn In<br />

S := Λ −1/2 A(η) + A(η) T Λ −1/2 − ξIn, T := ξS + Λ −1/2 ρ 2 In − Σ Λ −1/2 .<br />

By assumption, there exists a ξ ∈ R such that T is positive <strong>de</strong>finite. Then conjugation of ξM(ξ)<br />

by diag{ξIn,T 1/2 } yields the matrix<br />

ξS ξT 1/2<br />

ξT 1/2 ξIn<br />

The <strong>la</strong>tter, being Hermitian, is diagonalizab<strong>le</strong> with a real spectrum. This proves the c<strong>la</strong>im, and<br />

the fact that every solution of (30) belongs to Mn(R). Of course, only that one associated to the<br />

eigenvalues iμ which enter into the right half-p<strong>la</strong>ne when ℜτ increases is re<strong>le</strong>vant.<br />

Let us <strong>de</strong>note this matrix by M(ρ,η). Then <strong>de</strong>finition (27) can be exten<strong>de</strong>d, and yields<br />

<br />

.<br />

0n<br />

(iρ, η) = i n <strong>de</strong>t ΛM(ρ, η) − A(η) T ,<br />

where the matrix in the <strong>de</strong>terminant has real entries, hence the <strong>de</strong>terminant is a real number. We<br />

<strong>le</strong>ave open the fact that the matrix P(τ,η) can be <strong>de</strong>fined continuously and analytically along a<br />

neighbourhood of the real axis within the right half-p<strong>la</strong>ne. In this case, we should have a unique<br />

analytic function, real-valued along R and along (−ih(η), ih(η)), and real-valued up to the factor<br />

i n in the far field of the imaginary axis. A typical examp<strong>le</strong> of such a function is τ ↦→ √ τ 2 + 1,<br />

which is real for τ ∈ (−i, +i) and pure imaginary in the rest of the imaginary axis.<br />

<br />

.


3.4. The zeroes of ( √ z, η)<br />

D. Serre / Journal of Functional Analysis 236 (2006) 409–446 425<br />

Theorem 3.3. The function z ↦→ ( √ z, η) vanishes at <strong>le</strong>ast once in [−h(η) 2 , +∞), and at most<br />

n − 1 times in (−h(η) 2 , +∞). The <strong>la</strong>rgest zero is the number a such that ΛP (a) + A ∗ is nonpositive<br />

and singu<strong>la</strong>r.<br />

If ΛP (0) + A ∗ has a positive eigenvalue, then the IBVP is strongly unstab<strong>le</strong>.<br />

Remark. In particu<strong>la</strong>r, the so-cal<strong>le</strong>d uniform Lopatinskiĭ condition is never satisfied, except in<br />

one space dimension, since (τ, η) must vanish somewhere in ℜτ 0, (τ, η) = (0, 0). This<br />

implies that the estimates in the non-homogeneous IBVP (g ≡ 0 in (2)) do suffer a loss of <strong>de</strong>rivatives.<br />

Proof. The signature of ΛP (z) + A∗ varies monotonically and ( √ z, η) vanishes only when<br />

this signature changes. Thus it may vanish at most n times in (−h(η) 2 , +∞), and this maximum<br />

is achieved when ΛP0(η) + A∗ η is positive <strong>de</strong>finite. However, Proposition 3.1 ru<strong>le</strong>s out this<br />

possibility. Likewise Proposition 3.1 prevents ΛP0(η) + A∗ η to be negative <strong>de</strong>finite, whence the<br />

existence of one zero at <strong>le</strong>ast in [−h(η) 2 , +∞).<br />

By continuity and monotonicity, ( √ z, η) does not vanish in (a, +∞) when ΛP (a) + A∗ is non-positive. The <strong>la</strong>st c<strong>la</strong>im is the special case a = 0; un<strong>de</strong>r the assumption, the Lopatinskiĭ<br />

condition is not satisfied. ✷<br />

3.5. The converse to Theorems 2.2 and 3.3<br />

Let S(K) be the matrix occurring in Theorem 2.2. We begin with the observation that if we<br />

choose K =−(ΛP (a) + A ∗ ), which is Hermitian, then the i<strong>de</strong>ntities (17) and (18) imply that<br />

<br />

P(a) ∗ <br />

S(K) = Λ<br />

−In<br />

<br />

P(a),−In − a<br />

which is positive <strong>de</strong>finite when a0n<br />

and S>02n in the computation above. ✷<br />

This proposition has a remarkab<strong>le</strong> consequence for the characterization of strongly well-posed<br />

homogeneous IBVP. We recall that a necessary (though not sufficient) condition for the strong<br />

well-posedness is that the Lopatinskiĭ property be satisfied for every pair (τ, η) with either<br />

ℜτ >0orτ = 0. In particu<strong>la</strong>r, it must hold true when τ ∈ R + . We want to prove a converse<br />

<br />

,


426 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

of the <strong>la</strong>st sentence. For this, <strong>le</strong>t us assume only that (·,η) does not vanish over R + . Applying<br />

Proposition 3.2 and Theorem 2.2, we see that the functional I <strong>de</strong>fined by (8) is convex and coercive<br />

over H 1 (R + ). Thus we may apply the Hil<strong>le</strong>–Yosida theorem and obtain the well-posedness<br />

of the homogeneous IBVP at fixed η, in1+ 1 dimension. In turn, this well-posedness en<strong>sur</strong>es<br />

that the Lopatinskiĭ property holds true for every ℜτ >0. We summarize this analysis in the<br />

following statement.<br />

Theorem 3.4. Assume that (·,η)does not vanish over R + . Then<br />

Iη[v]:= 1<br />

2<br />

<br />

+∞<br />

is convex and coercive over H 1 (R + ). In particu<strong>la</strong>r:<br />

0<br />

w(v,v ′ )dxd<br />

• The corresponding homogeneous IBVP at frequency η is well-posed in H 1 (R + ).<br />

• ˆB(η): E(τ,η) ↦→ C n is an isomorphism when ℜτ >0.<br />

• The same holds true in the non-elliptic part of the boundary ℜτ = 0.<br />

To our know<strong>le</strong>dge, it is the first time that a structural assumption yields the conclusion that a<br />

strong instability (the vanishing of (τ, η) for some ℜτ >0, or more generally the Lopatinskiĭ<br />

condition) must happen either for a real τ , or nowhere. We point out that Theorem 3.4 is a kind<br />

of converse to Theorem 3.3, in the sense that if the homogeneous IBVP is strongly unstab<strong>le</strong>,<br />

then (·,η) must have a zero over R + , and therefore ΛP (0) + A ∗ must have a non-negative<br />

eigenvalue (presumably positive).<br />

3.5.1. Well-posedness of the full homogeneous IBVP<br />

We now <strong>le</strong>t varying the space frequency η. Looking for a criterion of well-posedness for the<br />

full homogeneous IBVP (1), (2) (thus with g ≡ 0), we make the natural assumption that does<br />

not vanish at all over R + × R. As shown above, each Iη is convex and coercive over H 1 (R + ).<br />

Let us remark that the various objects of the theory are homogeneous in their arguments. If κ is<br />

a positive real number, then we have<br />

A(κη) = κA(η), Σκη = κ 2 Ση, E(κτ,κη)= E(τ,η),<br />

P(κτ,κη)= κP (τ, η), h(κη) = κh(η), (κτ, κη) = κ(τ, η).<br />

Using these properties and the fact that the unit sphere of R d−1 is compact, we see that the<br />

number a = a(η) in the construction above can be chosen in the form a(η) =−ɛ|η| 2 ,forafixed<br />

ɛ>0, small enough. We <strong>de</strong><strong>du</strong>ce that<br />

+∞<br />

′ 2 2 2<br />

Iη[v] Cɛ |v | +|η| |v| dxd, (32)<br />

0<br />

where C does not <strong>de</strong>pend on η. Then inverse Fourier transform gives that W dominates the norm<br />

of H˙ 1 (Ω). In conclusion, W is convex and coercive over H˙ 1 (Ω). Whence our main result:


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 427<br />

Theorem 3.5. We assume that W is strictly rank-one convex. Then the following statements are<br />

equiva<strong>le</strong>nt to each other:<br />

(1) The stored energy W is convex and coercive over H˙ 1 (Ω).<br />

(2) The homogeneous IBVP is strongly well-posed in H˙ 1 (Ω).<br />

(3) One has (τ, η) = 0 for every τ ∈ R + , η ∈ R \{0}.<br />

(4) For every η = 0, the Hermitian matrix H(η):= −(ΛP (0,η)+ A∗ η ) is positive <strong>de</strong>finite.<br />

Proof. Essentially everything has been proved yet. We content ourselves to recall the main arguments.<br />

(1) ⇒ (2). Apply Hil<strong>le</strong>–Yosida theorem.<br />

(2) ⇒ (3). The strong well-posedness implies the Lopatinskiĭ condition for ℜτ >0, whence<br />

the special case of τ ∈ (0, +∞). We have seen the necessity of Lopatinskiĭ forτ = 0inthe<br />

intro<strong>du</strong>ction of Section 3.<br />

(3) ⇒ (4). The assumption tells that the Hermitian matrix ΛP (τ, η) + A∗ η is non-singu<strong>la</strong>r for<br />

every τ ∈ R + . Since it is negative <strong>de</strong>finite for τ <strong>la</strong>rge, it remains so for every τ 0, by continuity.<br />

(4) ⇒ (1). In the present paragraph, we have shown that the assumption implies a uniform<br />

inequality (32) for some ɛ>0, at <strong>le</strong>ast for η = 0. By continuity, it remains true for η = 0, whence<br />

the coercivity of W. ✷<br />

Remark. The matrix K(η) constructed above is homogeneous of <strong>de</strong>gree one in η, but is not<br />

linear in η in general. There is no special reason why a sing<strong>le</strong> non-tangential null-form could be<br />

ad<strong>de</strong>d to W in such a way that the new <strong>de</strong>nsity and the boundary term be strictly convex.<br />

3.6. The influence of non-tangential null-forms to W<br />

We observe on the one hand that the matrix P(z)<strong>de</strong>pends only on z, Ση and ℑAη = i 2 (A∗η −<br />

Aη) = 1 2 (A(η) + A(η)T ), but not on ℜAη, whi<strong>le</strong> the matrix ΛP (τ 2 ) + A∗ η that enters in the<br />

Lopatinskiĭ <strong>de</strong>terminant does involve ℜAη. On the other hand, the addition of a non-tangential<br />

null-form is a mean for modifying ℜAη and only that. This remark is consistent with that ma<strong>de</strong><br />

above: the addition of a non-tangential null-form does not change the differential operator, but<br />

modifies the boundary operator.<br />

Recall that the matrix A(η) has real entries. Thus<br />

ℜAη = i T<br />

A(η) − A(η)<br />

2<br />

<br />

involves only the skew-symmetric part of A(η). Of course, h(η) <strong>de</strong>pends on the symmetric part<br />

of A(η), but not on its skew-symmetric part. It turns out that the non-tangential null-forms are<br />

in one-to-one correspon<strong>de</strong>nce with the skew-symmetric matrices. Therefore p<strong>la</strong>ying with nontangential<br />

null-forms allows us to add to ΛP + A ∗ an arbitrary matrix of the form iM, with<br />

M ∈ Skewn(R).<br />

Proposition 3.3. Given Λ, Ση and the symmetric part of A(η), one can choose the skewsymmetric<br />

part in such a way that (±ih(η), η) vanishes.<br />

In other words, given the energy <strong>de</strong>nsity, one can add a null-form in such a way that<br />

(±ih(η), η) vanishes.


428 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

Proof. Let Y <strong>de</strong>note the real vector (ξ0Λ − A(η) T )X0. From (23), Y is orthogonal to X0. Thus<br />

there exists a skew-symmetric matrix M with real entries such that MX0 = Y . If we rep<strong>la</strong>ce A(η)<br />

by à := A(η) − M, we obtain (ΛP0(η) + à ∗ )X0 = 0, whence<br />

as <strong>de</strong>sired. ✷<br />

Remark.<br />

<strong>de</strong>t ΛP0(η) + Ã ∗ = 0,<br />

• There is no reason why the additional null-form of the proposition be in<strong>de</strong>pen<strong>de</strong>nt of η. Thus<br />

it is not possib<strong>le</strong> in general to find a sing<strong>le</strong> null-form such that (±ih(η), η) = 0 for every η.<br />

This could be achieved by adding a null-form, pseudo-differential in the tangential variab<strong>le</strong>s.<br />

• One may wan<strong>de</strong>r whether it is possib<strong>le</strong> to adjust the skew-symmetric part of A(η) in such a<br />

way that ΛP0(η) + A∗ η be non-positive. In this case, (√z, η) vanishes only at the extremity<br />

−h(η) 2 , whence a well-posedness at frequency η, plus the property that the <strong>sur</strong>face waves<br />

have an infinite energy! We <strong>le</strong>ave the rea<strong>de</strong>r check that such a choice is possib<strong>le</strong> if, and<br />

only if, the restriction of the Hermitian form X ↦→ X∗ (ΛP0(η) + A∗ η )X to real vectors is<br />

non-positive.<br />

4. Surface waves<br />

Assume that a given hyperbolic IBVP satisfies the Lopatinskiĭ condition, though not necessarily<br />

in a uniform way. Let us assume also that the stab<strong>le</strong> space E(τ,η) admits a continuous<br />

extension E(iρ,η) at some boundary point (iρ, η). We say that the IBVP admits a <strong>sur</strong>face wave<br />

with frequency (ρ, η) if ˆB(η): E(iρ,η) → C n is singu<strong>la</strong>r. This amounts to the existence of a<br />

non-trivial solution v of (12) for τ = iρ, with the properties that ˆB(η)v(0) = 0 and that v is<br />

polynomially boun<strong>de</strong>d at infinity. Then<br />

u(x, t) := e i(ρt+η·y) v(xd)<br />

is a solution of (1), (2) with f ≡ 0 and g ≡ 0. When v tends to zero at +∞, the <strong>de</strong>cay is exponential<br />

and we say that the <strong>sur</strong>face wave has finite energy. When E(iρ,η) equals the stab<strong>le</strong> subspace<br />

of (12), a <strong>sur</strong>face wave at frequency (ρ, η) is automatically of finite energy. This is, of course,<br />

the case when (iρ, η) lies in the elliptic region of the frequency boundary. Thus the existence of<br />

a <strong>sur</strong>face wave of elliptic frequency (ρ, η) is equiva<strong>le</strong>nt to (iρ, η) = 0. Theorem 3.3 tells that<br />

if W is convex and coercive, a finite energy <strong>sur</strong>face wave (FESW) must exist at space frequency<br />

η for some time frequency ρ, except in the marginal case where ΛP0(η) + A∗ η is non-positive.<br />

A well-known examp<strong>le</strong> of FESW is the Ray<strong>le</strong>igh wave in isotropic e<strong>la</strong>sticity.<br />

See [3] for an analysis of the ro<strong>le</strong> of <strong>sur</strong>face waves in the homogeneous IBVP. It is also<br />

exp<strong>la</strong>ined in [4] that among the c<strong>la</strong>ss of strictly (or symmetric) hyperbolic IBVPs, the prob<strong>le</strong>ms<br />

that admit FESW are non-generic. They form a hyper<strong>sur</strong>face in the space of hyperbolic IBVPs:<br />

un<strong>de</strong>r a small perturbation in the differential operator, or in the boundary operator, the IBVP falls<br />

either in the strongly well-posed c<strong>la</strong>ss or in the strongly ill-posed c<strong>la</strong>ss, generically. We shall<br />

show below that this picture becomes false when the perturbation keeps our IBVP within the<br />

c<strong>la</strong>ss of variational prob<strong>le</strong>ms (1), (2). Of course, this restriction is non-generic in the context of<br />

[4], but does have a physical re<strong>le</strong>vance.


4.1. Persistence of <strong>sur</strong>face waves<br />

D. Serre / Journal of Functional Analysis 236 (2006) 409–446 429<br />

Let us assume that for a given strictly rank-one convex energy W0, the IBVP (1), (2) admits a<br />

<strong>sur</strong>face wave of finite energy for some frequency η. Then there exists a ρ0 ∈ (−h(η), h(η)) such<br />

that (iρ0,η)= 0. Because of Lemma 3.2, there is a change in the signature of ΛP (iρ, η) + A∗ η<br />

across ρ0. Say that for ɛ>0 small enough, ΛP (i(ρ0 ± ɛ),η) + A∗ η are non-singu<strong>la</strong>r, and the<br />

numbers p± of their positive eigenvalues satisfy p+ >p−.<br />

When we make a small enough perturbation W = W0 + δW, the Hermitian matrices<br />

ΛP (i(ρ0 ± ɛ),η) + A∗ η still have p± positive eigenvalues. Therefore there exist a ρ in (ρ0 − ɛ,<br />

ρ0 + ɛ) such that ΛP (iρ, η) + A∗ η be singu<strong>la</strong>r. Then (iρ, η)) = 0 and there exists a FESW at<br />

frequency (ρ, η). This is the phenomenon of persistence of FESW. Remark that there may be<br />

(and there are, generically) p+ − p− such roots ρ in the interval (ρ0 − ɛ,ρ0 + ɛ). Notice that the<br />

same kind of persistence occurs when we perturb the frequency η (consi<strong>de</strong>ring instead η + δη)<br />

instead of the energy <strong>de</strong>nsity.<br />

Let us write η0 instead of η, and <strong>le</strong>t us vary the space frequency. When p+ = p− + 1, ρ0 is<br />

a simp<strong>le</strong> root of (·,η0) in the unperturbed prob<strong>le</strong>m. Then there exists an analytic function<br />

η ↦→ ρ(η) <strong>de</strong>fined in a neighbourhood of η0, satisfying ρ(η0) = ρ0 and (ρ(η), η) = 0. If ρ<br />

is globally <strong>de</strong>fined, then the wave front of the trace of solutions of (∂ 2 t + P)u∈ C∞ , along the<br />

boundary {xd = 0}, is inclu<strong>de</strong>d in the characteristic cone<br />

(ρ, η) ∈ R × R d−1 ; ρ = ρ(η) .<br />

Finally, we point out that the dimension of the space of FESWs is re<strong>la</strong>ted to (p−,p+):<br />

Proposition 4.1. With the notations above, the dimension of the kernel of ΛP (iρ0,η)+ A∗ η , that<br />

is the dimension of the space of FESWs at frequency (ρ0,η), equals p+ − p−.<br />

Proof. When ρ → ρ0 − 0, ΛP (iρ, η) + A ∗ η has p− positive and n − p− negative eigenvalues.<br />

Since this Hermitian matrix increases when ρ increases (notice that ρ ↦→ z = (iρ) 2 is <strong>de</strong>creasing!),<br />

we <strong>de</strong><strong>du</strong>ce by continuity that ΛP (iρ0,η)+A ∗ η has exactly p− positive eigenvalues. Letting<br />

ρ → ρ0 + 0, we find likewise that ΛP (iρ0,η) + A ∗ η has exactly n − p+ negative eigenvalues.<br />

The dimension of its kernel being equal to the multiplicity of the null eigenvalue, it is<br />

n − p− − (n − p+) = p+ − p−. ✷<br />

4.2. Transition to instability<br />

From Theorem 3.5, a transition towards instability happens precisely when the matrix H(η)=<br />

−(ΛP (0,η)+ A∗ η ) stops to be positive <strong>de</strong>finite. This can be achieved for instance by choosing<br />

ℜAη so as to have X∗ (ΛP (0) + A∗ η )X > 0 for some vector X, following an i<strong>de</strong>a employed in<br />

Section 3.6.<br />

Of course, since the IBVP is ill-posed whenever there exists at <strong>le</strong>ast one frequency η for<br />

which (·,η)has a root of positive real part, this transition happens when the sign of<br />

min d−1<br />

λ− H(η) : η ∈ S <br />

changes.


430 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

We now enlight a connection between the transition <strong>de</strong>scribed above and the nature of the<br />

elliptic BVP<br />

Lu = f in Ω, (33)<br />

Bu = g on ∂Ω. (34)<br />

Such prob<strong>le</strong>ms have been studied systematically by Agmon et al. [1]. See also the <strong>semi</strong>nal work<br />

by Lopatinskiĭ [11]. They show that a priori estimates are avai<strong>la</strong>b<strong>le</strong> if and only if a comp<strong>le</strong>menting<br />

condition holds true at non-zero frequencies η. It turns out that this condition coinci<strong>de</strong>s with<br />

(0,η)= 0 for all non-zero vector η ∈ R d−1 , where is the Lopatinskiĭ <strong>de</strong>terminant. Therefore,<br />

if a transition between well- and ill-posedness occurs at some hyperbolic IBVP, then the<br />

corresponding elliptic BVP is ill-posed.<br />

We warn the rea<strong>de</strong>r that the converse is not true when d 3, since the set of ill-posed elliptic<br />

BVPs has a non-void interior (in the set of parameters). As a matter of fact, in most cases,<br />

the vanishing of the function η ↦→ (0,η) at some point η0 = 0 means that the sign of this<br />

function changes. 7 Then a small perturbation in L or B yields a small perturbation in , so that<br />

(0, ·) still vanishes somewhere and the modified elliptic BVP remains ill-posed. Within the<br />

set of ill-posed elliptic BVPs in the sense of ADN, only those at the boundary may correspond<br />

to a transition in the hyperbolic IBVP. More precisely, the comp<strong>le</strong>ment function must vanish<br />

somewhere, without changing sign.<br />

We point out that this analysis <strong>le</strong>aves the possibility that the hyperbolic IBVP is ill-posed<br />

whi<strong>le</strong> the steady elliptic BVP is well-posed. This possibility is confirmed by explicit examp<strong>le</strong>s;<br />

see, for instance, Theorem 6.1.<br />

5. An extremum prob<strong>le</strong>m<br />

We solve in this section an abstract prob<strong>le</strong>m of extremum. Theorem 5.1 gives an alternate<br />

proof of the fact that every variational IBVP either admits <strong>sur</strong>face waves, or is unstab<strong>le</strong> in the<br />

Hadamard sense.<br />

We start with a functional I as in (8), where w is a sesquilinear form satisfying (9). Let us<br />

<strong>de</strong>fine the finite number<br />

I[v]<br />

β := inf , E[v]:=1<br />

v≡0 E[v] 2<br />

<br />

+∞<br />

0<br />

|v| 2 dxd. (35)<br />

We have the property that I − γE is convex over H 1 (R + ) if and only if γ β. Testing with the<br />

fields v of the form e −ωxd V , we immediately find the property<br />

In particu<strong>la</strong>r, we have<br />

(ℜω 0) ⇒ βIn Θ(ω) , Θ(ω):= |ω| 2 In − ωA −¯ωA ∗ + Σ. (36)<br />

βIn Θ(iξ), ∀ξ ∈ R.<br />

7 This is, of course, not the case if d = 2, because of homogeneity.


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 431<br />

When the <strong>la</strong>tter inequalities are strict, a fact that occurs for instance when<br />

min<br />

ℜω0 λ−<br />

<br />

Θ(ω) < min<br />

ξ∈R λ−<br />

<br />

Θ(iξ) , (37)<br />

we have the following result.<br />

Theorem 5.1. Let us assume (9), together with the strict inequality<br />

β


432 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

for every z ∈ H 1 (R), a property that follows easily from Fourier analysis. We thus have I[zm] <br />

λ0E[zm]. Passing to the limit in (39), we <strong>de</strong><strong>du</strong>ce<br />

β<br />

λ0ℓ<br />

lim I[Vm]+<br />

2 2 .<br />

Since I + rE is convex, it is weakly lsc over H 1 (R + ). Using the fact that Vm converges H 1 -<br />

weakly and L 2 -strongly towards v0, we conclu<strong>de</strong> that<br />

β − λ0ℓ<br />

I[v0] lim I[Vm] .<br />

2<br />

Finally, with I[v0] βE[v0]=β(1 − ℓ)/2, we obtain<br />

ℓ(λ0 − β) 0.<br />

The assumption (38) then implies ℓ = 0, meaning that vm converges towards v0 strongly in<br />

L 2 (R + ). ✷<br />

To comp<strong>le</strong>te the proof, we have to prove Lemma 5.1. For that purpose, we may assume that<br />

vm converges towards v0 in the following senses: weakly in H 1 (R + ), strongly in L 2 loc (R+ ) and<br />

almost everywhere. Let ψ ∈ C ∞ (R) be such that ψ(y) ≡ 0fory −1 and ψ(y)≡ 1fory 1,<br />

and 0 ψ 1 everywhere. Define also χ := 1 − √ 1 − ψ. We shall choose<br />

Vm(x) := 1 − ψ(x − xm) 1/2 vm(x), zm(x) := χ(x − xm)vm(x),<br />

for a suitab<strong>le</strong> sequence xm. We notice first that Vm → v0 almost everywhere provi<strong>de</strong>d<br />

lim xm =+∞. (40)<br />

This implies that Vm converges weakly in L 2 towards v0. In or<strong>de</strong>r to have the strong convergence,<br />

we only need that E[Vm]→E[v0], which is equiva<strong>le</strong>nt to<br />

<br />

lim<br />

+∞<br />

0<br />

Finally, the difference I[vm]−I[Vm]−I[zm] can be recast into<br />

ψ(x − xm) vm(x) 2 dx = 1 − 2ℓ. (41)<br />

<br />

xm+1<br />

xm−1<br />

B(Vm,zm)dxd,<br />

where B is a bilinear form with smooth coefficients, which involves the arguments and their first<br />

<strong>de</strong>rivatives. In or<strong>de</strong>r to satisfy (39), it is enough to have<br />

lim<br />

<br />

xm+1<br />

xm−1<br />

<br />

v ′ <br />

<br />

m<br />

2 +|vm| 2 dxd = 0. (42)


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 433<br />

There remains to choose xm so as to satisfy all of (40)–(42). To do so, we employ the diagonal<br />

extraction to build a subsequence, still <strong>de</strong>noted by (vm)m1, such that<br />

From (43) and<br />

we infer that<br />

and therefore<br />

m<br />

0<br />

m<br />

lim |vm − v0| 2 dxd = 0. (43)<br />

0<br />

|vm| 2 m<br />

<br />

dxd − 2ℓ =ℜ 〈vm − v0,vm + v0〉 dxd −<br />

A trivial consequence of (43) is that<br />

0<br />

0<br />

+∞<br />

m<br />

|v0| 2 dxd,<br />

m<br />

lim |vm| 2 dxd = 2ℓ, (44)<br />

<br />

lim<br />

+∞<br />

m<br />

0<br />

|vm| 2 dxd = 1 − 2ℓ.<br />

am<br />

lim |vm − v0| 2 dxd = 0<br />

for every sequence such that am


434 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

On the one hand, condition (41) is certainly satisfied for xm ∈ (am + 1,m− 1). On the other<br />

hand, the <strong>le</strong>ngth of this interval tends to +∞, whi<strong>le</strong> the integral<br />

<br />

m−1<br />

am+1<br />

<br />

v ′ <br />

m<br />

2 +|vm| 2 dxd<br />

remains boun<strong>de</strong>d (as it is when we integrate over R + ). Thus there exists a subinterval of <strong>le</strong>ngth<br />

two on which the integral is <strong>le</strong>ss than M/(m − am), thus tends to zero. We can choose xm the<br />

center of this subinterval to satisfy condition (42).<br />

Note. The calcu<strong>la</strong>tions of Section 3.6 show that assumption (38) is often satisfied.<br />

6. Examp<strong>le</strong>s<br />

We point out that in most examp<strong>le</strong>s, we are ab<strong>le</strong> to compute E(τ,η) even for non-real τ 2 ,<br />

thanks to the rather simp<strong>le</strong> structure of the energy <strong>de</strong>nsity. Remark that the Lopatinskiĭ <strong>de</strong>terminants<br />

computed below are not those given by formu<strong>la</strong> (27), since the explicit computation of the<br />

matrix P may be cumbersome. We have followed the more c<strong>la</strong>ssical way, where we compute an<br />

explicit basis of the stab<strong>le</strong> subspace. This method has the disadvantage that such a set of “incoming<br />

mo<strong>de</strong>s” may fail to be a basis at some frequency pairs (τ, η), whence spurious zeroes<br />

of .<br />

6.1. 2-dimensional wave-like systems<br />

Let us fix n = d = 2. The <strong>de</strong>nsity<br />

W0(F ) := 1<br />

|F |2<br />

2<br />

yields the differential operator P = x ⊗ I2. Equation (12) is<br />

u ′′ = τ 2 + η 2 u,<br />

and its stab<strong>le</strong> subspace is given by<br />

u ′ =−ω(τ,η)u, ω(τ,η) :=<br />

<br />

τ 2 + η2 ,<br />

where the square root is that of positive real part. We recognize P(τ,η)=−ω(τ,η)I2, whence<br />

P(0,η)=−|η|I2. In particu<strong>la</strong>r, h(η) =|η|.<br />

Since A ≡ 02 and Λ = I2, wehave<br />

ΛP (τ, η) + A ∗ η =−ω(τ,η)I2.<br />

This is a negative real number when τ 2 ∈ (−h(η) 2 , +∞). Hence ( √ z, η) vanishes only at the<br />

extremity z =−h(η) 2 . In conclusion, the IBVP is well-posed, with <strong>sur</strong>face waves of infinite<br />

energy.


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 435<br />

If we add to W0 a null-form, we do not change the PDEs but we change the boundary operator<br />

B. We thus consi<strong>de</strong>r the <strong>de</strong>nsity<br />

Wκ(F ) := W0(F ) + κ <strong>de</strong>t F.<br />

The boundary operator associated to Wκ is thus<br />

<br />

∂2u1 − κ∂1u2,<br />

u ↦→<br />

.<br />

∂2u2 + κ∂1u1<br />

Using Fourier–Lap<strong>la</strong>ce variab<strong>le</strong>s, we <strong>de</strong><strong>du</strong>ce the Lopatinskiĭ <strong>de</strong>terminant<br />

<br />

−ω −iκη<br />

(τ, η; κ) = <strong>de</strong>t<br />

= τ<br />

iκη −ω<br />

2 + 1 − κ 2 η 2 .<br />

As expected, (·,η; κ) is real analytic, with D(z,η; κ) = z + (1 − κ 2 )η 2 . When η = 0, the root<br />

(κ 2 − 1)η 2 of D(·,η; κ) hasthesamesignasκ 2 − 1. Therefore the hyperbolic IBVP is ill-posed<br />

for |κ| > 1, <strong>de</strong>spite the fact that the corresponding elliptic BVP is well-posed within this range<br />

of parameter (the effect of d = 2).<br />

The transition towards instability occurs when κ =±1. Say that κ = 1. The corresponding<br />

elliptic BVP consists in xu = f for x2 > 0, together with the boundary conditions ∂2u1 −<br />

∂1u2 = g1 and ∂2u2 +∂1u1 = g2. This is the Cauchy prob<strong>le</strong>m for the Cauchy–Riemann equations<br />

in terms of the comp<strong>le</strong>x function γ := ∂1u2 − ∂2u1 + i(∂2u2 + ∂1u1):<br />

∂γ<br />

= f, γ = g.<br />

∂ ¯z<br />

This is the most known ill-posed prob<strong>le</strong>m.<br />

6.2. 3-dimensional wave-like systems<br />

In or<strong>de</strong>r to have a non-trivial family of ill-posed elliptic BVPs, we take n = d = 3 and we<br />

form an energy <strong>de</strong>nsity through the same strategy as above:<br />

W(∇xu) := 1<br />

2 |∇xu| 2 + ∂3u1V ·∇yu2 − ∂3u2V ·∇yu1 + ∂3u2Y ·∇yu3 − ∂3u3Y ·∇yu2<br />

+ ∂3u3Z ·∇yu1 − ∂3u1Z ·∇yu3. (45)<br />

Here, vectors V,Y,Z∈ R2 are given and p<strong>la</strong>y the ro<strong>le</strong> of parameters. The differential operator is<br />

L =−x ⊗ I3. The incoming mo<strong>de</strong>s are given by the equation<br />

v ′ <br />

=−ωv, ω = τ 2 +|η| 2 ,<br />

whence P(τ,η)=−ω(τ,η)I3. The boundary operator is<br />

⎛<br />

⎞<br />

∂3u1 + V ·∇yu2 − Z ·∇yu3<br />

u ↦→ ⎝ ∂3u2 + Y ·∇yu3 − V ·∇yu1 ⎠ ,<br />

∂3u3 + Z ·∇yu1 − Y ·∇yu2


436 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

from which we find<br />

v ∗ H(η)v=|η||v| 2 + 2Y · ηℑ(v3 ¯v2) + 2Z · ηℑ(v1 ¯v3) + 2V · ηℑ(v2 ¯v1). (46)<br />

The Lopatinskiĭ <strong>de</strong>terminant writes<br />

with<br />

<br />

<br />

−ω iV · η −iZ · η <br />

<br />

<br />

(τ, η) = <br />

−iV · η −ω iY · η <br />

<br />

<br />

<br />

iZ · η −iY · η −ω<br />

= ωq(η)− ω 2 ,<br />

q(η) := (V · η) 2 + (Y · η) 2 + (Z · η) 2 . (47)<br />

Let λ±(q) <strong>de</strong>note the eigenvalues of the quadratic form q. The roots of (·,η) are given by<br />

τ =±i|η| (for ω = 0), and by<br />

τ 2 = q(η)−|η| 2 . (48)<br />

On the one hand, this equation has a positive real root for some η if and only if 1 1.<br />

When λ+(q) < 1, the well-posedness of the hyperbolic IBVP is a consequence of Theorem<br />

3.5. Alternately, we may remark that W is convexifiab<strong>le</strong> in the sense of Section 2.1, and<br />

then apply the Hil<strong>le</strong>–Yosida theorem. Thanks to Theorem 2.1, and since d = 3,weonlyhaveto<br />

verify that W(G,F3·) is positive whenever G ∈ M2×3(R) has rank one. To check this property,<br />

we rewrite<br />

where B is the boundary operator and<br />

2W(∇xu) =|Bu| 2 + W1(∇yu),<br />

W1(∇yu) =|∇yu| 2 − (V ·∇yu2 − Z ·∇yu3) 2 − (Y ·∇yu3 − V ·∇yu1) 2<br />

− (Z ·∇yu1 − Y ·∇yu2) 2 .


When G = η ⊗ r, we obtain<br />

Because of |r × σ | |r||σ |, we <strong>de</strong><strong>du</strong>ce<br />

D. Serre / Journal of Functional Analysis 236 (2006) 409–446 437<br />

W1(η ⊗ r) =|η| 2 |r| 2 2 Y · η <br />

<br />

− r<br />

× Z · η .<br />

<br />

V · η<br />

W1(η ⊗ r) |η| 2 − q(η) |r| 2 ,<br />

which is positive if λ+(q) 1. Whence W is convexifiab<strong>le</strong>.<br />

6.3. Isotropic e<strong>la</strong>sticity<br />

We turn now to a practical application in e<strong>la</strong>sticity, where again n = d = 3. We consi<strong>de</strong>r the<br />

case of an isotropic stored energy <strong>de</strong>nsity given by (5). We recall that W is convex if λ 0 and<br />

2λ + 3μ 0, and that it is convexifiab<strong>le</strong> by a TNF if λ 0 and λ + μ 0, the <strong>la</strong>tter condition<br />

being weaker than the former.<br />

The differential operator. The PDEs for x3 > 0are<br />

∂ 2 t<br />

u = λu + (λ + μ)∇ div u. (49)<br />

Let us perform the Lap<strong>la</strong>ce–Fourier transform. In the new unknown, we distinguish the tangential<br />

components v⊥ := (v1,v2) and the normal component vd. Because of isotropy, it is<br />

enough to work with the sca<strong>la</strong>r quantity w := iv⊥ · η:<br />

τ 2 w = λw ′′ − (2λ + μ)|η| 2 w − (λ + μ)|η| 2 vd,<br />

τ 2 vd = (2λ + μ)v ′′<br />

d − λ|η|2vd + (λ + μ)w ′ .<br />

Let us assume that ℜτ >0, so that τ 2 ∈ C \ R − . When looking for mo<strong>de</strong>s in exp(−ωxd), we<br />

find that ω ∈{ωP ,ωS}, where<br />

and<br />

ωP,S :=<br />

τ 2<br />

c2 +|η|<br />

P,S<br />

2<br />

cP := 2λ + μ, cS := √ λ<br />

are the velocities of the pres<strong>sur</strong>e ans shear waves. These are the waves propagated by Eq. (49).<br />

They are distinct provi<strong>de</strong>d λ + μ = 0 and τ = 0. Notice that cP is <strong>la</strong>rger than cS in the convexifiab<strong>le</strong><br />

case, and smal<strong>le</strong>r otherwise. When either λ + μ = 0orτ = 0, we have ωP = ωS, and there<br />

is an additional mo<strong>de</strong> of the form (axd + b)exp(−ωxd).


438 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

Pres<strong>sur</strong>e waves and shear waves. Simp<strong>le</strong> calcu<strong>la</strong>tions show that the mo<strong>de</strong> associated to ωP is<br />

proportional to the vector<br />

<br />

w,vd,w ′ ,v ′ <br />

d P := |η| 2 ,ωP , −ωP |η| 2 , −ω 2 <br />

P ,<br />

and that the mo<strong>de</strong> associated to ωS is proportional to the vector<br />

<br />

w,vd,w ′ ,v ′ <br />

d S := ωS, 1, −ω 2 <br />

S , −ωS .<br />

The boundary condition. In this variational context, the boundary condition is that of zero<br />

normal stress:<br />

λ(∂<strong>du</strong> +∇ud) + μ(div u)ed = 0, x3 = 0.<br />

After the Fourier–Lap<strong>la</strong>ce transformation, we find the equiva<strong>le</strong>nt system<br />

w ′ −|η| 2 vd = 0,<br />

(2λ + μ)u ′ d<br />

+ μw = 0.<br />

The Lopatinskiĭ <strong>de</strong>terminant. We now write that (w, vd,w ′ ,v ′ d ) is a linear combination of the<br />

pres<strong>sur</strong>e and shear mo<strong>de</strong>s:<br />

<br />

w,vd,w ′ ,v ′ <br />

d = α w,vd,w ′ ,v ′ <br />

d S + βw,vd,w ′ ,v ′ <br />

d P .<br />

Writing the boundary condition, we obtain a 2 × 2 linear system, of which the <strong>de</strong>terminant is the<br />

Lopatinskiĭ <strong>de</strong>terminant:<br />

<br />

ω<br />

(τ, η) = <br />

<br />

2 S +|η|2 2|η| 2ωP −2λωS μ|η| 2 − (2λ + μ)ω2 <br />

<br />

<br />

<br />

P<br />

.<br />

This gives<br />

λ −1 (τ, η) = 4|η| 2 <br />

ωSωP − 2|η| 2 +<br />

This formu<strong>la</strong> extends by continuity when ℜτ = 0.<br />

Discussion. If vanishes, then<br />

16|η| 4<br />

<br />

|η| 2 +<br />

τ 2<br />

c 2 S<br />

<br />

|η| 2 +<br />

an equation that can be recast into Q(τ 2 /|η| 2 ) = 0, where<br />

τ 2<br />

c 2 P<br />

<br />

= 2|η| 2 +<br />

<br />

z<br />

Q(z) :=<br />

c2 4 <br />

z<br />

+ 2 − 16<br />

S<br />

c2 <br />

z<br />

+ 1<br />

S c2 P<br />

τ 2 2 . (50)<br />

λ<br />

τ 2<br />

c 2 S<br />

<br />

+ 1 .<br />

4<br />

,


This polynomial satisfies<br />

D. Serre / Journal of Functional Analysis 236 (2006) 409–446 439<br />

Q −c 2 <br />

′ 1<br />

S = 1, Q(0) = 0, Q (0) = 16<br />

c2 −<br />

S<br />

1<br />

c2 <br />

, Q(±∞) =+∞.<br />

P<br />

If λ + μ is positive, W is convexifiab<strong>le</strong> by a TNF, and therefore the homogeneous hyperbolic<br />

IBVP is well-posed. This can be checked directly by showing that does not vanish, except at<br />

boundary points of elliptic type τ =±icR|η|, where −c2 R is the unique root of Q in the interval<br />

(−c2 S , 0). Notice that Q′ (0) is positive in this case, so that such a root must exist. The number<br />

cR, with the dimension of a velocity, is the speed of Ray<strong>le</strong>igh waves, the FESW of this prob<strong>le</strong>m.<br />

If λ + μ is negative, then Q ′ (0) is negative, and Q must have a positive root c2 ,bythe<br />

Intermediate Value theorem. For τ = c|η|, we then have<br />

<br />

4|η| 2 <br />

ωSωP + 2|η| 2 +<br />

τ 2 2 (τ, η) = 0,<br />

λ<br />

where the parenthesis is a positive real number. Therefore (c|η|,η) ≡ 0 and the hyperbolic<br />

IBVP is strongly ill-posed.<br />

Theorem 6.2. When n = d = 3 and the energy <strong>de</strong>nsity is given by (5), with λ>0 and 2λ + μ>0<br />

for uniform rank-one convexity, we have:<br />

(1) The homogeneous hyperbolic IBVP is well-posed provi<strong>de</strong>d λ + μ>0, or equiva<strong>le</strong>ntly<br />

cP >cS.<br />

(2) The hyperbolic IBVP is strongly ill-posed when λ + μ0, certainly give a control of the L2-norms of ∂iui (i = 1,...,3), ∂1u2 and ∂2u1, and<br />

of ∂2u3 + ∂3u2, ∂1u3 + ∂3u1. But since Ω has a boundary, Korn’s inequality does not apply<br />

and this does not give a control neither of ∂3u nor of ∇u3. As remarked in Section 3.5.1, the<br />

coercivity of W comes from a corrector that is neither a null-from (it does yield a boundary<br />

integral), nor differential (the matrix K does <strong>de</strong>pend on η). Following Section 3.5, we may<br />

take<br />

<br />

λ|η|I2 i(ν − λ)η<br />

K(η) =<br />

i(λ− ν)ηT <br />

λ|η|


440 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

with ν := c0 min{λ,λ+ μ} and c0 is some suitab<strong>le</strong> positive number, in<strong>de</strong>pen<strong>de</strong>nt of the other<br />

parameters.<br />

The steady elliptic BVP. Because of the coinci<strong>de</strong>nce between ωP and ωS when τ = 0, the<br />

calcu<strong>la</strong>tion above does not give an answer for the steady BVP. This is <strong>du</strong>e to our choice of the<br />

mo<strong>de</strong>s, because we do not obtain the affine mo<strong>de</strong>s in the limit; it has the effect that becomes<br />

trivial in this limit case. A more careful analysis would give us a non-trivial function .<br />

Here, the differential equations imply<br />

2<br />

∂d −|η| 22 w = 0,<br />

and the same for vd. The mo<strong>de</strong>s that <strong>de</strong>cay at +∞ thus satisfy<br />

<br />

∂d +|η| 2 <br />

w = 0, ∂d +|η| 2 vd = 0.<br />

In other words, we have w = (αxd + β)exp(−|η|xd), with an analogous formu<strong>la</strong> for vd. Going<br />

back to the very ODEs, we find the general formu<strong>la</strong> for the <strong>de</strong>caying mo<strong>de</strong>s:<br />

w = (λ + μ)axd + b |η| 2 e −|η|xd ,<br />

vd = (λ + μ)a|η|xd + (3λ + μ)a + b|η| e −|η|xd .<br />

Inserting this into the boundary conditions, we obtain again a linear 2 × 2 system, whence a<br />

correct Lopatinskiĭ <strong>de</strong>terminant<br />

(0,η)=−4λ(λ + μ)|η| 4 .<br />

This shows that the elliptic BVP is well-posed if and only if λ + μ = 0.<br />

Remark that the ill-posed elliptic BVPs form a hyper<strong>sur</strong>face in the space of parameters (λ, μ),<br />

<strong>de</strong>spite the dimension three of the physical domain. This is <strong>du</strong>e to the isotropy assumption. For<br />

a non-isotropic energy <strong>de</strong>nsity, we expect that these ill-posed steady prob<strong>le</strong>ms form a set of<br />

non-void interior. Then the transition between well-posed and ill-posed hyperbolic IBVPs would<br />

occur along a boundary of this set.<br />

6.4. Phase transition in a van <strong>de</strong>r Waals fluid<br />

Kreiss’ and Sakamoto’s theory of hyperbolic IBVPs has been adapted by A. Majda [12] to<br />

treat the linearized stability of shock waves in systems of conservation <strong>la</strong>ws (a second paper<br />

[13] treats the non-linear stability). The context differs slightly, in the sense that the boundary<br />

condition is coup<strong>le</strong>d with a PDE that <strong>de</strong>scribes the evolution of the shock front. These boundary<br />

conditions come from the linearization of the Rankine–Hugoniot condition. Despite these<br />

technical differences, the same notions of Kreiss–Lopatinskiĭ condition, UKL and Lopatinskiĭ<br />

<strong>de</strong>terminant remain re<strong>le</strong>vant.<br />

Majda’s method is re<strong>le</strong>vant for Lax shocks, where the Rankine–Hugoniot condition provi<strong>de</strong>s<br />

the right number of boundary conditions. It has been adapted by H. Freistüh<strong>le</strong>r [7] to<br />

so-cal<strong>le</strong>d un<strong>de</strong>rcompressive shocks, when additional jump conditions are given in comp<strong>le</strong>ment<br />

of the Rankine–Hugoniot condition. This extension of the theory is particu<strong>la</strong>rly re<strong>le</strong>vant to phase


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 441<br />

boundaries in a van <strong>de</strong>r Waals fluid. Such a fluid is <strong>de</strong>scribed by the usual Eu<strong>le</strong>r equations of an<br />

inviscid, isentropic compressib<strong>le</strong> fluid. The unknowns are the <strong>de</strong>nsity ρ and the velocity v. The<br />

equations govern the conservation of mass and momentum:<br />

∂tρ + div(ρv) = 0, (51)<br />

∂t(ρv) + Div(ρv ⊗ v) +∇xp = 0. (52)<br />

The pres<strong>sur</strong>e is given as a non-monotone equation of state p = π(ρ). The function π is increasing<br />

on (0,ρ−) (gas phase) and (ρ+, +∞) (liquid phase), whi<strong>le</strong> being <strong>de</strong>creasing over (ρ−,ρ+).<br />

Notice that (51), (52) imply formally the conservation of energy<br />

<br />

1<br />

∂t<br />

2 ρ|v|2 <br />

1<br />

+ ρε(ρ) + div<br />

2 ρ|v|2 <br />

+ ρε(ρ) + π(ρ) v = 0, (53)<br />

where ε is the function <strong>de</strong>fined by<br />

dε<br />

dρ<br />

π(ρ)<br />

= .<br />

ρ2 The gas and liquid phases are the states for which system (51), (52) is hyperbolic. The sound<br />

speed c is then the real number √ π ′ (ρ). Given a discontinuity across a smooth hyper<strong>sur</strong>face, it<br />

fulfills the Lax shock condition if the normal velocity V of the shock front satisfies the inequalities<br />

(v · ν + c)+,(v· ν − c)−


442 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

that the linearized prob<strong>le</strong>m around a phase boundary has a variational structure too, although of a<br />

slightly different form than the one studied in the present artic<strong>le</strong>. Thus the main result of [2], the<br />

existence of finite energy <strong>sur</strong>face waves in this linearized prob<strong>le</strong>m is likely to be a consequence<br />

of this structure, in the same spirit as in our Theorem 3.3. We <strong>le</strong>ave this justification for a future<br />

work.<br />

7. General domain and variab<strong>le</strong> coefficient<br />

We turn towards the realistic case where Ω is a smooth open domain in Rd and the quadratic<br />

energy <strong>de</strong>nsity W(x;∇xu) may <strong>de</strong>pend smoothly upon the space variab<strong>le</strong>. For the sake of simplicity,<br />

we limit ourselves to boun<strong>de</strong>d domains. The smoothness required for the boundary and<br />

the coefficients is C2 . The Lagrangian<br />

<br />

<br />

L[u]:= |∂tu| 2 − W(x;∇xu) dxdt<br />

Ω<br />

<strong>de</strong>fines an initial boundary-value prob<strong>le</strong>m in Ω. From this IBVP, we <strong>de</strong>fine a family of IBVPs<br />

with constant coefficients in half-spaces, parametrized by the e<strong>le</strong>ments of the boundary ∂Ω.To<br />

each point x0 ∈ ∂Ω, we associate the Lagrangian<br />

<br />

<br />

L[u]:= |∂tu| 2 − W(x0;∇xu) dxdt,<br />

ω(x0)<br />

where ω(x0) is the half-space with the same outer normal ν(x0) at x0 as Ω:<br />

ω(x0) ={x: (x − x0) · ν(x0)


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 443<br />

This result follows immediately from the Hil<strong>le</strong>–Yosida theorem and the following estimate.<br />

Theorem 7.2. Un<strong>de</strong>r the assumptions of Theorem 7.1, there exists two positive constants ɛ and C,<br />

such that<br />

Proof. If x1 ∈ Ω, <strong>le</strong>t us <strong>de</strong>fine<br />

W[u] ɛ∇xu 2<br />

L2 − Cu2<br />

(ω) L2 (ω) .<br />

<br />

Y[x1; u]:=<br />

R d<br />

W(x1;∇xu) dx.<br />

By rank-one convexity, there exists a positive number α(x1) such that Y[x1; u] α(x1)∇xu 2 ,<br />

where ·stands for the L 2 (R d )-norm. Since W varies smoothly with x, and since Ω is compact,<br />

we have<br />

inf<br />

x1∈Ω α(x1)>0.<br />

Likewise, Theorem 3.5 tells that if x0 is a boundary point, then W[x0; u] dominates<br />

<br />

β(x0) ∇xu 2 dx<br />

ω(x0)<br />

for some positive number β(x0). Again, continuity and compactness of the boundary imply<br />

inf<br />

x0∈∂Ω β(x0)>0.<br />

In short, there exists a number γ>0 such that for every interior point x1 or boundary point x0,<br />

there holds true<br />

Y[x1; u] γ ∇xu 2 , W[x0; u] γ ∇xu 2 ,<br />

where we employ the L2-norm, either of Rd or the half-space ω(x0), according to the context.<br />

Let the ball B = B(x1; r) be contained in Ω. Ifu∈ H 1 (Ω) has support contained in B,<br />

extension by zero yields a ũ ∈ H˙ 1 (Rd ). Then<br />

<br />

<br />

W[u]=Y[x1;ũ]+ W(x;∇xu) − W(x1;∇xu) dx.<br />

Because of smoothness, we <strong>de</strong>rive<br />

B<br />

W[u] γ ∇xu 2 − O(r)∇xu 2 .<br />

Therefore, choosing r small enough, we are certain that<br />

W[u] γ 2<br />

∇xu<br />

2


444 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

holds true. We point out that the same positive r may be chosen for every point x1. To treat the<br />

case of a boundary point x0, we employ the same argument, but we need another technical tool.<br />

If r is small enough, then B ∩ Ω is diffeomorphic to a half-ball<br />

B+ := x: (x − x0) · ν(x0)0 as above. By compactness we may cover Ω by a finite col<strong>le</strong>ction of balls Bj :=<br />

B(x j ; r) where either Bj ⊂ Ω or x j ∈ ∂Ω (here it is useful to take r <strong>le</strong>ss than the minimum of<br />

the curvature of the boundary). Let (ρ1,...,ρN) be a partition of unity over Ω adapted to the<br />

B ′ j s, with ρj 0. If u ∈ H 1 (Ω), we <strong>de</strong>fine<br />

Using the po<strong>la</strong>r form Ψ of W ,wehave<br />

uj := ρj u, ujk := √ ρj ρk u.


D. Serre / Journal of Functional Analysis 236 (2006) 409–446 445<br />

W(x;∇u) = <br />

W(x;∇uj ) + 2 <br />

Ψ(x;∇uj , ∇uk)<br />

j<br />

j


446 D. Serre / Journal of Functional Analysis 236 (2006) 409–446<br />

where Λ, S and Σ are Hermitian, S is linear and Σ quadratic in η and<br />

ξ 2 Λ(x0) − ξS(x0; η) + Σ(x0; η) α ξ 2 +|η| 2 , ∀(ξ, η) ∈ R d , ∀x0 ∈ ∂Ω,<br />

for some positive α (strict rank-one convexity).<br />

Acknow<strong>le</strong>dgments<br />

This work was initiated by a conversation with Luc Tartar, whose fundamental questions are<br />

unfortunately not solved here. The i<strong>de</strong>a that steady ill-posed prob<strong>le</strong>ms have something to do with<br />

the transition to instability was suggested by Constantin Dafermos. I am grateful to both of them<br />

for their questions and their interest in the topic. Several crucial results were obtained <strong>du</strong>ring a<br />

stay in the Marais Poitevin; I thank Pasca<strong>le</strong> for having <strong>le</strong>t me steal hours from our nice vacations.<br />

References<br />

[1] S. Agmon, A. Douglis, L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations<br />

satisfying general boundary conditions. II, Comm. Pure Appl. Math. 17 (1964) 35–92.<br />

[2] S. Benzoni-Gavage, Stability of multi-dimensional phase transitions in a van <strong>de</strong>r Waals fluid, Nonlinear Anal. 31<br />

(1998) 243–263.<br />

[3] S. Benzoni-Gavage, D. Serre, Multi-dimensional Hyperbolic Partial Differential Equations; First Or<strong>de</strong>r Systems and<br />

Applications, Oxford Univ. Press, Oxford, Fall 2006, in press.<br />

[4] S. Benzoni-Gavage, F. Rousset, D. Serre, K. Zumbrun, Generic types and transitions in hyperbolic initial-boundary<br />

value prob<strong>le</strong>ms, Proc. Roy. Soc. Edinburgh Sect. A 132 (2002) 1073–1104.<br />

[5] L.C. Evans, Weak Convergence Methods for Nonlinear Partial Differential Equations, CBMS Reg. Conf. Ser. Math.,<br />

vol. 74, Amer. Math. Soc., Provi<strong>de</strong>nce, RI, 1990.<br />

[6] H. Fan, M. S<strong>le</strong>mrod, Dynamic flows with liquid/vapor phase transitions, in: S. Fried<strong>la</strong>n<strong>de</strong>r, D. Serre (Eds.), Handbook<br />

of Mathematical Fluid Dynamics, vol. I, North-Hol<strong>la</strong>nd, Amsterdam, 2002, pp. 373–420.<br />

[7] H. Freistüh<strong>le</strong>r, The persistence of i<strong>de</strong>al shock waves, Appl. Math. Lett. 7 (1994) 7–11.<br />

[8] R. Gardner, K. Zumbrun, The gap <strong>le</strong>mma and geometric criteria for instability of viscous shock profi<strong>le</strong>s, Comm.<br />

Pure Appl. Math. 51 (1998) 797–855.<br />

[9] R. Hersh, Mixed prob<strong>le</strong>ms in several variab<strong>le</strong>s, J. Math. Mech. 12 (1963) 317–334.<br />

[10] H.-O. Kreiss, Initial boundary value prob<strong>le</strong>ms for hyperbolic systems, Comm. Pure Appl. Math. 23 (1970) 277–298.<br />

[11] Y. Lopatinskiĭ, On a method of re<strong>du</strong>cing boundary prob<strong>le</strong>ms for a system of differential equations of elliptic type to<br />

regu<strong>la</strong>r integral equations, Ukrain. Mat. Zh. 5 (1953) 123–151 (in Russian).<br />

[12] A. Majda, The stability of multidimensional shock fronts, Mem. Amer. Math. Soc. 41 (275) (1983).<br />

[13] A. Majda, The existence of multidimensional shock fronts, Mem. Amer. Math. Soc. 43 (281) (1983).<br />

[14] G. Métivier, The block structure condition for symmetric hyperbolic systems, Bull. London Math. Soc. 32 (2000)<br />

689–702.<br />

[15] R. Sakamoto, Hyperbolic Boundary Value Prob<strong>le</strong>ms, Cambridge Univ. Press, Cambridge, 1982.<br />

[16] D. Serre, Formes quadratiques et calcul <strong>de</strong>s variations, J. Math. Pures Appl. 62 (1983) 177–196.<br />

[17] D. Serre, La transition vers l’instabilité pour <strong>le</strong>s on<strong>de</strong>s <strong>de</strong> choc multidimensionnel<strong>le</strong>s, Trans. Amer. Math. Soc. 353<br />

(2001) 5071–5093.<br />

[18] F.J. Terpstra, Die Darstellung biquadratischer Formen als Summen von Quadraten mit Anwen<strong>du</strong>ng auf die Variationsrechnung,<br />

Math. Ann. 116 (1938) 166–180.<br />

[19] K. Zumbrun, D. Serre, Viscous and inviscid stability of multidimensional shock fronts, Indiana Univ. Math. J. 48<br />

(1999) 937–992.


Journal of Functional Analysis 236 (2006) 447–456<br />

www.elsevier.com/locate/jfa<br />

On the super fixed point property in pro<strong>du</strong>ct spaces<br />

Andrzej Wi´snicki<br />

Institute of Mathematics, Maria Curie-Skłodowska University, 20-031 Lublin, Po<strong>la</strong>nd<br />

Received 29 October 2005; accepted 10 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 2 May 2006<br />

Communicated by G. Pisier<br />

Abstract<br />

We prove that if F is a finite-dimensional Banach space and X has the super fixed point property for<br />

nonexpansive mappings, then F ⊕ X has the super fixed point property with respect to a <strong>la</strong>rge c<strong>la</strong>ss of<br />

norms including all lp norms, 1 p


448 A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456<br />

sufficient condition for X to have WFPP. (Recall that a Banach space X has normal structure if<br />

the Chebyshev radius rC(C) < diam C for all boun<strong>de</strong>d closed convex subsets C of X consisting<br />

of more than one point.)<br />

The prob<strong>le</strong>m of whether FPP or WFPP is preserved un<strong>de</strong>r direct sum of Banach spaces is an<br />

old one. In 1968 Belluce, Kirk and Steiner [3] proved that the direct sum of two Banach spaces<br />

with normal structure, endowed with the “maximum” norm, also has normal structure. Since<br />

then, the preservation of normal structure and conditions which guarantee normal structure have<br />

been studied extensively and the prob<strong>le</strong>m is now quite well un<strong>de</strong>rstood (see [28–30,35] and<br />

the references given there). But the situation is much more difficult if at <strong>le</strong>ast one of these spaces<br />

<strong>la</strong>cks (weak) normal structure. To the author’s know<strong>le</strong>dge the only known results are given in [10,<br />

19,27,30,39]. (We should also mention [5,21,23,26,37], where nonexpansive mappings <strong>de</strong>fined<br />

on rectang<strong>le</strong>s C1 × C2 are consi<strong>de</strong>red.)<br />

In this paper we prove in Section 3 that if F is finite-dimensional and X has the super fixed<br />

point property (see Section 2 for the <strong>de</strong>finition), then F ⊕ X has the super fixed point property<br />

with respect to a <strong>la</strong>rge c<strong>la</strong>ss of norms including all strictly monotone norms. This provi<strong>de</strong>s a<br />

solution to the “super-version” of the prob<strong>le</strong>m of M.A. Khamsi [21, p. 999] for this c<strong>la</strong>ss of<br />

norms. Some consequences of this theorem and examp<strong>le</strong>s of Banach spaces with the super fixed<br />

point property are given in Section 4.<br />

2. Basic notions and tools<br />

In this section we shall briefly recall basic tools used in the sequel. For more <strong>de</strong>tails we refer<br />

the rea<strong>de</strong>r to [1,2,7,14,24,32,34]. Let X and Y be Banach spaces and <strong>le</strong>t 0


A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456 449<br />

Here lim U <strong>de</strong>notes the ultralimit over U. One can prove that the quotient norm on (X)U is given<br />

by<br />

<br />

(xn)U<br />

= limU xn,<br />

where (xn)U is the equiva<strong>le</strong>nce c<strong>la</strong>ss of (xn). It is also c<strong>le</strong>ar that X is isometric to a subspace<br />

of (X)U by the embedding x → (x)U . We shall not distinguish between x and (x)U .<br />

The connection between finite representability and ultrapowers was in<strong>de</strong>pen<strong>de</strong>ntly observed<br />

by Henson and Moore [17], and Stern [36] (see also [1,15,34]).<br />

Theorem 2.2. A Banach space Y is finitely representab<strong>le</strong> in X if and only if there exists an<br />

ultrafilter U such that Y is isometric to a subspace of (X)U .<br />

From Theorem 2.2 we obtain immediately<br />

Proposition 2.3. For any Banach space X:<br />

(i) X is superref<strong>le</strong>xive iff each ultrapower (X)U is ref<strong>le</strong>xive;<br />

(ii) X has SFPP iff each ultrapower (X)U has FPP.<br />

We shall also need the following result concerning iterated ultrapowers, see for instance<br />

[34, Theorem 13.2].<br />

Theorem 2.4. Let U and V be ultrafilters on I and J , respectively, and <strong>le</strong>t X be a Banach<br />

space. Then there exists an ultrafilter W <strong>de</strong>fined on I × J such that ((X)U )V is isometric to the<br />

ultrapower (X)W .<br />

Assume now that C is a nonempty weakly compact, convex subset of a Banach space X and<br />

that there exists a nonexpansive mapping T : C → C without fixed points. Then, by Zorn <strong>le</strong>mma,<br />

there exists a minimal convex and weakly compact set K ⊂ C which is invariant un<strong>de</strong>r T and<br />

which is not a sing<strong>le</strong>ton. Recall that a sequence (xn) is cal<strong>le</strong>d an approximate fixed point sequence<br />

for T if<br />

lim<br />

n→∞ Txn − xn=0.<br />

The following <strong>le</strong>mma was in<strong>de</strong>pen<strong>de</strong>ntly proved by Goebel [13] and Karlovitz [20].<br />

Lemma 2.5. If (xn) is an approximate fixed point sequence for T in K, then<br />

for all x ∈ K.<br />

lim<br />

n→∞ xn − x=diam K<br />

In 1980 the Banach space ultrapower construction was applied in fixed point theory by Maurey,<br />

see [31]. Let U be a free ultrafilter on N and <strong>de</strong>note by K ⊂ (X)U the set<br />

K = (xn)U ∈ (X)U : xn ∈ K for all n ∈ N .


450 A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456<br />

We may extend the mapping T by setting T ((xn)U ) = (T xn)U . It is not difficult to see that<br />

T : K → K is a well-<strong>de</strong>fined nonexpansive mapping. Moreover, Fix T , the set of fixed points<br />

of T , is nonempty and is characterized as those points from K which are represented by sequences<br />

(xn) in K for which lim U Txn − xn=0. We can now rephrase Lemma 2.5 as follows.<br />

Lemma 2.6. For every x ∈ K and y ∈ Fix T<br />

Let (xn) be a sequence in K and put<br />

x − y=diam K.<br />

H(xn) = (xkn )U : (xkn ) is a subsequence of (xn) .<br />

It is not difficult to see that if (xn) is an approximate fixed point sequence for T , then<br />

H(xn) ⊂ Fix T . The following <strong>le</strong>mma is a reformu<strong>la</strong>tion of known facts, see the arguments in<br />

[25, Theorem 4.4], [38, Theorem 3.1], [1, p. 86]. We sketch the proof for the convenience of the<br />

rea<strong>de</strong>r.<br />

Lemma 2.7. Let X be superref<strong>le</strong>xive and assume that (xn) is a sequence in K which converges<br />

weakly to 0. Then<br />

0 ∈ convH(xn).<br />

Proof. It is sufficient to prove that 0 is in the weak clo<strong>sur</strong>e of H(xn). Letε>0 and <strong>le</strong>t<br />

F1, F2,...,Fm be functionals in (X) ∗ U . Since X is superref<strong>le</strong>xive, then, by the Henson–Moore<br />

theorem [16,17], see also [15,34], there exist sequences (f 1 n ), (f 2 n ),...,(fm n ) ⊂ X∗ such that<br />

<br />

F1 (un)U = limU f 1 n (un), ...,<br />

<br />

Fm (un)U = limU f m n (un)<br />

for every (un)U ∈ X. Since (xn) converges weakly to 0, there exists an increasing sequence of<br />

integers (kn) such that<br />

for all n ∈ N. Hence<br />

<br />

f 1 n (xkn ) ε, ...,<br />

<br />

f m n (xkn ) ε<br />

<br />

F1 (xkn )U<br />

<br />

ε, ..., Fm (xkn )U<br />

<br />

ε<br />

and it follows that (xkn )U ∈ H(xn) belongs to the weak neighbourhood of 0. This comp<strong>le</strong>tes the<br />

proof since ε>0 and F1, F2,...,Fm ∈ (X) ∗ U were arbitrary. ✷<br />

3. Main theorem<br />

Let X and Y be Banach spaces and p ∈[1, ∞). We <strong>de</strong>note by X ⊕p Y the pro<strong>du</strong>ct space<br />

X ⊕ Y equipped with the norm<br />

<br />

(x, y) = x p +y p 1/p .


A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456 451<br />

In [21] a <strong>la</strong>rger c<strong>la</strong>ss of norms were consi<strong>de</strong>red. We shall also consi<strong>de</strong>r another c<strong>la</strong>ss of norms.<br />

A norm ·on X ⊕ Y is said to be of type (UL) if:<br />

(i) (x, 0)=x and (0,y)=y for every x ∈ X, y ∈ Y ;<br />

(ii) x (x, y) and y (x, y) for every x ∈ X, y ∈ Y ;<br />

(iii) there exists M>0 such that for every x,x ′ ∈ X and y ∈ Y we have<br />

x−x ′ M (x, y) − (x ′ ,y) .<br />

It was asked in [21, p. 999] whether WFPP (FPP) is preserved un<strong>de</strong>r the pro<strong>du</strong>ct X ⊕ Y if one of<br />

these spaces is finite-dimensional.<br />

Below we solve the “super-version” of this prob<strong>le</strong>m for strictly monotone norms and norms<br />

of type (UL). Notice that (X ⊕ Y)U is isometric to (X)U ⊕ (Y )U for any ultrafilter U and<br />

<br />

<br />

(xn)U ,(yn)U<br />

(X)U ⊕(Y )U = <br />

(xn,yn)U<br />

<br />

(X⊕Y)U = lim <br />

(xn,yn)<br />

U<br />

. X⊕Y<br />

From now on we shall use the same symbol ·for the norms above.<br />

Lemma 3.1. If the pro<strong>du</strong>ct X ⊕ Y is endowed with a norm of type (UL), then the in<strong>du</strong>ced space<br />

(X)U ⊕ (Y )U is endowed with a norm of type (UL), too.<br />

Proof. It is not difficult to see that<br />

<br />

(xn)U , 0 <br />

= lim(xn, 0)<br />

U<br />

= lim xn=<br />

U <br />

(xn)U<br />

<br />

for every (xn)U ∈ (X)U and, simi<strong>la</strong>rly, (0,(yn)U )=(yn)U for (yn)U ∈ (Y )U . To prove (ii)<br />

notice that<br />

Hence<br />

lim U xn lim U<br />

<br />

(xn,yn) <br />

and limU yn lim (xn,yn)<br />

U<br />

.<br />

<br />

(xn)U<br />

(xn)U ,(yn)U<br />

and (yn)U (xn)U ,(yn)U<br />

<br />

for every (xn)U ∈ (X)U ,(yn)U ∈ (Y )U . Finally<br />

lim xn−lim x<br />

U U ′ <br />

n M lim (xn,yn)<br />

U<br />

<br />

− limU ′<br />

(x n ,yn) <br />

which proves (iii). ✷<br />

We are now in a position to prove the following theorem.<br />

Theorem 3.2. Let X be a Banach space with the super fixed point property and F be a finitedimensional<br />

space. Then F ⊕ X, endowed with a norm of type (UL), has the super fixed point<br />

property.


452 A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456<br />

Proof. Assume conversely that F ⊕ X does not have SFPP. Then, by Theorem 2.2, there exists<br />

an ultrafilter V such that F ⊕ (X)V ∼ = (F ⊕ X)V does not have FPP. Since X has SFPP, it follows<br />

from Theorem 2.1 that X is superref<strong>le</strong>xive and, by Proposition 2.3, (X)V is superref<strong>le</strong>xive. Hence<br />

F ⊕ (X)V is superrefexive, too. To simplify notation, put Y = (X)V.<br />

Let T : C → C be a nonexpansive mapping <strong>de</strong>fined on a boun<strong>de</strong>d closed and convex subset C<br />

of F ⊕ Y without fixed points. Since F ⊕ Y is superref<strong>le</strong>xive, C is weakly compact and hence<br />

there exists a closed convex set K ⊂ C which is minimal invariant un<strong>de</strong>r T .Let(xn,yn) be<br />

an approximate fixed point sequence for T in K. We may assume that diam K = 1 and that<br />

((xn,yn)) converges weakly to (0, 0) ∈ K. Then, by the Goebel–Karlovitz Lemma 2.5,<br />

<br />

lim (xn,yn) = 1.<br />

n→∞<br />

Let U be a free ultrafilter on N. Since F is finite-dimensional, (xn) converges strongly to 0. Thus<br />

(xkn )U = 0 for every subsequence (xkn ) of (xn) and<br />

H(xn,yn) := (xkn ,ykn )U : (xkn ,ykn ) is a subsequence of (xn,yn) <br />

By Lemma 2.7,<br />

= (0,ykn )U : (ykn ) is a subsequence of (yn) .<br />

(0, 0) ∈ convH(xn,yn)<br />

and, since H(xn,yn) ⊂ Fix T , there exist distinct points (0,ykn )U ,(0,yln )U ∈ H(xn,yn) such that<br />

<br />

0, 1 2ykn + 1 2yln <br />

<br />

U = 1<br />

ykn 2 + 1 2yln <br />

<br />

U = r 0 and put<br />

D = B (0,ykn )U , 1 2 d ∩ B (0,yln )U , 1 2 d ∩ B(K,r) ∩ K.<br />

It is easy to see that D = ∅is convex and T(D)⊂ D. We show that if (an,bn)U ∈ D, then<br />

(an)U = 0. In<strong>de</strong>ed,<br />

and thus<br />

Hence<br />

<br />

(ykn<br />

<br />

− yln )U<br />

(ykn<br />

and it follows from Lemma 3.1 that (an)U = 0.<br />

<br />

− bn)U<br />

+ (bn − yln )U<br />

<br />

<br />

= <br />

(0,ykn − bn)U<br />

+ (0,bn − yln )U<br />

<br />

<br />

<br />

(−an,ykn − bn)U<br />

+ (an,bn − yln )U<br />

<br />

<br />

= <br />

(0,ykn − yln )U<br />

= (ykn − yln )U<br />

<br />

<br />

(0,bn − yln )U<br />

<br />

= (an,bn − yln )U<br />

<br />

.<br />

<br />

0,(bn− yln )U<br />

= (an)U ,(bn− yln )U


Therefore,<br />

A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456 453<br />

D1 = (un)U : (0,un)U ∈ D <br />

is a subset of Y which is isometric to D.<br />

Define T1 : D1 → D1 as<br />

<br />

T1 (un)U = PrY<br />

T <br />

(0,un)U ,<br />

where PrY <strong>de</strong>notes the standard projection onto Y . Notice that T1 is nonexpansive. By assumption,<br />

X has SFPP, and it follows from Theorem 2.4 that (Y )U = ((X)V)U has FPP. Thus there<br />

exists (vn)U ∈ D1 such that<br />

<br />

T1 (vn)U = (vn)U<br />

and consequently<br />

T <br />

(0,vn)U = (0,vn)U .<br />

But this contradicts Lemma 2.6 because (0,vn)U ∈ D ⊂ B(K,r). ✷<br />

Let us consi<strong>de</strong>r another, more symmetric, c<strong>la</strong>ss of norms. A norm ·Z on R 2 is said to be<br />

strictly monotone if<br />

<br />

(x1,y1) Z < (x2,y2) Z<br />

whenever |x1| |x2|, |y1| < |y2| or |x1| < |x2|, |y1| |y2|. We say that a norm on X ⊕ Y is<br />

strictly monotone if<br />

<br />

(x, y) = x, y Z<br />

and the norm ·Z is strictly monotone.<br />

Strictly monotone norms do not necessarily satisfy the condition (i):<br />

<br />

(x, 0) =x and (0,y) =y for every x ∈ X, y ∈ Y.<br />

However, it is not difficult to see that the proof of Theorem 3.2 is also valid in this case.<br />

Theorem 3.3. Let X be a Banach space with the super fixed point property and F be a finitedimensional<br />

space. Then F ⊕ X, endowed with a strictly monotone norm, has the super fixed<br />

point property.<br />

In particu<strong>la</strong>r, the result holds for l p -pro<strong>du</strong>cts F ⊕p X, p ∈[1, ∞).


454 A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456<br />

4. Consequences<br />

It is a long-standing open prob<strong>le</strong>m whether all superref<strong>le</strong>xive spaces have FPP (and hence<br />

SFPP). However, there are two important c<strong>la</strong>sses of spaces which have the super fixed point<br />

property. Recall first that<br />

ɛ0(X) = sup ɛ 0: δX(ɛ) = 0 ,<br />

where δX <strong>de</strong>notes the mo<strong>du</strong>lus of convexity of a Banach space X. It is well known that ɛ0(X) < 2<br />

implies superref<strong>le</strong>xivity of X. A recent result of García Falset et al. [12] (see also [32]), states that<br />

if ɛ0(X) < 2, then X has FPP. In consequence, X has SFPP since ɛ0(X) = ɛ0((X)U ). Combining<br />

it with Theorems 3.2 and 3.3 we obtain<br />

Proposition 4.1. Let X be a Banach space with ɛ0(X) < 2 and <strong>le</strong>t F be a finite-dimensional<br />

space. Then F ⊕ X, endowed with a norm of type (UL) or a strictly monotone norm, has SFPP.<br />

In 1997 Prus [33] intro<strong>du</strong>ced the notion of uniformly noncreasy spaces. A real Banach space<br />

X is uniformly noncreasy if for every ε>0 there is δ>0 such that if f,g ∈ SX∗ and f −g ε,<br />

then diam S(f,g,δ) ε, where<br />

S(f,g,δ) = x ∈ BX: f(x) 1 − δ ∧ g(x) 1 − δ .<br />

To be precise, we put diam ∅=0. It is well known that uniformly convex as well as uniformly<br />

smooth spaces are uniformly noncreasy. The Bynum space l2,∞ , which is l2 space endowed with<br />

the norm<br />

x2,∞ = max x + 2, x − <br />

2 ,<br />

(see [4]), and the space X √ 2 , which is l2 space endowed with the norm<br />

x√ 2 = maxx2, √ <br />

2x∞ ,<br />

(see [3]), are examp<strong>le</strong>s of uniformly noncreasy spaces without normal structure. It was proved<br />

in [33] that all uniformly noncreasy spaces are superref<strong>le</strong>xive and have SFPP. This yields<br />

Proposition 4.2. Let X be uniformly noncreasy and <strong>le</strong>t F be a finite-dimensional space. Then<br />

F ⊕ X, endowed with a norm of type (UL) or a strictly monotone norm, has SFPP.<br />

Remark. The above result is also valid for some generalizations of uniformly noncreasy spaces<br />

given in [8,9,11].<br />

Other examp<strong>le</strong>s of spaces with the super fixed point property are given by the author’s result<br />

[39, Theorem 2.3]. In particu<strong>la</strong>r, the l p -pro<strong>du</strong>cts of uniformly noncreasy spaces have SFPP.<br />

Banach spaces X with the property that R ⊕ X has WFPP were studied in [27]. The following<br />

theorem was established for the l 1 -norm but the proof works for all strictly monotone norms.


A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456 455<br />

Theorem 4.3. (See [27, Theorem 1]) Let X be a Banach space which is uniformly convex in<br />

every direction. If Y is a Banach space such that R ⊕ Y , endowed with a strictly monotone norm,<br />

has WFPP, then X ⊕ Y also has WFPP.<br />

Theorems 3.3 and 4.3 give<br />

Proposition 4.4. Let X be a Banach space which is uniformly convex in every direction. If Y has<br />

SFPP, then X ⊕ Y , endowed with a strictly monotone norm, has WFPP.<br />

We conclu<strong>de</strong> with the observation concerning spaces with the Schur property. The proof is<br />

simi<strong>la</strong>r to that in Section 3.<br />

Proposition 4.5. Let X be a Banach space with the Schur property and <strong>le</strong>t Y has SFPP. Then<br />

X ⊕ Y , endowed with a norm of type (UL) or a strictly monotone norm, has WFPP.<br />

Prob<strong>le</strong>m. It is natural to ask whether the results of this paper are valid for the pro<strong>du</strong>ct space<br />

endowed with the “maximum” norm or, more generally, with a monotone norm.<br />

Acknow<strong>le</strong>dgment<br />

The author thanks the referee for valuab<strong>le</strong> comments on the manuscript.<br />

References<br />

[1] A.G. Aksoy, M.A. Khamsi, Nonstandard Methods in Fixed Point Theory, Springer-Ver<strong>la</strong>g, New York, 1990.<br />

[2] J.M. Ayerbe To<strong>le</strong>dano, T. Domínguez Benavi<strong>de</strong>s, G. López Acedo, Mea<strong>sur</strong>es of Noncompactness in Metric Fixed<br />

Point Theory, Birkhäuser, Basel, 1997.<br />

[3] L.P. Belluce, W.A. Kirk, E.F. Steiner, Normal structure in Banach spaces, Pacific J. Math. 26 (1968) 433–440.<br />

[4] W.L. Bynum, A c<strong>la</strong>ss of spaces <strong>la</strong>cking normal structure, Compos. Math. 25 (1972) 233–236.<br />

[5] S. Dhompongsa, A. Kaewcharoen, A. Kaewkhao, Fixed point property of direct sums, Nonlinear Anal. 63 (2005)<br />

e2177–e2188.<br />

[6] D. van Dulst, A.J. Pach, On f<strong>la</strong>tness and some ergodic super-properties of Banach spaces, Indag. Math. 43 (1981)<br />

153–164.<br />

[7] J. Elton, P.K. Lin, E. O<strong>de</strong>ll, S. Szarek, Remarks on the fixed point prob<strong>le</strong>m for nonexpansive maps, in: R. Sine (Ed.),<br />

Fixed Points and Nonexpansive Mappings, in: Contemp. Math., vol. 18, Amer. Math. Soc., Provi<strong>de</strong>nce, RI, 1983,<br />

pp. 87–120.<br />

[8] H. Fetter, B. Gamboa <strong>de</strong> Buen, (r,k,l)-Somewhat uniformly noncreasy Banach spaces, in: J. García Falset,<br />

E. Lloréns Fuster, B. Sims (Eds.), Fixed Point Theory and Applications, Yokohama Publ., Yokohama, 2004, pp. 71–<br />

80.<br />

[9] H. Fetter, B. Gamboa <strong>de</strong> Buen, J. García Falset, Banach spaces which are somewhat uniformly noncreasy, J. Math.<br />

Anal. Appl. 285 (2003) 444–455.<br />

[10] J. García Falset, E. Lloréns Fuster, Normal structure and fixed point property, G<strong>la</strong>sgow Math. J. 38 (1996) 29–37.<br />

[11] J. García Falset, E. Lloréns Fuster, E.M. Mazcuñan Navarro, Banach spaces which are r-uniformly noncreasy,<br />

Nonlinear Anal. 53 (2003) 957–975.<br />

[12] J. García Falset, E. Lloréns Fuster, E.M. Mazcuñan Navarro, Uniformly nonsquare Banach spaces have the fixed<br />

point property for nonexpansive mappings, J. Funct. Anal. 233 (2006) 494–514.<br />

[13] K. Goebel, On the structure of minimal invariant sets for nonexpansive mappings, Ann. Univ. Mariae Curie-<br />

Skłodowska Sect. A 29 (1975) 73–77.<br />

[14] K. Goebel, W.A. Kirk, Topics in Metric Fixed Point Theory, Cambridge Univ. Press, Cambridge, 1990.<br />

[15] S. Heinrich, Ultrapro<strong>du</strong>cts in Banach space theory, J. Reine Angew. Mat. 313 (1980) 72–104.<br />

[16] C.W. Henson, L.C. Moore Jr., The nonstandard theory of topological vector spaces, Trans. Amer. Math. Soc. 172<br />

(1972) 405–435.


456 A. Wi´snicki / Journal of Functional Analysis 236 (2006) 447–456<br />

[17] C.W. Henson, L.C. Moore Jr., Subspaces of the nonstandard hull of a normed space, Trans. Amer. Math. Soc. 197<br />

(1974) 131–143.<br />

[18] R.C. James, Some self-<strong>du</strong>al properties of normed linear spaces, in: Symp. on Infinite-Dimensional Topology, Ann.<br />

Math. Stud. 69 (1972) 159–175.<br />

[19] M.Á. Japón Pineda, A new constant in Banach spaces and stability of the fixed point property, in: J. Krzy˙z, S. Prus<br />

(Eds.), Proc. of Workshop on Fixed Point Theory, Ann. Univ. Mariae Curie-Skłodowska Sect. A 51 (1997) 135–141.<br />

[20] L.A. Karlovitz, Existence of fixed points of nonexpansive mappings in a space without normal structure, Pacific J.<br />

Math. 66 (1976) 153–159.<br />

[21] M.A. Khamsi, On normal structure, fixed point property and contractions of type (γ ), Proc. Amer. Math. Soc. 106<br />

(1989) 995–1001.<br />

[22] W.A. Kirk, A fixed point theorem for mappings which do not increase distances, Amer. Math. Monthly 72 (1965)<br />

1004–1006.<br />

[23] W.A. Kirk, C. Martinez-Yanez, Nonexpansive and locally nonexpansive mappings in pro<strong>du</strong>ct spaces, Nonlinear<br />

Anal. 12 (1988) 719–725.<br />

[24] W.A. Kirk, B. Sims (Eds.), Handbook of Metric Fixed Point Theory, Kluwer Aca<strong>de</strong>mic Publ., Dordrecht, 2001.<br />

[25] W.A. Kirk, C. Martinez-Yanez, S.S. Shin, Asymptotically nonexpansive mappings, Nonlinear Anal. 33 (1998) 1–12.<br />

[26] T. Kuczumow, Fixed point theorems in pro<strong>du</strong>ct spaces, Proc. Amer. Math. Soc. 108 (1990) 727–729.<br />

[27] T. Kuczumow, S. Reich, M. Schmidt, A fixed point property of l1-pro<strong>du</strong>ct spaces, Proc. Amer. Math. Soc. 119<br />

(1993) 457–463.<br />

[28] T. Lan<strong>de</strong>s, Permanence properties of normal structure, Pacific J. Math. 110 (1984) 125–143.<br />

[29] T. Lan<strong>de</strong>s, Normal structure and the sum-property, Pacific J. Math. 123 (1986) 127–147.<br />

[30] G. Marino, P. Pietrama<strong>la</strong>, H.K. Xu, Geometrical conditions in pro<strong>du</strong>ct spaces, Nonlinear Anal. 46 (2001) 1063–<br />

1071.<br />

[31] B. Maurey, Points fixes <strong>de</strong>s contractions <strong>de</strong> certains faib<strong>le</strong>ment compacts <strong>de</strong> L 1 , in: Semin. Anal. Fonct. 1980–1981,<br />

Éco<strong>le</strong> Polytechnique, Pa<strong>la</strong>iseau, 1981.<br />

[32] E.M. Mazcuñan Navarro, Geometry of Banach spaces in metric fixed point theory, PhD thesis, Va<strong>le</strong>ncia, 2003.<br />

[33] S. Prus, Banach spaces which are uniformly noncreasy, in: V. Lakshmikantham (Ed.), Proc. 2nd World Congress of<br />

Nonlinear Analysts, Nonlinear Anal. 30 (1997) 2317–2324.<br />

[34] B. Sims, Ultra-Techniques in Banach Space Theory, Queen’s Papers in Pure Appl. Math., vol. 60, Queen’s University<br />

Kingston, Ontario, 1982.<br />

[35] B. Sims, M.A. Smyth, On some Banach space properties sufficient for weak normal structure and their permanence<br />

properties, Trans. Amer. Math. Soc. 351 (1999) 497–513.<br />

[36] J. Stern, Some applications of mo<strong>de</strong>l theory in Banach space theory, Ann. Math. Logic 9 (1976) 49–122.<br />

[37] K.K. Tan, H.K. Xu, On fixed point theorems of nonexpansive mappings in pro<strong>du</strong>ct spaces, Proc. Amer. Math.<br />

Soc. 113 (1991) 983–989.<br />

[38] A. Wi´snicki, Towards the fixed point property for superref<strong>le</strong>xive spaces, Bull. Austral. Math. Soc. 64 (2001) 435–<br />

444.<br />

[39] A. Wi´snicki, Pro<strong>du</strong>cts of uniformly noncreasy spaces, Proc. Amer. Math. Soc. 130 (2002) 3295–3299.


Journal of Functional Analysis 236 (2006) 457–489<br />

Brown mea<strong>sur</strong>es of sets of commuting<br />

operators in a type II1 factor<br />

Hanne Schultz 1<br />

www.elsevier.com/locate/jfa<br />

Department of Mathematics and Computer Science, University of Southern Denmark, Denmark<br />

Received 14 November 2005; accepted 9 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 18 April 2006<br />

Communicated by G. Pisier<br />

Abstract<br />

Using the spectral subspaces obtained in [U. Haagerup, H. Schultz, Invariant subspaces of operators in a<br />

general II1-factor, preprint, 2005], Brown’s results (cf. [L.G. Brown, Lidskii’s theorem in the type II case,<br />

in: H. Araki, E. Effros (Eds.), Geometric Methods in Operator Algebras, Kyoto, 1983, in: Pitman Res.<br />

Notes Math. Ser., vol. 123, Longman Sci. Tech., 1986, pp. 1–35]) on the Brown mea<strong>sur</strong>e of an operator<br />

in a type II1 factor (M,τ) are generalized to finite sets of commuting operators in M. Itisshownthat<br />

whenever T1,...,Tn ∈ M are mutually commuting operators, there exists one and only one compactly<br />

supported Borel probability mea<strong>sur</strong>e μT1,...,Tn on B(Cn ) such that for all α1,...,αn ∈ C,<br />

τ log |α1T1 +···+αnTn − 1| <br />

=<br />

C n<br />

log |α1z1 +···+αnzn − 1| dμT1,...,Tn (z1,...,zn).<br />

Moreover, for every polynomial q in n commuting variab<strong>le</strong>s, μ q(T1,...,Tn) is the push-forward mea<strong>sur</strong>e of<br />

μT1,...,Tn via the map q : Cn → C.<br />

In addition it is shown that, as in [U. Haagerup, H. Schultz, Invariant subspaces of operators in a general<br />

II1-factor, preprint, 2005], for every Borel set B ⊆ C n there is a maximal closed T1-,...,Tn-invariant subspace<br />

K affiliated with M, such that μT1|K,...,Tn|K is concentrated on B. Moreover, τ(P K) = μT1,...,Tn (B).<br />

This generalizes the main result from [U. Haagerup, H. Schultz, Invariant subspaces of operators in a general<br />

II1-factor, preprint, 2005] to n-tup<strong>le</strong>s of commuting operators in M.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

E-mail address: schultz@imada.s<strong>du</strong>.dk.<br />

1 As a stu<strong>de</strong>nt of the PhD-school OP-ALG-TOP-GEO the author is partially supported by the Danish Research Training<br />

Council. Partially supported by The Danish National Research Foundation.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.003


458 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Keywords: Brown mea<strong>sur</strong>e; Invariant subspaces; II1-factor; Unboun<strong>de</strong>d i<strong>de</strong>mpotents<br />

1. Intro<strong>du</strong>ction<br />

In [4] Brown showed that for every operator T in a type II1 factor (M,τ) there is one and<br />

only one compactly supported Borel probability mea<strong>sur</strong>e μT on C, theBrown mea<strong>sur</strong>e of T ,<br />

such that for all λ ∈ C,<br />

τ log |T − λ1| <br />

= log |z − λ| dμT (z).<br />

C<br />

In [7] it was shown that given T ∈ M and B ∈ B(C), there is a maximal closed T -invariant<br />

projection P = PT (B) ∈ M, such that the Brown mea<strong>sur</strong>e of PTP (consi<strong>de</strong>red as an e<strong>le</strong>ment<br />

of P MP ) is concentrated on B. Moreover, PT (B) is hyper-invariant for T ,<br />

τ PT (B) = μT (B), (1.1)<br />

and the Brown mea<strong>sur</strong>e of P ⊥ TP ⊥ (consi<strong>de</strong>red as an e<strong>le</strong>ment of P ⊥ MP ⊥ ) is concentrated<br />

on B c .<br />

In particu<strong>la</strong>r, if S,T ∈ M are commuting operators and A,B ∈ B(C), then PS(A) ∧ PT (B) is<br />

S- and T -invariant. Thus, it is tempting to <strong>de</strong>fine a Brown mea<strong>sur</strong>e for the pair (S, T ), μS,T ,by<br />

μS,T (A × B) = τ PS(A) ∧ PT (B) , A,B ∈ B(C). (1.2)<br />

In or<strong>de</strong>r to show that μS,T extends (uniquely) to a Borel probability mea<strong>sur</strong>e on C2 , we intro<strong>du</strong>ce<br />

the notion of an i<strong>de</strong>mpotent valued mea<strong>sur</strong>e (cf. Section 3), and for T ∈ M and B ∈ B(C) we<br />

<strong>de</strong>fine an unboun<strong>de</strong>d i<strong>de</strong>mpotent affiliated with M, eT (B), byD(eT (B)) = KT (B) + KT (Bc )<br />

and<br />

<br />

ξ, ξ ∈ KT (B),<br />

eT (B)ξ =<br />

0, ξ ∈ KT (Bc ),<br />

where KT (B) is the range of PT (B) (cf. Section 3). We prove that the map B ↦→ eT (B) is an<br />

i<strong>de</strong>mpotent valued mea<strong>sur</strong>e and this enab<strong>le</strong>s us to show that μS,T given by (1.2) has an extension<br />

to a mea<strong>sur</strong>e on C2 (cf. Section 4). As in [7] we also prove the existence of spectral subspaces<br />

for a set of commuting operators in M. In the case of two commuting operators S,T ∈ M, we<br />

show that for B ∈ B(C2 ) there is a maximal S- and T -invariant projection P ∈ P(M), such<br />

that μS|P(H),T |P(H)<br />

(1.3)<br />

(computed re<strong>la</strong>tive to P MP ) is concentrated on B. In the case where B =<br />

B1 × B2, B1,B2 ∈ B(C), P is simply the intersection PS(B1) ∧ PT (B2) (cf. Section 5).<br />

Finally, in Section 6 we show that μS,T has a characterization simi<strong>la</strong>r to the one given by<br />

Brown in [4]. That is, we show that μS,T is the unique compactly supported Borel probability<br />

mea<strong>sur</strong>e on B(C2 ) such that for all α, β ∈ C,<br />

τ log |αS + βT − 1| <br />

= log |αz + βw − 1| dμS,T (z, w),<br />

C 2


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 459<br />

and there is a simi<strong>la</strong>r characterization of the Brown mea<strong>sur</strong>e of an arbitrary finite set of commuting<br />

operators. In particu<strong>la</strong>r, μαS+βT is the push-forward mea<strong>sur</strong>e of μS,T via the map<br />

(z, w) ↦→ αz + βw. In fact we can show that for every polynomial q in two commuting variab<strong>le</strong>s,<br />

μq(S,T ) is the push-forward mea<strong>sur</strong>e of μS,T via the map q : C2 → C. All this can (and<br />

will) be generalized to arbitrary finite sets of mutually commuting operators.<br />

Section 2 is <strong>de</strong>voted to a summary of [10] in which we recall the <strong>de</strong>finition of the mea<strong>sur</strong>e<br />

topology on M and how the clo<strong>sur</strong>e of M with respect to the mea<strong>sur</strong>e topology may be i<strong>de</strong>ntified<br />

with the set of closed, <strong>de</strong>nsely <strong>de</strong>fined operators affiliated with M. Afterwards we focus our<br />

attention on the set of unboun<strong>de</strong>d i<strong>de</strong>mpotents affiliated with M, I( ˜M). We prove that these<br />

are in one-to-one correspon<strong>de</strong>nce with pairs of projections P,Q ∈ M with P ∧ Q = 0 and<br />

P ∨ Q = 1. More precisely, when E is a closed, <strong>de</strong>nsely <strong>de</strong>fined, unboun<strong>de</strong>d operator affiliated<br />

with M with E · E = E, then P , the range projection of E, and Q, the projection onto the kernel<br />

of E, satisfy that P ∧ Q = 0 and P ∨ Q = 1. Moreover, D(E) = P(H) + Q(H), and E is given<br />

by<br />

<br />

ξ, ξ ∈ P(H),<br />

Eξ =<br />

(1.4)<br />

0, ξ ∈ Q(H).<br />

Vice versa, when P,Q∈ M with P ∧ Q = 0 and P ∨ Q = 1, then (1.4) <strong>de</strong>fines a closed, <strong>de</strong>nsely<br />

<strong>de</strong>fined, unboun<strong>de</strong>d operator affiliated with M with E · E = E. In particu<strong>la</strong>r, (1.3) <strong>de</strong>fines a<br />

closed, <strong>de</strong>nsely <strong>de</strong>fined i<strong>de</strong>mpotent affiliated with M. In Section 2 it is also shown that I( ˜M) is<br />

stab<strong>le</strong> with respect to addition of countably many i<strong>de</strong>mpotents (En) ∞ n=1 with EnEm = EmEn = 0,<br />

n = m. These are results which are nee<strong>de</strong>d in the succeeding sections.<br />

2. I<strong>de</strong>mpotents in ˜M<br />

We begin this section with a summary of E. Nelson’s “Notes on non-commutative integration”<br />

[10]. We consi<strong>de</strong>r a finite von Neumann algebra M represented on a Hilbert space H and<br />

we fix a faithful, normal, tracial state τ on M.We<strong>le</strong>tP(M) <strong>de</strong>note the set of projections in M.<br />

Nelson <strong>de</strong>fines the mea<strong>sur</strong>e topology on M as follows. For ε, δ > 0, <strong>le</strong>t<br />

N(ε, δ) = T ∈ M |∃P ∈ P(M): TP ε, τ P ⊥ δ .<br />

The mea<strong>sur</strong>e topology on M is then the trans<strong>la</strong>tion invariant topology on M for which the<br />

N(ε, δ)’s form a fundamental system of neighborhoods of 0. ˜M is the comp<strong>le</strong>tion of M with<br />

respect to the mea<strong>sur</strong>e topology.<br />

Simi<strong>la</strong>rly, Nelson <strong>de</strong>fines ˜H to be the comp<strong>le</strong>tion of H with respect to the trans<strong>la</strong>tion invariant<br />

topology on H for which the sets<br />

O(ε, δ) = ξ ∈ H |∃P ∈ P(M): Pξ ε, τ P ⊥ δ <br />

form a fundamental system of neighborhoods of 0. According to [10, Theorem 2], the natural<br />

mappings M → ˜M and H → ˜H are both injections.<br />

Theorem 2.1. [10, Theorem 1] The mappings<br />

M → M : T ↦→ T ∗ , M × M → M : (S, T ) ↦→ S + T, M × M → M : (S, T ) ↦→ S · T,<br />

H × H → H : (ξ, η) ↦→ ξ + η, M × H → H : (T , ξ) ↦→ Tξ


460 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

all have unique continuous extensions as mappings ˜M → ˜M, ˜M × ˜M → ˜M, ˜M × ˜M → ˜M,<br />

˜H × ˜H → ˜H and ˜M × ˜H → ˜H, respectively. In particu<strong>la</strong>r, ˜H is a comp<strong>le</strong>x vector space and ˜M<br />

is a comp<strong>le</strong>x ∗-algebra with a continuous representation on ˜H.<br />

For x ∈ ˜M <strong>de</strong>fine<br />

and <strong>de</strong>fine Mx : D(Mx) → H by<br />

D(Mx) ={ξ ∈ H | xξ ∈ H} (2.1)<br />

Mxξ = xξ, ξ ∈ D(Mx). (2.2)<br />

Recall that a (not necessarily boun<strong>de</strong>d or everywhere <strong>de</strong>fined) operator A on H is said to be<br />

affiliated with M if AU = UA for every unitary U ∈ M ′ .<br />

Theorem 2.2. [10, Theorem 4] For every x ∈ ˜M, Mx is a closed, <strong>de</strong>nsely <strong>de</strong>fined operator<br />

affiliated with M, and<br />

Moreover, for x,y ∈ ˜M,<br />

M ∗ x<br />

where A <strong>de</strong>notes the clo<strong>sur</strong>e of a closab<strong>le</strong> operator A.<br />

= Mx ∗. (2.3)<br />

Mx+y = Mx + My, (2.4)<br />

Mx·y = Mx · My, (2.5)<br />

A closed, <strong>de</strong>nsely <strong>de</strong>fined operator A on H has a po<strong>la</strong>r <strong>de</strong>composition A = V |A|, and if A is<br />

affiliated with M, then V ∈ M and all the spectral projections (E|A|([0,t[))t>0 belong to M.<br />

Put<br />

An = V<br />

n<br />

0<br />

t dE|A|(t).<br />

Assuming that A is affiliated with M, we get that (An) ∞ n=1 is a Cauchy sequence with respect to<br />

the mea<strong>sur</strong>e topology. In<strong>de</strong>ed,<br />

<br />

(An+k − An) · E|A| [0,n[ = 0,<br />

and τ(E|A|([n, ∞[)) → 0asn →∞. Hence, there exists a ∈ ˜M such that An → a in mea<strong>sur</strong>e,<br />

and according to [10, Theorem 3], A = Ma. It follows that every closed, <strong>de</strong>nsely <strong>de</strong>fined operator<br />

A affiliated with M is of the form A = Ma for some a ∈ ˜M, and this a is uniquely <strong>de</strong>termined.<br />

In<strong>de</strong>ed, if Ma = Mb, then a and b agree on D(Ma) which is <strong>de</strong>nse in H with respect to the norm<br />

topology and hence <strong>de</strong>nse in ˜H with respect to the mea<strong>sur</strong>e topology. Since the representation of<br />

˜M on H is continuous, it follows that a and b agree on all of ˜H. By the same argument, if S and


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 461<br />

T are closed, <strong>de</strong>nsely <strong>de</strong>fined operators affiliated with M which agree on a <strong>de</strong>nse subset of H,<br />

then S = T .<br />

In summary, ˜M is the comp<strong>le</strong>tion of M with respect to the mea<strong>sur</strong>e topology but it may also<br />

be viewed as the set of closed, <strong>de</strong>nsely <strong>de</strong>fined operators affiliated with M. In particu<strong>la</strong>r, if S and<br />

T belong to the <strong>la</strong>tter, then Theorem 2.2 tells us that S ∗ , S + T and S · T are also closed, <strong>de</strong>nsely<br />

<strong>de</strong>fined and affiliated with M.<br />

Notation 2.3. In what follows we shall i<strong>de</strong>ntify ˜M with the set of closed, <strong>de</strong>nsely <strong>de</strong>fined operators<br />

affiliated with M, but whenever necessary, we will specify which one of the two pictures<br />

mentioned above we are using. In general, lower case <strong>le</strong>tters will represent e<strong>le</strong>ments of the comp<strong>le</strong>tion<br />

of M with respect to the mea<strong>sur</strong>e topology, whereas upper case <strong>le</strong>tters will represent<br />

closed, <strong>de</strong>nsely <strong>de</strong>fined operators affiliated with M. Forx ∈ ˜M we will <strong>de</strong>note by ker(x),<br />

range(x) and supp(x) the kernel of Mx, the range of Mx and the support projection of Mx,<br />

respectively.<br />

In the rest of this section we will study the set of i<strong>de</strong>mpotent e<strong>le</strong>ments in ˜M,<br />

I( ˜M) ={e ∈ ˜M | e · e = e}. (2.6)<br />

Alternatively, if ˜M is viewed as the set of closed, <strong>de</strong>nsely <strong>de</strong>fined operators affiliated with M,<br />

then I( ˜M) is the subset of operators E fulfilling that E · E = E.<br />

The following proposition shows that I( ˜M) is in one-to-one correspon<strong>de</strong>nce with the set of<br />

pairs of projections P,Q∈ M such that P ∧ Q = 0 and P ∨ Q = 1.<br />

Proposition 2.4. Let E ∈ I( ˜M). Then the range of E, range(E), and the kernel of E, ker(E)<br />

(which is the range of 1 − E), are closed subspaces of H, with<br />

and<br />

range(E) ∩ ker(E) ={0} (2.7)<br />

range(E) + ker(E) = H. (2.8)<br />

Moreover, D(E) = range(E)+ker(E). Conversely, if P , Q ∈ M are projections with P ∧Q = 0<br />

and P ∨ Q = 1, then there is a unique i<strong>de</strong>mpotent E ∈ ˜M with D(E) = P(H) + Q(H),<br />

range(E) = P(H) and ker(E) = Q(H), and E is <strong>de</strong>termined by<br />

<br />

ξ, ξ ∈ P(H),<br />

Eξ =<br />

(2.9)<br />

0, ξ ∈ Q(H).<br />

Proof. Write E = Me for a (uniquely <strong>de</strong>termined) e<strong>le</strong>ment e ∈ ˜M with e · e = e and recall<br />

from (2.1) that<br />

D(E) ={ξ ∈ H | eξ ∈ H}.<br />

Let E · E <strong>de</strong>note the <strong>de</strong>nsely <strong>de</strong>fined operator on H obtained by composing E with itself. Then<br />

D(E · E) = ξ ∈ D(E) | Eξ ∈ D(E) = ξ ∈ D(E) | eEξ ∈ H = ξ ∈ D(E) | e 2 ξ ∈ H .


462 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

But e 2 = e as everywhere <strong>de</strong>fined operators on ˜H. Hence, for ξ ∈ D(E),<br />

e 2 ξ = eξ = Eξ,<br />

and it follows that D(E · E) = D(E) and that E · E = E (without taking clo<strong>sur</strong>e). In particu<strong>la</strong>r,<br />

range(E) ⊆ D(E). Moreover, since E · E = E,<br />

range(E) = ξ ∈ D(E) | Eξ = ξ = ker(1 − E).<br />

According to [8, Exercise 2.8.45], the kernel of a closed operator is closed. Hence, ker(E) and<br />

range(E) = ker(1 − E) are closed. Moreover,<br />

range(E) ∩ ker(1 − E) = ker(1 − E) ∩ ker(E) ={0}.<br />

C<strong>le</strong>arly, range(E) + ker(E) ⊆ D(E), and since<br />

the converse inclusion also holds. That is,<br />

ξ = Eξ + (1 − E)ξ, ξ ∈ D(E),<br />

D(E) = range(E) + ker(E).<br />

Let P and Q <strong>de</strong>note the projections onto range(E) and ker(E), respectively. Then by the above,<br />

P ∧ Q = 0 and P ∨ Q = 1, and E is the operator on D(E) = P(H) + Q(H) <strong>de</strong>termined by (2.9).<br />

In particu<strong>la</strong>r, E is uniquely <strong>de</strong>termined by its range and its kernel.<br />

Conversely, assume that P and Q are projections in M, such that P ∧ Q = 0 and P ∨ Q = 1,<br />

and <strong>le</strong>t E be the operator on D(E) = P(H)+Q(H) <strong>de</strong>termined by (2.9). Then c<strong>le</strong>arly, E ·E = E<br />

(without taking clo<strong>sur</strong>e), range(E) = P(H), and ker(E) = Q(H). Moreover, the graph of E is<br />

given by<br />

G(E) = (ξ + η,ξ) | ξ ∈ P(H), η ∈ Q(H) <br />

= (u, v) ∈ H × H | u − v ∈ Q(H), v ∈ P(H) ,<br />

which is c<strong>le</strong>arly a closed subspace of H × H. Hence, E ∈ I( ˜M). ✷<br />

Definition 2.5. We <strong>de</strong>fine tr : I( ˜M) →[0, 1] by<br />

where supp(e) ∈ P(M) <strong>de</strong>notes the support projection of Me.<br />

tr(e) = τ supp(e) , e∈ I( ˜M), (2.10)<br />

Remark 2.6. For every x ∈ ˜M, supp(x) ∼ Prange(x) so one also has that for e ∈ I( ˜M),<br />

tr(e) = τ(Prange(e)). (2.11)<br />

Throughout the paper we will, without further mentioning, make use of this i<strong>de</strong>ntity.


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 463<br />

Proposition 2.7. Let e1,...,en ∈ I( ˜M) (seen as everywhere <strong>de</strong>fined operators on ˜H) with<br />

eiej = 0 when i = j. Then e1 +···+en ∈ I( ˜M), and<br />

(a) ker(e1 +···+en) = n i=1 ker(ei);<br />

(b) supp(e1 +···+en) = n i=1 supp(ei);<br />

(c) range(e1 +···+en) = n i=1 range(ei);<br />

(d) tr(e1 +···+en) = n i=1 tr(ei).<br />

Proof. It suffices to consi<strong>de</strong>r the case n = 2. The general case follows by in<strong>du</strong>ction over n ∈ N.<br />

If e1,e2 ∈ I( ˜M) with e1e2 = e2e1 = 0, then obviously, e1 + e2 ∈ I( ˜M).<br />

(a) C<strong>le</strong>arly, ker(e1) ∩ ker(e2) ⊆ ker(e1 + e2). On the other hand, if ξ ∈ ker(e1 + e2), then<br />

eiξ = ei(e1 + e2)ξ = 0(i = 1, 2), whence ker(e1 + e2) ⊆ ker(e1) ∩ ker(e2).<br />

(b) Since supp(e1 + e2) is the projection onto [ker(e1 + e2)] ⊥ , (b) follows from (a).<br />

(c) For a closed, <strong>de</strong>nsely <strong>de</strong>fined operator S affiliated with M, range(S) = ker(S∗ ) ⊥ .Using<br />

that every i<strong>de</strong>mpotent has closed range (cf. Proposition 2.4) and applying (a) to e∗ 1 +···+e∗ n ,we<br />

find that<br />

range(e1 +···+en) = range(e1 +···+en) = ker e ∗ 1 +···+e∗ ⊥ n<br />

=<br />

n<br />

i=1<br />

ker e ∗⊥ i =<br />

n<br />

range(ei).<br />

(d) For i = 1, 2 put Pi = Prange(ei). Since e1e2 = e2e1 = 0, for ξ ∈ P1(H) ∩ P2(H) we have<br />

that<br />

i=1<br />

ξ = e1ξ = e2e1ξ = 0.<br />

Then by Kap<strong>la</strong>nsky’s formu<strong>la</strong> (cf. [8, Theorem 6.1.7]),<br />

so that<br />

P1 ∨ P2 − P1 ∼ P2 − P1 ∧ P2 = P2,<br />

τ(P1 ∨ P2) = τ(P1) + τ(P2).<br />

According to (c), P1 ∨ P2 is the projection onto range(e1 + e2), and then by Remark 2.6,<br />

tr(e1 + e2) = tr(e1) + tr(e2). ✷<br />

Proposition 2.8. Let (en) ∞ n=1 be a sequence of i<strong>de</strong>mpotents in ˜M with enem = emen = 0 when<br />

n = m. Then there is an i<strong>de</strong>mpotent in ˜M, which we <strong>de</strong>note by ∞ n=1 en, such that N n=1 en →<br />

∞n=1 en in mea<strong>sur</strong>e as N →∞. Moreover, supp( N n=1 en) ↗ supp( ∞ n=1 en), whence<br />

<br />

∞<br />

supp<br />

<br />

∞<br />

= supp(en), (2.12)<br />

n=1<br />

en<br />

n=1


464 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

and<br />

Also,<br />

Proof. Let<br />

range<br />

<br />

∞<br />

tr<br />

n=1<br />

∞<br />

n=1<br />

en<br />

en<br />

fn =<br />

k=1<br />

<br />

∞<br />

= tr(en). (2.13)<br />

n=1<br />

<br />

∞<br />

= range(en). (2.14)<br />

n=1<br />

n<br />

ek, n∈N. (2.15)<br />

Then, according to Proposition 2.7, supp(fn) = n k=1 supp(ek), and tr(fn) = n k=1 tr(ek). We<br />

prove that (fn) ∞ n=1 is a Cauchy sequence in ˜M.Forn, k ∈ N,<strong>le</strong>t<br />

Then<br />

so for every ε>0,<br />

Pn,k = supp(fn+k − fn) =<br />

(fn+k − fn)P ⊥ n,k<br />

k<br />

supp(en+l). (2.16)<br />

l=1<br />

fn+k − fn ∈ N ε, τ(Pn,k) <br />

= N ε,<br />

= 0, (2.17)<br />

k<br />

<br />

tr(en+l) . (2.18)<br />

Now, n k=1 tr(ek) = tr(fn) 1, so for arbitrary δ>0 there is an n0 ∈ N such that<br />

l=1<br />

∞<br />

tr(ek) δ. (2.19)<br />

k=n0<br />

It follows from (2.18) and (2.19) that when n n0 and k 1, then fn+k − fn ∈ N(ε, δ). Thus,<br />

(fn) ∞ n=1 is a Cauchy sequence in ˜M. Put<br />

For all k,n ∈ N, fn+kfn = fnfn+k = fn, and hence<br />

so that e · e = e.<br />

e = lim<br />

n→∞ fn ∈ ˜M. (2.20)<br />

efn = fne = fn, (2.21)


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 465<br />

Let Pn = supp(fn) = n k=1 supp(ek). Then Pn ↗ P := ∞ k=1 supp(ek) as n →∞, and<br />

∞<br />

tr(ek) = lim<br />

k=1<br />

n<br />

n→∞<br />

k=1<br />

tr(ek) = lim<br />

n→∞ tr(fn) = lim<br />

n→∞ τ(Pn) = tr(P ).<br />

It follows from (2.21) that for every n ∈ N, Pn supp(e). Hence P supp(e). On the other<br />

hand, for every n ∈ N, fn(1 − Pn) = 0, so<br />

<br />

e(1 − P)= lim fn(1 − Pn)<br />

n→∞<br />

= 0 (2.22)<br />

(the limit refers to the mea<strong>sur</strong>e topology). Thus, supp(e) P , and we have shown that supp(e) =<br />

P = ∞ k=1 supp(ek) and that (2.13) holds.<br />

In or<strong>de</strong>r to prove (2.14), <strong>le</strong>t ξ ∈ range(em). Then emξ = ξ and<br />

Thus,<br />

range<br />

∞<br />

en<br />

n=1<br />

∞<br />

n=1<br />

<br />

ξ = emξ = ξ.<br />

en<br />

<br />

∞<br />

⊇ range(en). (2.23)<br />

On the other hand, we know that range( ∞ n=1 en) ⊆ H, and since range( N n=1 en) ⊆<br />

∞n=1 range(en) for all N ∈ N,wealsohavethat<br />

range<br />

∞<br />

n=1<br />

en<br />

n=1<br />

<br />

∞<br />

⊆ range(en)<br />

where ∼ <strong>de</strong>notes clo<strong>sur</strong>e with respect to the mea<strong>sur</strong>e topology. Intersecting by H on both si<strong>de</strong>s<br />

of the inclusion, we get that<br />

This proves (2.14). ✷<br />

range<br />

∞<br />

n=1<br />

en<br />

n=1<br />

∼<br />

,<br />

<br />

∞<br />

⊆ range(en). (2.24)<br />

We shall make use of the following theorem from [2]. For a published proof of it, we refer the<br />

rea<strong>de</strong>r to [1].<br />

n=1


466 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Theorem 2.9. [2] Let E and F be (not necessarily closed) subspaces of H which are affiliated<br />

with M. 2 Then E ∩ F is affiliated with A, and<br />

E ∩ F = E ∩ F. (2.25)<br />

Lemma 2.10. Consi<strong>de</strong>r i<strong>de</strong>mpotents e, f ∈ ˜M. Let P = Prange(e), Q = Prange(1−e), R = Prange(f )<br />

and S = Prange(1−f). Then ef = feif and only if<br />

Proof. C<strong>le</strong>arly,<br />

(P ∧ R) ∨ (P ∧ S) ∨ (Q ∧ R) ∨ (Q ∧ S) = 1. (2.26)<br />

1 = ef + e(1 − f)+ (1 − e)f + (1 − e)(1 − f). (2.27)<br />

Suppose that ef = fe, and <strong>le</strong>t g1 = ef , g2 = e(1 − f), g3 = (1 − e)f and g4 = (1 − e)(1 − f).<br />

Then g1,...,g4 are i<strong>de</strong>mpotents with support projections P1, P2, P3 and P4, respectively, such<br />

that<br />

<br />

4<br />

<br />

(H) = H. (2.28)<br />

Pi<br />

i=1<br />

Moreover, P1 P ∧ R, P2 P ∧ S, P3 Q ∧ R and P4 Q ∧ S. This shows that (2.26) holds.<br />

On the other hand, assume that (2.26) holds. According to (2.26) and Theorem 2.9,<br />

H0 := D(ef ) ∩ D(f e) ∩ (P ∧ R)(H) + (P ∧ S)(H) + (Q ∧ R)(H) + (Q ∧ S)(H) <br />

is <strong>de</strong>nse in H so it suffices to prove that ef and fe agree on H0. To see this, <strong>le</strong>t ξ ∈ D(ef ) ∩<br />

D(f e) ∩ (P ∧ R)(H). Then<br />

ef ξ = eξ = ξ = fξ = feξ, (2.29)<br />

and simi<strong>la</strong>rly, when ξ ∈ D(ef ) ∩ D(f e) ∩ (P ∧ S)(H), ξ ∈ D(ef ) ∩ D(f e) ∩ (Q ∧ R)(H) or<br />

ξ ∈ D(ef ) ∩ D(f e) ∩ (Q ∧ S)(H). Thus, ef agrees with feon H0. ✷<br />

3. An i<strong>de</strong>mpotent valued mea<strong>sur</strong>e associated with T ∈ M<br />

As in the previous section, consi<strong>de</strong>r a finite von Neumann algebra M with a faithful, normal,<br />

tracial state τ . Inspired by the notion of a spectral mea<strong>sur</strong>e we make the following <strong>de</strong>finition.<br />

Definition 3.1. Let (X, F) <strong>de</strong>note a mea<strong>sur</strong>ab<strong>le</strong> space. An i<strong>de</strong>mpotent valued mea<strong>sur</strong>e on (X, F)<br />

(with values in ˜M) isamape from F into I( ˜M) such that:<br />

(i) e(X) = 1,<br />

(ii) e(F1)e(F2) = e(F2)e(F1) = 0 when F1,F2 ∈ F with F1 ∩ F2 =∅,<br />

2 A subspace E of H is said to be affiliated with M if for all T ∈ M ′ , T(E)⊆ E. Note that if E is affiliated with M,<br />

then the projection onto E belongs to M,andifE is closed, then this is a necessary and sufficient condition for E to be<br />

affiliated with M.


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 467<br />

(iii) when (Fn) ∞ n=1 is a sequence of mutually disjoint sets from F, then N n=1 e(Fn) converges<br />

in mea<strong>sur</strong>e as N →∞to e( ∞ n=1 Fn), i.e.<br />

<br />

∞<br />

<br />

∞<br />

e = e(Fn).<br />

n=1<br />

Fn<br />

Note that because of (ii) and Proposition 2.8, the limit in (iii) actually exists.<br />

From now on we will assume that M is in fact a type II1 factor. Recall from [7] that for<br />

T ∈ M and B ⊆ C a Borel set there is a maximal T -invariant projection P = PT (B) ∈ M, such<br />

that the Brown mea<strong>sur</strong>e of PTP (consi<strong>de</strong>red as an e<strong>le</strong>ment of P MP ) is concentrated on B.<br />

Moreover, PT (B) is hyper-invariant for T ,<br />

n=1<br />

τ PT (B) = μT (B), (3.1)<br />

and the Brown mea<strong>sur</strong>e of P ⊥ TP ⊥ (consi<strong>de</strong>red as an e<strong>le</strong>ment of P ⊥ MP ⊥ ) is concentrated<br />

on B c .We<strong>le</strong>tKT (B) <strong>de</strong>note the range of PT (B). Then the aim of this section is to prove:<br />

Theorem 3.2. Let T ∈ M, and for B ∈ B(C), <strong>le</strong>teT (B) with D(eT (B)) = KT (B) + KT (B c ) be<br />

given by<br />

<br />

ξ, ξ ∈ KT (B),<br />

eT (B)ξ =<br />

0, ξ ∈ KT (Bc ).<br />

Then eT (B) ∈ I( ˜M), and B ↦→ eT (B) is an i<strong>de</strong>mpotent valued mea<strong>sur</strong>e.<br />

The proof of this theorem uses various results which we state and prove below. The first one<br />

of these is a <strong>le</strong>mma which we proved in [7], but for the sake of comp<strong>le</strong>teness we give the proof<br />

here as well.<br />

Lemma 3.3. Let T ∈ M, and <strong>le</strong>t P ∈ M be a non-zero, T -invariant projection. Then for every<br />

B ∈ B(C),<br />

(3.2)<br />

KT |P(H) (B) = KT (B) ∩ P(H), (3.3)<br />

where T |P(H) is consi<strong>de</strong>red as an e<strong>le</strong>ment of the type II1 factor P MP .<br />

Proof. Let Q ∈ P MP <strong>de</strong>note the projection onto KT |P(H) (B), and <strong>le</strong>t R = PT (B) ∧ P . We will<br />

prove that Q R and R Q.<br />

C<strong>le</strong>arly, Q P . In or<strong>de</strong>r to see that Q PT (B), recall that PT (B) is maximal with respect<br />

to the properties:<br />

(i) PT (B)T PT (B) = TPT (B);<br />

(ii) μPT (B)T PT (B) (computed re<strong>la</strong>tive to PT (B)MPT (B)) is concentrated on B.<br />

Since<br />

QT Q = QT P Q = TPQ= TQ, (3.4)


468 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

and μQT Q = μQT P Q (computed re<strong>la</strong>tive to QMQ) is concentrated on B, we get that Q <br />

PT (B), and hence Q R.<br />

Simi<strong>la</strong>rly, to prove that R Q, we must show that:<br />

(i ′ ) RT PR = TPR,i.e.RT R = TR;<br />

(ii ′ ) μRT PR = μRT R (computed re<strong>la</strong>tive to RMR) is concentrated on B.<br />

Note that if PT (B) = 0, then R Q, so we may assume that PT (B) = 0. (i ′ ) holds, because<br />

R(H) = P(H) ∩ PT (B)(H) is T -invariant when P(H) and PT (B)(H) are T -invariant. In or<strong>de</strong>r<br />

to prove (ii ′ ), at first note that R(H) is TPT (B)-invariant. Hence<br />

⊥<br />

μTPT (B) = τ1(R) · μRT R + τ1 R · μR⊥TR⊥, (3.5)<br />

where<br />

It follows that<br />

τ1 =<br />

1<br />

τ(PT (B)) · τ|PT (B)MPT (B).<br />

c<br />

τ1(R) · μRT R B c<br />

μTPT (B) B = 0, (3.6)<br />

and thus, if R = 0, then μRT R(B c ) = 0, and (ii ′ ) holds. If R = 0, then R Q is trivially fulfil<strong>le</strong>d.<br />

✷<br />

Proposition 3.4. For every Borel set B ⊆ C,<br />

KT (B) = KT ∗<br />

c<br />

B ∗⊥, (3.7)<br />

where A ∗ := {z | z ∈ A} for A ⊆ C. Moreover, for all Borel sets A,B ⊆ C,<br />

and<br />

KT (A) ∩ KT (B) = KT (A ∩ B), (3.8)<br />

KT (A ∪ B) = KT (A) + KT (B). (3.9)<br />

Proof. Let B ∈ B(C) and <strong>le</strong>t P = PT (B). Then P ⊥ is T ∗-invariant, and<br />

∗<br />

μP ⊥T ∗P ⊥ B = μ (P ⊥T ∗P ⊥ ) ∗(B) = μP ⊥TP⊥(B) = 0 (3.10)<br />

(recall that μP ⊥TP⊥ is concentrated on Bc ). Thus, μP ⊥T ∗P ⊥ is concentrated on C \ B∗ , and<br />

maximality of PT ∗(C \ B∗ ) implies that<br />

PT (B) ⊥ = P ⊥ PT ∗<br />

∗<br />

C \ B . (3.11)<br />

Since


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 469<br />

τ PT ∗<br />

∗<br />

C \ B = μT ∗<br />

∗<br />

C \ B = 1 − μT ∗<br />

∗<br />

B = 1 − μT (B) = τ PT (B) ⊥ ,<br />

we get from (3.11) that PT (B) ⊥ = PT ∗(C \ B∗ ).<br />

Next, <strong>le</strong>t A,B ∈ B(C). By maximality of PT (A) and PT (B), PT (A ∩ B) PT (A) ∧ PT (B),<br />

so ⊇ holds in (3.8). We <strong>le</strong>t K := KT (A) ∩ KT (B), and we <strong>le</strong>t Q <strong>de</strong>note the projection onto K.<br />

Then, according to Lemma 3.3,<br />

K = KT |K T (A) (B) = KT |K T (B) (A),<br />

proving that μQT Q is concentrated on A and on B and therefore on A ∩ B. Consequently, Q <br />

PT (A ∩ B),so⊆ also holds in (3.8).<br />

Finally, we infer from (3.7) and (3.8) that<br />

c c<br />

KT (A ∪ B) = KT A ∩ B c = KT ∗<br />

c c<br />

A ∩ B ∗⊥ = KT ∗<br />

c<br />

A ∗ c<br />

∩ B ∗⊥ = KT ∗<br />

c<br />

A ∗ ∩ KT ∗<br />

c<br />

B ∗⊥ = KT ∗<br />

<br />

Ac ∗⊥ + KT ∗<br />

<br />

Bc ∗⊥ = KT (A) + KT (B). ✷<br />

It follows from Propositions 3.4 and 2.4 that for B ∈ B(C), eT (B) given by (3.2) belongs to<br />

I( ˜M), as stated in Theorem 3.2.<br />

Lemma 3.5. Let (xn) ∞ n=1 be a sequence in ˜M, and suppose τ(supp(xn)) → 0 as n →∞. Then<br />

xn → 0 in the mea<strong>sur</strong>e topology.<br />

Proof. This is standard. ✷<br />

If S ∈ M commutes with T ∈ M, then for every B ∈ B(C), KT (B) and KT (B c ) are<br />

S-invariant, and therefore S commutes with eT (B) as well. We prove that, as a consequence<br />

of this, [eS(A), eT (B)]=0 for every A ∈ B(C).<br />

Lemma 3.6. Let T ∈ M, and <strong>le</strong>t e ∈ I( ˜M) with [e,T ]=0. Then for every B ∈ B(C),<br />

[e,eT (B)]=0. In particu<strong>la</strong>r, if S ∈ M commutes with T , then [eS(·), eT (·)]=0.<br />

Proof. Let P = Prange(e), Q = Prange(1−e), R = Prange(eT (B)) and S = Prange(1−eT (B)). We prove<br />

that (2.26) holds. Since eT = Te, P(H) and Q(H) are T -invariant. Then by Lemma 3.3,<br />

and<br />

KT |P(H) (B) = KT (B) ∩ P(H) = R(H) ∩ P(H),<br />

c<br />

KT |P(H) B c<br />

= KT B ∩ P(H) = S(H) ∩ P(H).<br />

Hence (R ∧ P)∨ (S ∧ P)= P , and simi<strong>la</strong>rly, (R ∧ Q) ∨ (S ∧ Q) = Q. It follows that<br />

as <strong>de</strong>sired. ✷<br />

1 = P ∨ Q = (R ∧ P)∨ (S ∧ P)∨ (R ∧ Q) ∨ (S ∧ Q),


470 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Proof of Theorem 3.2. eT (∅) = 0, because PT (∅) = 0. If B1,B2 ∈ B(C) with B1 ∩ B2 =∅, then<br />

KT (B1) ∩ KT (B2) ={0}, i.e. range(eT (B1)) ∩ range(eT (B2)) ={0}. According to Lemma 3.6,<br />

[eT (B1), eT (B2)]=0 so that eT (B1)eT (B2) ∈ I( ˜M). Moreover,<br />

range eT (B1)eT (B2) ⊆ range eT (B1) ∩ range eT (B2) ={0},<br />

and we conclu<strong>de</strong> that eT (B1)eT (B2) = eT (B2)eT (B1) = 0.<br />

Now, <strong>le</strong>t (Bn) ∞ n=1 be a sequence of mutually disjoint Borel sets. Then for each N ∈ N we get<br />

from Proposition 3.4 and Lemma 2.7 that<br />

<br />

N<br />

<br />

N<br />

<br />

range eT Bn = KT Bn = KT (B1) +···+KT (BN )<br />

n=1<br />

n=1<br />

= range eT (B1) +···+eT (BN ) <br />

and<br />

ker<br />

<br />

eT<br />

N<br />

Bn<br />

n=1<br />

<br />

N<br />

= KT<br />

n=1<br />

Bn<br />

c <br />

N<br />

= KT<br />

n=1<br />

= ker eT (B1) +···+eT (BN ) .<br />

B c n<br />

<br />

N<br />

=<br />

c<br />

KT Bn =<br />

n=1<br />

n=1<br />

N<br />

ker eT (Bn) <br />

Since an e<strong>le</strong>ment e in I( ˜M) is uniquely <strong>de</strong>termined by its kernel and its range, it follows that eT<br />

is additive, i.e.<br />

<br />

N<br />

<br />

= eT (B1) +···+eT (BN ) (N ∈ N). (3.12)<br />

n=1<br />

eT<br />

n=1<br />

Bn<br />

Additivity of eT implies that<br />

<br />

∞<br />

<br />

∞<br />

<br />

eT Bn − eT (Bn) = lim<br />

N→∞<br />

n=1<br />

eT<br />

= lim<br />

N→∞ eT<br />

(the limits refer to the mea<strong>sur</strong>e topology), where<br />

<br />

τ supp<br />

<br />

∞<br />

<br />

= τ<br />

<br />

∞<br />

eT<br />

as N →∞.<br />

Bn<br />

n=N+1<br />

PT<br />

∞<br />

Bn<br />

n=1<br />

∞<br />

Bn<br />

n=N+1<br />

n=N+1<br />

<br />

<br />

N<br />

−<br />

Bn<br />

<br />

= μT<br />

eT (Bn)<br />

n=1<br />

∞<br />

n=N+1<br />

Bn<br />

<br />

<br />

→ 0,<br />

Combining this with (3.13) and Lemma 3.5, we find that eT is σ -additive as well. ✷<br />

(3.13)<br />

Note that in the case where T is a normal operator, B ↦→ PT (B) is just the spectral mea<strong>sur</strong>e<br />

of T , and eT (B) = PT (B).


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 471<br />

4. The Brown mea<strong>sur</strong>e of a set of commuting operators in M<br />

As in the previous section, <strong>le</strong>t M be a type II1 factor. The purpose of this section is to prove:<br />

Theorem 4.1. Let n ∈ N, and <strong>le</strong>t T1,...,Tn ∈ M be commuting operators. Then there is a probability<br />

mea<strong>sur</strong>e μT1,...,Tn on (Cn , B(C n )), which is uniquely <strong>de</strong>termined by<br />

μT1,...,Tn (B1 ×···×Bn) = τ<br />

n<br />

i=1<br />

PTi (Bi)<br />

where PTi (Bi) ∈ M is the projection onto KTi (Bi) (cf. Section 2).<br />

<br />

, B1,...,Bn ∈ B(C), (4.1)<br />

The i<strong>de</strong>a of proof is as follows. As mentioned in the previous section (cf. Lemma 3.6), if<br />

S ∈ M commutes with T ∈ M, then [eS(A), eT (B)]=0 for all A,B ∈ B(C). We may therefore<br />

<strong>de</strong>fine a map eT1,...,Tn from B(C)n into I( ˜M) by<br />

eT1,...,Tn (B1,...,Bn) = eT1 (B1)eT2 (B2) ···eTn (Bn), B1,...,Bn ∈ B(C). (4.2)<br />

We will then <strong>de</strong>fine ν on B(C) n by<br />

ν(B1,...,Bn) = τ supp eT1,...,Tn (B1,...,Bn) = τ(Prange(eT1 ,...,Tn (B1,...,Bn)))<br />

<br />

n<br />

= τ PTi (Bi)<br />

<br />

, B1,...,Bn ∈ B(C), (4.3)<br />

i=1<br />

and we will prove that ν extends (uniquely) to a probability mea<strong>sur</strong>e, μT1,...,Tn on (Cn , B(C n )).<br />

Theorem 4.2. Consi<strong>de</strong>r uncountab<strong>le</strong>, comp<strong>le</strong>te, separab<strong>le</strong> metric spaces (X1,d1),...,(Xn,dn).<br />

Suppose ν : B(X1) ×···×B(Xn) →[0, ∞[ is a map satisfying:<br />

(1) for all B2 ∈ B(X2), B3 ∈ B(X3),...,Bn ∈ B(Xn), B ↦→ ν(B,B2,...,Bn) is a mea<strong>sur</strong>e on<br />

(X1, B(X1));<br />

(2) for all B1 ∈ B(X1), B3 ∈ B(X3),...,Bn ∈ B(Xn), B ↦→ ν(B1,B,B3,...,Bn) is a mea<strong>sur</strong>e<br />

on (X2, B(X2));<br />

.<br />

(n) for all B1 ∈ B(X1), B2 ∈ B(X2), . . . , Bn−1 ∈ B(Xn−1), B ↦→ ν(B1,B2,...,Bn−1,B) is a<br />

mea<strong>sur</strong>e on (Xn, B(Xn)).<br />

Then there is a unique mea<strong>sur</strong>e μ on n i=1 B(Xi), such that for all B1 ∈ B(X1), B2 ∈<br />

B(X2), . . . , Bn ∈ B(Xn),<br />

μ(B1 × B2 ×···×Bn) = ν(B1,B2,...,Bn). (4.4)<br />

Proof. According to [9, Remark 1, p. 358], (Xi, B(Xi)) is Borel equiva<strong>le</strong>nt to ([0, 1], B([0, 1])),<br />

i.e. there is a bijective bimea<strong>sur</strong>ab<strong>le</strong> map φi : (Xi, B(Xi)) → ([0, 1], B([0, 1])). Therefore we


472 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

may as well assume that Xi = R (i = 1,...,n). We may also assume that ν(R, R,...,R) = 1.<br />

We <strong>de</strong>fine F : R n →[0, 1] by<br />

F(x1,...,xn) = ν ]−∞,x1],...,]−∞,xn] , x1,...,xn ∈ R. (4.5)<br />

Because of (1)–(n), F is increasing in each variab<strong>le</strong> separately and satisfies:<br />

(a) if x (k)<br />

i ↘ xi, i = 1,...,n, then F(x (k)<br />

1 ,...,x(k) n ) ↘ F(x1,...,xn),<br />

(b) if xi ↘−∞for some i ∈{1,...,n}, then F(x1,...,xn) ↘ 0,<br />

(c) if xi ↗∞for all i ∈{1,...,n}, then F(x1,...,xn) ↗ 1.<br />

Then, according to [3, Corol<strong>la</strong>ry 2.27], there is a (unique) probability mea<strong>sur</strong>e μ on (R n , B(R n ))<br />

such that for all x1,...,xn ∈ R,<br />

and<br />

μ ]−∞,x1]×···×]−∞,xn] = F(x1,...,xn). (4.6)<br />

Let x2,...,xn ∈ R be fixed but arbitrary. Then the (finite) mea<strong>sur</strong>es<br />

B ↦→ μ B ×]−∞,x2]×···×]−∞,xn] <br />

B ↦→ ν B,]−∞,x2],...,]−∞,xn] <br />

have the same distribution functions. Hence they must be i<strong>de</strong>ntical. That is, for all B ∈ B(R),<br />

μ B ×]−∞,x2]×···×]−∞,xn] = ν B,]−∞,x2],...,]−∞,xn] . (4.7)<br />

Now, <strong>le</strong>t B1 ∈ B(R) and x3,...,xn ∈ R be fixed but arbitrary. Then (4.7) shows that the (finite)<br />

mea<strong>sur</strong>es<br />

B ↦→ μ B1 × B ×]−∞,x3]×···×]−∞,xn] <br />

and<br />

B ↦→ ν B1,B,]−∞,x3],...,]−∞,xn] <br />

have the same distribution functions, so they must be i<strong>de</strong>ntical as well. That is, for all B ∈ B(R),<br />

μ B1 × B ×]−∞,x3]×···×]−∞,xn] = ν B1,B,]−∞,x3],...,]−∞,xn] . (4.8)<br />

Continuing like this we find that (4.4) holds. ✷<br />

It follows from Theorem 4.2 that in or<strong>de</strong>r to show that μT1,...,Tn exists, we must prove that<br />

(1)–(n) of Theorem 4.2 hold in the case where X1 =···=Xn = C, and where ν is given by (4.3).<br />

From now on we will, in or<strong>de</strong>r to simplify notation a litt<strong>le</strong>, consi<strong>de</strong>r the case n = 2, and we<br />

will assume that S,T ∈ M are commuting operators. It should be c<strong>le</strong>ar that the proof given<br />

below may be generalized to the case of arbitrary n ∈ N.


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 473<br />

Lemma 4.3. For fixed A ∈ B(C), νA : B(C) →[0, 1] given by<br />

(cf. (4.3)) is a mea<strong>sur</strong>e on (C, B(C)).<br />

νA(B) = ν(A,B) = τ PS(A) ∧ PT (B) , B ∈ B(C) (4.9)<br />

Proof. According to Theorem 3.2 and Definition 3.1, eT (∅) = 0, so<br />

νA(∅) = τ supp eS(A)eT (∅) = 0.<br />

Let (Bn) ∞ n=1 be a sequence of mutually disjoint sets from B(C). Then eT ( ∞ n=1 Bn) =<br />

∞n=1<br />

eT (Bn),so<br />

<br />

∞<br />

<br />

∞<br />

eS(A)eT Bn = eS(A)eT (Bn) (4.10)<br />

n=1<br />

with eS(A)eT (Bn)eS(A)eT (Bm) = eS(A)eT (Bn)eT (Bm) = 0 when n = m. Hence, by Proposition<br />

2.8,<br />

<br />

∞<br />

<br />

∞<br />

tr eS(A)eT Bn = tr eS(A)eT (Bn) . (4.11)<br />

This shows that νA is a mea<strong>sur</strong>e. ✷<br />

n=1<br />

It now follows from Lemma 4.3 and Theorem 4.2 that there is one and only one (probability)<br />

mea<strong>sur</strong>e μS,T on B(C 2 ) such that for all A,B ∈ B(C),<br />

n=1<br />

n=1<br />

μS,T (A × B) = τ supp eS,T (A, B) = τ PS(A) ∧ PT (B) , (4.12)<br />

and this proves Theorem 4.1 in the case n = 2.<br />

5. Spectral subspaces for commuting operators S,T ∈ M<br />

Theorem 5.1. Let S, T ∈ M be commuting operators, and <strong>le</strong>t B ⊆ C 2 be any Borel set. Then<br />

there is a maximal, closed, S- and T -invariant subspace K = KS,T (B) affiliated with M, such<br />

that the Brown mea<strong>sur</strong>e μS|K,T |K is concentrated on B. Let PS,T (B) ∈ M <strong>de</strong>note the projection<br />

onto KS,T (B). Then more precisely:<br />

(i) if B = B1 × B2 with B1,B2 ∈ B(C), then<br />

(ii) if B is a disjoint union of the sets (B (k)<br />

1<br />

then<br />

PS,T (B) = PS(B1) ∧ PT (B2); (5.1)<br />

PS,T (B) =<br />

∞<br />

k=1<br />

× B(k)<br />

2 )∞ k=1 , where B(k)<br />

i ∈ B(C), k ∈ N, i = 1, 2,<br />

PS<br />

(k) (k) <br />

B 1 ∧ PT B 2 ; (5.2)


474 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

(iii) and for general B ∈ B(C 2 ),<br />

Moreover,<br />

PS,T (B) =<br />

<br />

B⊆U,U⊆C 2 open<br />

PS,T (U). (5.3)<br />

μS,T (B) = τ PS,T (B) , B ∈ B C 2 . (5.4)<br />

Remark 5.2. Every non-empty, open subset of C 2 ∼ = R 4 is a disjoint union of countably many<br />

standard intervals, i.e.setsoftheform 4 i=1 ]ai,bi], where −∞


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 475<br />

μS|K,T |K<br />

(B) =<br />

=<br />

∞<br />

k=1<br />

∞<br />

k=1<br />

<br />

τP MP PS<br />

<br />

τP MP PS<br />

(k) <br />

B 1 ∧ PT<br />

(k) <br />

B 1 ∧ PT<br />

(k) <br />

B 2 ∧ P<br />

(k) <br />

B 2<br />

= 1<br />

∞<br />

tr<br />

τ(P)<br />

k=1<br />

(k) (k) <br />

eS B 1<br />

eT B 2<br />

= 1<br />

τ(P) tr<br />

<br />

∞<br />

(k) (k) <br />

eS B 1<br />

eT B 2<br />

k=1<br />

<br />

= 1<br />

τ(P) τ<br />

<br />

∞<br />

(k) (k) <br />

PS B 1 ∧ PT B 2<br />

<br />

= 1.<br />

k=1<br />

Thus, (b) holds.<br />

Now, suppose that Q ∈ M is an S- and T -invariant projection, and that μS|L,T |L is concentrated<br />

on B, where L = Q(H). Then by Lemma 3.3 and Proposition 2.8,<br />

P ∧ Q =<br />

=<br />

∞<br />

∞<br />

k=1<br />

k=1<br />

PS<br />

PS|L<br />

Hence, Proposition 2.8 and (5.6) imply that<br />

τQMQ(P ∧ Q) = trQMQ<br />

(k) (k) <br />

B 1 ∧ PT B 2<br />

<br />

∧ Q<br />

(k) (k) <br />

B 1 ∧ PT |L B 2<br />

= P range( ∞k=1 eS| L (B (k)<br />

1 )eT | L (B (k)<br />

2 )).<br />

=<br />

=<br />

∞<br />

k=1<br />

∞<br />

k=1<br />

∞<br />

k=1<br />

eS|L<br />

<br />

trQMQ eS|L<br />

<br />

τQMQ PS|L<br />

= μS|L,T |L (B)<br />

= 1.<br />

Thus, P ∧ Q = Q, and this shows that (c) holds.<br />

(k) (k) <br />

B 1<br />

eT |L B 2<br />

<br />

(k) <br />

B 1<br />

eT |L<br />

(k) <br />

B 1 ∧ PT |L<br />

(k) <br />

B 2<br />

(k) <br />

B 2


476 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

As mentioned in Remark 5.2, every open set U ⊆ C2 may be written as a union of countably<br />

many mutually disjoint sets from K. Thus, we have now proved existence of PS,T (U) for every<br />

such U, and for general B ∈ B(C2 ) we will <strong>de</strong>fine<br />

<br />

PS,T (B) :=<br />

PS,T (U). (5.7)<br />

Then again, P := PS,T (B) satisfies that<br />

(a) P is S- and T -invariant.<br />

Moreover, we prove that with K = P(H),<br />

B⊆U,U⊆C 2 open<br />

(b) μS|K,T |K is concentrated on B, and<br />

(c) P is maximal with respect to the properties (a) and (b).<br />

These properties will entail that when B happens to be a union of countably many mutually<br />

disjoint sets from K, then (5.7) agrees with the previous <strong>de</strong>finition of PS,T (B) (cf. (5.5)).<br />

Now, to see that (b) holds, note that μS|K,T |K is regu<strong>la</strong>r (cf. [6, Theorem 7.8]), and hence<br />

μS|K,T |K (B) = inf μS|K,T |K (U) | B ⊆ U, U ⊆ C2 open . (5.8)<br />

Let U be any open subset of C 2 containing B. Write U as a union of countably many mutually<br />

disjoint sets from K:<br />

Then, according to (5.6),<br />

μS|K,T |K<br />

(U) =<br />

U =<br />

∞<br />

k=1<br />

∞ (k)<br />

B<br />

k=1<br />

1<br />

<br />

τP MP PS<br />

and using Proposition 2.8 and Lemma 3.3 we find that<br />

μS|K,T |K (U) = trP MP<br />

= τP MP<br />

∞<br />

k=1<br />

∞<br />

k=1<br />

= τP MP<br />

= τP MP (P )<br />

= 1,<br />

eS|K<br />

PS|K<br />

<br />

× B(k)<br />

2 .<br />

(k) <br />

B 1 ∧ PT<br />

PS,T (U) ∧ P <br />

(k) <br />

B 2 ∧ P ,<br />

(k) (k) <br />

B 1<br />

eT |K B 2<br />

<br />

(k) (k) <br />

B 1 ∧ PT |K B 2<br />

<br />

where PS,T (U) is given by (5.5). Hence by (5.8), μS|K,T |K is concentrated on B.


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 477<br />

Finally, if Q ∈ M is any S- and T -invariant projection, and if μS|L,T |L is concentrated on B,<br />

where L = Q(H), then μS|L,T |L is concentrated on U for every open set U containing B. Hence,<br />

by the first part of the proof, Q PS,T (U) for every such U, and it follows from the <strong>de</strong>finition<br />

of PS,T (B) that Q P .<br />

Concerning (5.4), note that if B = B1 × B2, where B1,B2 ∈ B(C), then, by the <strong>de</strong>finitions<br />

of μS,T and PS,T (B), (5.4) holds. If B is a disjoint union of sets (B (k) ) ∞ k=1 = (B(k)<br />

1 × B(k)<br />

2 )∞ k=1 ,<br />

where B (k)<br />

i ∈ B(C), k ∈ N, i = 1, 2, then<br />

μS,T (B) =<br />

∞<br />

τ (k)<br />

PS,T B 1<br />

k=1<br />

Applying Proposition 2.8 we thus find that<br />

<br />

× B(k)<br />

2 =<br />

<br />

∞<br />

μS,T (B) = τ supp<br />

k=1<br />

k=1<br />

eS<br />

∞<br />

τ supp (k) (k) <br />

eS B 1<br />

eT B 2 .<br />

k=1<br />

(k) (k) <br />

B 1<br />

eT B 2<br />

<br />

<br />

∞<br />

= τ supp (k) (k) <br />

eS B 1<br />

eT B 2<br />

<br />

= τ<br />

∞<br />

k=1<br />

PS,T<br />

= τ PS,T (B) .<br />

B (k)<br />

1<br />

Finally, for general B ∈ B(C 2 ), since μS,T is regu<strong>la</strong>r,<br />

<br />

× B(k)<br />

2<br />

<br />

μS,T (B) = inf μS,T (U) | B ⊆ U ⊆ C 2 ,Uopen <br />

= inf τ PS,T (U) | B ⊆ U ⊆ C 2 ,Uopen <br />

<br />

<br />

<br />

= τ<br />

PS,T (U)<br />

B⊆U⊆C 2 ,Uopen<br />

= τ PS,T (B) . ✷<br />

The proof given above may c<strong>le</strong>arly be generalized to the case of n commuting operators<br />

T1,...,Tn ∈ M, so that Theorem 5.1 has a slightly more general version:<br />

Theorem 5.3. Let n ∈ N, <strong>le</strong>tT1,...,Tn ∈ M be commuting operators, and <strong>le</strong>t B ⊆ Cn be any<br />

Borel set. Then there is a maximal closed subspace, K = KT1,...,Tn (B), affiliated with M which<br />

is Ti-invariant for every i ∈{1,...,n}, and such that the Brown mea<strong>sur</strong>e μT1|K,...,Tn|K is concentrated<br />

on B. Let PT1,...,Tn (B) ∈ M <strong>de</strong>note the projection onto KT1,...,Tn (B). Then more precisely:<br />

(i) if B = B1 ×···×Bn with Bi ∈ B(C), then<br />

PT1,...,Tn<br />

(B) =<br />

n<br />

PTi (Bi); (5.9)<br />

i=1


478 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

(ii) if B is a disjoint union of sets (B (k) ) ∞ k=1<br />

k ∈ N, i = 1,...,n, then<br />

(iii) and for general B ∈ B(C n ),<br />

Moreover, for every B ∈ B(C n ),<br />

PT1,...,Tn<br />

PT1,...,Tn<br />

(B) =<br />

(B) =<br />

6. An alternative characterization of μS,T<br />

= (B(k)<br />

1 ×···×B(k) n ) ∞ k=1 , where B(k)<br />

i ∈ B(C),<br />

∞<br />

k=1<br />

<br />

B⊆U, U⊆C n open<br />

(k)<br />

PT1,...,Tn, B ; (5.10)<br />

PT1,...,Tn (U). (5.11)<br />

μT1,...,Tn (B) = τ PT1,...,Tn (B) . (5.12)<br />

In this final section we are going to give a characterization of the Brown mea<strong>sur</strong>e of two<br />

commuting operators in M, which is different from the one we gave in Theorem 4.1. Recall<br />

from [4] that for T ∈ M, the Brown mea<strong>sur</strong>e of T , μT , is the unique compactly supported Borel<br />

probability mea<strong>sur</strong>e on C which satisfies the i<strong>de</strong>ntity<br />

τ log |T − λ1| <br />

= log |z − λ| dμT (z) (6.1)<br />

for all λ ∈ C.<br />

We are going to prove that a simi<strong>la</strong>r property characterizes μS,T .<br />

C<br />

Theorem 6.1. Let S,T ∈ M be commuting operators. Then μS,T is the unique compactly supported<br />

Borel probability mea<strong>sur</strong>e on C2 which satisfies the i<strong>de</strong>ntity<br />

τ log |αS + βT − 1| <br />

= log |αz + βw − 1| dμS,T (z, w) (6.2)<br />

for all α, β ∈ C.<br />

C 2<br />

Remark 6.2. Let S,T ∈ M be as in Theorem 6.1. Note that if μS,T satisfies (6.2) for all α, β ∈ C,<br />

then for all α, β, λ ∈ C,<br />

τ log |αS + βT − λ1| <br />

= log |αz + βw − λ| dμS,T (z, w). (6.3)<br />

C 2<br />

This is c<strong>le</strong>ar for λ = 0, and for λ = 0, (6.3) follows from the fact that two subharmonic functions<br />

<strong>de</strong>fined in C coinci<strong>de</strong> iff they agree almost everywhere with respect to Lebesgue mea<strong>sur</strong>e. It now<br />

follows from Brown’s characterization of μαS+βT that μαS+βT is the push-forward mea<strong>sur</strong>e να,β<br />

of μS,T via the map (z, w) ↦→ αz + βw. On the other hand, if να,β = μαS+βT , then (6.2) holds.


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 479<br />

Recall from [7] that the modified spectral radius of T ∈ M, r ′ (T ), is <strong>de</strong>fined by<br />

r ′ (T ) := max |z| z ∈ supp(μT ) . (6.4)<br />

Also recall from [7, Corol<strong>la</strong>ry 2.6] that in fact<br />

r ′ <br />

(T ) = lim lim T<br />

p→∞ n→∞<br />

n 1/n<br />

<br />

. (6.5)<br />

p/n<br />

Lemma 6.3. Let S,T ∈ M be commuting operators. Then the modified spectral radii, r ′ (S),<br />

r ′ (T ), r ′ (ST ) and r ′ (S + T), satisfy the inequalities<br />

and<br />

r ′ (ST ) r ′ (S) · r ′ (T ), (6.6)<br />

r ′ (S + T) r ′ (S) + r ′ (T ). (6.7)<br />

Proof. (6.6) follows from (6.5) and the generalized Höl<strong>de</strong>r inequality (cf. [5]): for A,B ∈ M<br />

and for 0


480 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Repeating this argument, we find that for arbitrary θ ∈[0, 2π[,<br />

i.e.<br />

Since θ was arbitrary, we conclu<strong>de</strong> that<br />

and this proves (6.7). ✷<br />

supp(μ e iθ (S+T) ) ⊆ z ∈ C Re z r ′ (S) + r ′ (T ) ,<br />

supp(μS+T ) ⊆ z ∈ C Re e −iθ z r ′ (S) + r ′ (T ) .<br />

supp(μS+T ) ⊆ B 0,r ′ (S) + r ′ (T ) ,<br />

Lemma 6.4. Let S,T ∈ M be commuting operators, and <strong>le</strong>t α, β ∈ C. Then μαS,βT is the pushforward<br />

mea<strong>sur</strong>e of μS,T via the map hα,β : C × C → C × C given by<br />

hα,β(z, w) = (αz, βw).<br />

Proof. Recall that μαS,βT is uniquely <strong>de</strong>termined by the property that for all B1,B2 ∈ B(C),<br />

Now, it is easily seen that for α = 0 and β = 0,<br />

Hence,<br />

μαS,βT (B1 × B2) = τ PαS(B1) ∧ PβT (B2) . (6.8)<br />

PαS(B1) = PS<br />

<br />

1<br />

α B1<br />

<br />

<br />

1<br />

μαS,βT (B1 × B2) = τ PS<br />

α B1<br />

<br />

1<br />

∧ PT<br />

and PβT (B2) = PT<br />

β B2<br />

<br />

1<br />

β B2<br />

<br />

.<br />

<br />

1<br />

= μS,T<br />

α B1 × 1<br />

β B2<br />

<br />

−1<br />

= μS,T hα,β (B1 × B2) . (6.9)<br />

If for instance α = 0, then PαS(B1) = 0if0/∈ B1 and PαS(B1) = 1 if 0 ∈ B1. It then follows that<br />

(6.9) holds in this case as well. Simi<strong>la</strong>r arguments apply if β = 0. ✷<br />

Proof of Theorem 6.1. As noted in Remark 6.2, it suffices to prove that for all α, β ∈ C, μαS+βT<br />

is the push-forward mea<strong>sur</strong>e of μS,T via the map (z, w) ↦→ αz + βw. At first we will consi<strong>de</strong>r<br />

the case α = β = 1. Define a : C × C → C by<br />

We are going to prove that for all B ∈ B(C),<br />

a(z,w) = z + w (z,w∈ C).<br />

−1<br />

μS+T (B) = μS,T a (B) . (6.10)


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 481<br />

It suffices to show that for every open set U ⊆ C,<br />

−1<br />

μS+T (U) μS,T a (U) . (6.11)<br />

In<strong>de</strong>ed, if this holds, then by regu<strong>la</strong>rity of μS+T and a(μS,T ), for every Borel set B ⊆ C,<br />

μS+T (B) = inf μS+T (U) <br />

B ⊆ U, U open<br />

inf a(μS,T )(U) <br />

B ⊆ U, U open<br />

= μS,T<br />

a −1 B .<br />

Since both mea<strong>sur</strong>es are probability mea<strong>sur</strong>es, and the above inequality holds for both B and B c ,<br />

we must have i<strong>de</strong>ntity. That is, (6.10) holds.<br />

Now, <strong>le</strong>t U ⊆ C be any open set. Then V := a −1 (U) is open in C 2 and we may write V as a<br />

countab<strong>le</strong> union of mutually disjoint “boxes,”<br />

where for z ∈ C and δ>0,<br />

V =<br />

∞<br />

I(zn,δn) × I(wn,δn),<br />

n=1<br />

I(z,δ):= w ∈ C Re(z) − δ


482 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

r ′ S + T − (zn + wn)1 <br />

′<br />

r P(H)<br />

[S − zn1] <br />

′<br />

+ r P(H)<br />

[T − wn1] <br />

<br />

P(H)<br />

r ′ [S − zn1] <br />

′<br />

+ r PS(I (zn,δn))(H)<br />

[T − wn1] <br />

<br />

PT (I (wn,δn))(H)<br />

2 √ 2 δn,<br />

and it follows that μS+T |P(H) is concentrated on B(zn + wn, 2 √ 2δn) ⊆ U. Hence, P <br />

PS+T (U), and we are done.<br />

Now, if α, β ∈ C, then we conclu<strong>de</strong> from the above and Lemma 6.4 that<br />

−1 −1 −1<br />

μαS+βT (B) = μαS,βT a (B) = μS,T hα,β a (B) = μS,T (a ◦ hα,β) −1 (B) ,<br />

and since (a ◦ hα,β)(z, w) = αz + βw, this comp<strong>le</strong>tes the proof of the i<strong>de</strong>ntity (6.2).<br />

To prove uniqueness of μS,T , suppose that ν is a compactly supported Borel probability mea<strong>sur</strong>e<br />

on C2 which satisfies the i<strong>de</strong>ntity (6.2) for all α, β ∈ C. That is, for all α, β ∈ C, μαS+βT is<br />

the push-forward mea<strong>sur</strong>e of ν via the map (z, w) ↦→ αz + βw. Then, to prove that ν = μS,T ,it<br />

suffices to prove that for all y = (y1,...,y4) ∈ R4 ,<br />

<br />

e i(y,x) <br />

dμS,T (x) = e i(y,x) dν(x) (6.15)<br />

R 4<br />

(here we i<strong>de</strong>ntify C with R 2 ). For x = (x1,...,x4) ∈ R 4 and y = (y1,...,y4) ∈ R 4 , note that<br />

R 4<br />

(y, x) = Re (y1 − iy2)(x1 + ix2) + (y3 − iy4)(x3 + ix4) ,<br />

and hence with α = y1 − iy2 and β = y3 − iy4 we find that<br />

<br />

e i(y,x) <br />

dμS,T (x) = e iRe(αz+βw) <br />

dμS,T (z, w) =<br />

as <strong>de</strong>sired. ✷<br />

C 2<br />

<br />

=<br />

C 2<br />

R 4<br />

e iRe(αz+βw) <br />

dν(z,w) =<br />

R 4<br />

C<br />

e iRez dμαS+βT (z)<br />

e i(y,x) dν(x),<br />

Remark 6.5. In the proof above it was shown that for U ⊆ C an open set, we have the following<br />

inequality:<br />

−1<br />

PS+T (U) PS,T a (U) . (6.16)<br />

But it was also shown that the two projections above have the same trace:<br />

τ PS+T (U) −1 −1<br />

= μS+T (U) = μS,T a (U) = τ PS,T a (U) .<br />

Hence, the two projections in (6.16) are i<strong>de</strong>ntical, and by Theorem 5.1(iii), for every Borel set<br />

B ⊆ C, we must have that<br />

−1<br />

PS+T (B) = PS,T a (B) . (6.17)


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 483<br />

As in the previous section, one can easily generalize the proof given above to the case of an<br />

arbitrary finite set of commuting operators, {T1,...,Tn}. That is, we actually have the following<br />

alternative <strong>de</strong>scription of μT1,...,Tn .<br />

Theorem 6.6. Let n ∈ N, and <strong>le</strong>t T1,...,Tn be mutually commuting operators in M. Then<br />

μT1,...,Tn is the unique compactly supported Borel probability mea<strong>sur</strong>e on Cn which satisfies<br />

the i<strong>de</strong>ntity<br />

τ log |α1T1 + ··· +αnTn − 1| <br />

= log |α1z1 + ··· +αnzn − 1| dμT1,...,Tn (z1,...,zn) (6.18)<br />

for all α1,...,αn ∈ C.<br />

Lemma 6.7. Define a : C 2 → C by<br />

C n<br />

a(z,w) = z + w,<br />

and <strong>le</strong>t U ⊆ C be an open set. Then for every pair (S, T ) of commuting operators in M,wemay<br />

write V := a−1 (U) as a countab<strong>le</strong> disjoint union of sets (I (zn,δn) × I(wn,δn)) ∞ n=1 , where for<br />

z ∈ C and δ>0,<br />

I(z,δ):= w ∈ C Re(z) − δ


484 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Remark 6.8. Let S ∈ M be invertib<strong>le</strong>, and <strong>le</strong>t T1,...,Tn be mutually commuting operators<br />

in M. Then<br />

μ S −1 T1S,...,S −1 TnS<br />

= μT1,...,Tn . (6.20)<br />

In<strong>de</strong>ed, this follows from the characterization of μT1,...,Tn given in Theorem 6.6 and from the fact<br />

that for all T ∈ M, μS−1TS = μT (cf. [4]).<br />

Proposition 6.9. Let S,T ∈ M be commuting operators. Then μST is the push-forward mea<strong>sur</strong>e<br />

of μS,T via the map m : (z, w) ↦→ zw.<br />

Proof. The proof is essentially the same as the one we gave above when consi<strong>de</strong>ring the map<br />

a : (z, w) ↦→ z + w. Again it suffices to show that for every open set U ⊆ C,<br />

−1<br />

μST (U) μS,T m (U) , (6.21)<br />

and for such an open set U we write V := m −1 (U) as a countab<strong>le</strong> union of mutually disjoint<br />

“boxes” as in (6.12), but this time we make <strong>sur</strong>e that δn > 0 is so small that<br />

B znwn, √ <br />

2 δn T +|zn| ⊆ U. (6.22)<br />

As in the previous case, one only has to show that for every n ∈ N,<br />

<br />

PST (U) PS I(zn,δn) <br />

∧ PT I(wn,δn) .<br />

Fix n ∈ N and set P = PS(I (zn,δn)) ∧ PT (I (wn,δn)). Since<br />

r ′ [S − zn1] <br />

′<br />

r P(H)<br />

[S − zn1] √<br />

2 δn,<br />

PS(I (zn,δn))(H)<br />

and<br />

and since<br />

r ′ [T − wn1] <br />

′<br />

r P(H)<br />

[T − wn1] √<br />

2 δn,<br />

PT (I (wn,δn))(H)<br />

ST − znwn1 = (S − zn1)T + zn(T − wn1),<br />

we have (cf. Lemma 6.3) that<br />

r ′ [ST − znwn1] <br />

′<br />

r P(H)<br />

(S − zn1)T <br />

+|zn|r P(H)<br />

′ [T − wn1] <br />

<br />

P(H)<br />

r ′ [S − zn1] <br />

T +|zn|r P(H)<br />

′ [T − wn1] <br />

<br />

P(H)<br />

√ <br />

2 δn T +|zn| .<br />

Thus, μST |P(H) is concentrated on B(znwn, √ 2δn(T +|zn|)) ⊆ U, and therefore P PST (U),<br />

as <strong>de</strong>sired. ✷


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 485<br />

Remark 6.10. As in the additive case, we infer from the proof given above that for every Borel<br />

set B ⊆ C we have:<br />

−1<br />

PST (B) = PS,T m (B) . (6.23)<br />

Proposition 6.11. Consi<strong>de</strong>r type II1 factors M1 and M2 with faithful tracial states τ1 and τ2,<br />

respectively. Let S ∈ M1 and T ∈ M2. Then<br />

and it follows that<br />

μS⊗1,1⊗T = μS ⊗ μT , (6.24)<br />

μS⊗1+1⊗T = μS ∗ μT , (6.25)<br />

μS⊗T = μS ⋆μT , (6.26)<br />

where ∗ (⋆, respectively) <strong>de</strong>notes additive (multiplicative, respectively) convolution.<br />

Proof. μS⊗1+1⊗T (μS⊗T , respectively) is the push-forward mea<strong>sur</strong>e of μS⊗1,1⊗T via the map<br />

a : (z, w) ↦→ z + w (m : (z, w) ↦→ zw), respectively), and μS ∗ μT (μS ⋆μT , respectively) is the<br />

push-forward mea<strong>sur</strong>e of μS ⊗μT via that same map. Thus, (6.25) and (6.26) follow from (6.24).<br />

To see that the <strong>la</strong>tter holds, <strong>le</strong>t B1,B2 ∈ B(C). It is easily seen that<br />

Hence,<br />

This proves (6.24). ✷<br />

PS⊗1(B1) = PS(B1) ⊗ 1 and P1⊗T (B2) = 1 ⊗ PT (B2).<br />

μS⊗1,1⊗T (B1 × B2) = (τ1 ⊗ τ2) PS(B1) ⊗ 1 ∩ 1 ⊗ PT (B2) <br />

<br />

= τ1 PS(B1) <br />

τ2 PT (B2) <br />

= μS(B1)μT (B2)<br />

7. Polynomials in n commuting variab<strong>le</strong>s<br />

In this final section we will prove:<br />

= (μS ⊗ μT )(B1 × B2).<br />

Theorem 7.1. Let n ∈ N, and <strong>le</strong>t q be a polynomial in n commuting variab<strong>le</strong>s, i.e. q ∈<br />

C[z1,...,zn]. Then for every n-tup<strong>le</strong> (T1,...,Tn) of commuting operators in M, one has that<br />

μq(T1,...,Tn) = q(μT1,...,Tn ), (7.1)<br />

where q(μT1,...,Tn ) is the push-forward mea<strong>sur</strong>e of μT1,...,Tn via q : Cn → C.<br />

The proof relies on the previous sections and a few technical <strong>le</strong>mmas.


486 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Lemma 7.2. Given n ∈ N and commuting operators T1,...,Tn ∈ M. Let 1 i


H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 487<br />

Lemma 7.3. Let n ∈ N and <strong>le</strong>t α ∈ C. Define an, m (α)<br />

n : C n+1 → C by<br />

an(z1,...,zn,zn+1) = (z1,...,zn + zn+1), (7.6)<br />

m (α)<br />

n (z1,...,zn,zn+1) = (z1,...,αznzn+1). (7.7)<br />

Then for any (n + 1)-tup<strong>le</strong> (T1,...,Tn+1) of commuting operators in M one has that<br />

and<br />

μT1,...,Tn−1,Tn+Tn+1<br />

μT1,...,Tn−1,αTnTn+1<br />

= an(μT1,...,Tn+1 ) (7.8)<br />

= m(α) n (μT1,...,Tn+1 ). (7.9)<br />

Proof. The proof is based on Lemma 7.2 and the fact that by (6.17) and (6.23), for any Borel set<br />

B ⊆ C we have<br />

−1<br />

PTn+Tn+1 (B) = PTn,Tn+1 1 (B) (7.10)<br />

and<br />

−1 (α)<br />

PαTnTn+1 (B) = PαTn,Tn+1 m (B) = PTn,Tn+1 m −1 <br />

(B) .<br />

In or<strong>de</strong>r to prove (7.8), consi<strong>de</strong>r arbitrary Borel sets B1,...,Bn ⊆ C. We must show that<br />

μT1,...,Tn−1,Tn+Tn+1 (B1 ×···×Bn) = μT1,...,Tn+1<br />

a −1<br />

n (B1 ×···×Bn) ,<br />

i.e. that<br />

μT1,...,Tn−1,Tn+Tn+1 (B1<br />

<br />

×···×Bn) = μT1,...,Tn+1 B1 ×···×Bn−1 × a −1 (Bn) . (7.11)<br />

But by Lemma 7.2 and by (7.10),<br />

PT1,...,Tn−1,Tn+Tn+1 (B1 ×···×Bn) = PT1 (B1) ∧···∧PTn−1 (Bn−1) ∧ PTn+Tn+1 (Bn)<br />

= PT1 (B1) ∧···∧PTn−1 (Bn−1)<br />

−1<br />

∧ PTn,Tn+1 a (Bn) <br />

<br />

B1 ×···×Bn−1 × a −1 (Bn) ,<br />

= PT1,...,Tn+1<br />

and this proves (7.11). (7.9) follows in a simi<strong>la</strong>r way. ✷<br />

Lemma 7.4. Let n ∈ N, and <strong>le</strong>t σ ∈ Sn (the group of permutations of {1, 2,...,n}). Then for any<br />

n-tup<strong>le</strong> (T1,...,Tn) of commuting operators in M,<br />

μTσ(1),...,Tσ(n)<br />

= σ(μT1,...,Tn ), (7.12)<br />

where i<strong>de</strong>ntify σ with the corresponding permutation of coordinates C n → C n .<br />

Proof. This follows easily from Theorem 4.1. ✷


488 H. Schultz / Journal of Functional Analysis 236 (2006) 457–489<br />

Lemma 7.5. For n ∈ N and 1 i n <strong>de</strong>fine fi : C n → C n+1 by<br />

fi(z1,...,zn) = (z1,...,zn,zi).<br />

Then for n commuting operators T1,...,Tn ∈ M, one has that<br />

μT1,...,Tn,Ti<br />

Proof. Given Borel sets B1,...,Bn+1 ⊆ C we must show that<br />

C<strong>le</strong>arly,<br />

μT1,...,Tn,Ti (B1 ×···×Bn+1) = μT1,...,Tn<br />

= fi(μT1,...,Tn ). (7.13)<br />

f −1<br />

i (B1 ×···×Bn+1) . (7.14)<br />

f −1<br />

i (B1 ×···×Bn+1) = B1 ×···×(Bi ∩ Bn+1) ×···×Bn<br />

so that the right-hand si<strong>de</strong> of (7.14) is<br />

τ PT1 (B1) ∧···∧PTi (Bi<br />

<br />

∩ Bn+1) ∧···∧PTn (Bn)<br />

= τ PT1 (B1) ∧···∧PTi (Bi) ∧ PTi (Bn+1) ∧···∧PTn (Bn) .<br />

But this is exactly the <strong>le</strong>ft-hand si<strong>de</strong> of (7.14) and we are done. ✷<br />

We will not give the proof of Theorem 7.1 in full generality but rather, by way of an examp<strong>le</strong>,<br />

illustrate how it goes. Consi<strong>de</strong>r for instance 3 commuting operators T1,T2,T3 ∈ M and the<br />

polynomial q ∈ C[z1,z2,z3] given by<br />

At first <strong>de</strong>fine φ1 : C 3 → C 5 by<br />

q(z1,z2,z3) = 1 + 2z 2 2 + z1z2z3. (7.15)<br />

φ1(z1,z2,z3) = (z2,z2,z1,z2,z3).<br />

By repeated use of Lemmas 7.4 and 7.5 we find that<br />

Next <strong>de</strong>fine φ2 : C 5 → C 2 by<br />

μT2,T2,T1,T2,T3<br />

= φ1(μT1,T2,T3 ).<br />

φ2(z1,...,z5) = (2z1z2,z3z4z5),<br />

and by repeated use of (7.9) and Lemma 7.4 conclu<strong>de</strong> that<br />

With φ3 : C 2 → C given by<br />

μ 2T 2 2 ,T1T2T3 = (φ2 ◦ φ1)(μT1,T2,T3 ).<br />

φ3(z1,z2) = z1 + z2


we now have (cf. (7.8)) that<br />

H. Schultz / Journal of Functional Analysis 236 (2006) 457–489 489<br />

μ(q−1)(T1,T2,T3) = (φ3 ◦ φ2 ◦ φ1)(μT1,T2,T3 ) = (q − 1)(μT1,T2,T3 ).<br />

It is now a simp<strong>le</strong> matter to show that for all λ ∈ C,<br />

τ log q(T1,T2,T3) − λ1 <br />

= log |z − λ| dq(μT1,T2,T3 )(z),<br />

and then by Brown’s characterization of μq(T1,T2,T3), μq(T1,T2,T3) = q(μT1,T2,T3 ), as <strong>de</strong>sired.<br />

Acknow<strong>le</strong>dgments<br />

C<br />

Part of this work was carried out whi<strong>le</strong> I visited the UCLA Department of Mathematics. I<br />

want to thank the <strong>de</strong>partment, and especially the Operator Algebra Group, for their hospitality.<br />

I also thank my advisor, Uffe Haagerup, with whom I had enlightening discussions about this<br />

work.<br />

References<br />

[1] L. Aagaard, The non-microstates free entropy dimension of DT-operators, J. Funct. Anal. 213 (1) (2004) 176–205.<br />

[2] P. Ainsworth, Ubegrænse<strong>de</strong> operatorer affilieret med en en<strong>de</strong>lig von Neumann algebra, Master thesis, University of<br />

O<strong>de</strong>nse, 1985.<br />

[3] L. Breiman, Probability, Addison–Wes<strong>le</strong>y, 1968.<br />

[4] L.G. Brown, Lidskii’s theorem in the type II case, in: H. Araki, E. Effros (Eds.), Geometric Methods in Operator<br />

Algebras, Kyoto, 1983, in: Pitman Res. Notes Math. Ser., vol. 123, Longman Sci. Tech., 1986, pp. 1–35.<br />

[5] T. Fack, H. Kosaki, Generalized s-numbers of τ -mea<strong>sur</strong>ab<strong>le</strong> operators, Pacific J. Math. 123 (1986) 269–300.<br />

[6] G.B. Fol<strong>la</strong>nd, Real Analysis, Mo<strong>de</strong>rn Techniques and Their Applications, Wi<strong>le</strong>y, 1984.<br />

[7] U. Haagerup, H. Schultz, Invariant subspaces of operators in a general II1-factor, preprint, 2005.<br />

[8] R.V. Kadison, J.R. Ringrose, Fundamentals of the Theory of Operator Algebras, vol. I, Aca<strong>de</strong>mic Press, 1983.<br />

[9] K. Kuratowski, Topology, vol. I, second ed., Aca<strong>de</strong>mic Press, London, 1966.<br />

[10] E. Nelson, Notes on non-commutative integration, J. Funct. Anal. 15 (1974) 103–116.


Journal of Functional Analysis 236 (2006) 490–516<br />

www.elsevier.com/locate/jfa<br />

Function spaces between BMO and critical<br />

Sobo<strong>le</strong>v spaces ✩<br />

Jean Van Schaftingen<br />

Département <strong>de</strong> Mathématique, Université Catholique <strong>de</strong> Louvain, 2 chemin <strong>du</strong> Cyclotron,<br />

1348 Louvain-<strong>la</strong>-Neuve, Belgium<br />

Received 29 November 2005; accepted 1 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 2 May 2006<br />

Communicated by H. Brezis<br />

Abstract<br />

The function spaces Dk(Rn ) are intro<strong>du</strong>ced and studied. The <strong>de</strong>finition of these spaces is based on a regu<strong>la</strong>rity<br />

property for the critical Sobo<strong>le</strong>v spaces Ws,p (Rn ),wheresp = n, obtained by J. Bourgain, H. Brezis,<br />

New estimates for the Lap<strong>la</strong>cian, the div–curl, and re<strong>la</strong>ted Hodge systems, C. R. Math. Acad. Sci. Paris<br />

338 (7) (2004) 539–543 (see also J. Van Schaftingen, Estimates for L1-vector fields, C. R. Math. Acad.<br />

Sci. Paris 339 (3) (2004) 181–186). The spaces Dk(Rn ) contain all the critical Sobo<strong>le</strong>v spaces. They are<br />

embed<strong>de</strong>d in BMO(Rn ), but not in VMO(Rn ). Moreover, they have some extension and trace properties<br />

that BMO(Rn ) does not have.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Critical Sobo<strong>le</strong>v spaces; BMO<br />

1. Intro<strong>du</strong>ction<br />

1.1. Integrals with divergence-free vector-fields<br />

When p


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 491<br />

p>nit is embed<strong>de</strong>d in the space of Höl<strong>de</strong>r continuous functions of exponent α, C 0,α (R n ), with<br />

α = 1 − n/p [1,5,19].<br />

The case p = n is more <strong>de</strong>licate. When n>1, functions in W 1,n (R n ) do not need to be continuous<br />

or boun<strong>de</strong>d, but have many properties in common with such functions. This is expressed<br />

for examp<strong>le</strong> by the embedding of W 1,n (R n ) in the spaces BMO(R n ) and VMO(R n ) of functions<br />

of boun<strong>de</strong>d and vanishing mean oscil<strong>la</strong>tion [6]. These consi<strong>de</strong>rations are also valid for fractional<br />

Sobo<strong>le</strong>v spaces W s,p (R n ), with sp = n.<br />

Another property of critical Sobo<strong>le</strong>v space was recently obtained by Bourgain and Brezis [3,<br />

23]: for every vector field ϕ ∈ (L 1 ∩ C)(R n ; R n ) and u ∈ W s,p (R n ),ifdivϕ = 0 in the sense of<br />

distributions, then<br />

<br />

<br />

<br />

uϕ dx<br />

Cs,pϕL1 (Rn ) uWs,p (Rn ). (1.1)<br />

R n<br />

There is no such property for BMO(R n ) or for VMO(R n ) (see [2] and Remark 5.2).<br />

A natural question is the re<strong>la</strong>tionship between (1.1) and the embedding of W s,p (R n ) in the<br />

spaces BMO(R n ) and VMO(R n ). In or<strong>de</strong>r to answer it, we <strong>de</strong>fine, for n 1, the <strong>semi</strong>norm<br />

and the vector space<br />

uDn−1(R n ) = sup<br />

ϕ∈D(R n ;R n )<br />

div ϕ=0<br />

ϕ L 1 (R n ) 1<br />

<br />

<br />

<br />

uϕ dx<br />

<br />

n<br />

Dn−1 R = u ∈ D ′ R n : uDn−1(Rn ) < ∞ .<br />

Here D(R N ; R N ) is the space of compactly supported smooth vector fields and D ′ (R n ) is the<br />

space of distributions [16]. The subscript n − 1 will be justified by further extensions. By the<br />

inequality (1.1), W s,p (R n ) is embed<strong>de</strong>d in Dn−1(R n ).<br />

The question of the previous paragraph is answered as follows: VMO(R n ) is not embed<strong>de</strong>d in<br />

Dn−1(R n ) (Proposition 5.1), and Dn−1(R n ) is embed<strong>de</strong>d in BMO(R n ) (Theorem 5.3). Moreover,<br />

if u ∈ Dn−1(R n ) is continuous, and k 2, then u| R k BMO(R k ) CuDn−1(R n ) (Theorems 3.4<br />

and 5.3). This inequality remains open when k = 1.<br />

The proof of the embedding of Dn−1(R n ) in BMO(R n ) is based on the <strong>du</strong>ality between<br />

BMO(R n ) and the Hardy space H 1 (R n ), and on a <strong>de</strong>composition of every function in H 1 (R n ) as<br />

a sum of some components of divergence-free vector-fields, with a suitab<strong>le</strong> control on the norms.<br />

The inequality (1.1) was prece<strong>de</strong>d by a geometric counterpart [4]: for every closed rectifiab<strong>le</strong><br />

curve γ ∈ C 1 (S 1 ; R n ) and u ∈ (C ∩ W 1,n )(R n ),<br />

R n<br />

R n<br />

(1.2)<br />

<br />

<br />

<br />

u γ(t) <br />

<br />

˙γ(t)dt<br />

Cs,p˙γ L1 (S1 ) uWs,p (Rn ). (1.3)


492 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

(See [22] for an e<strong>le</strong>mentary proof.) The right-hand si<strong>de</strong> of (1.3) could also be used to <strong>de</strong>fine<br />

a <strong>semi</strong>norm on continuous functions. By the arguments of [3], based on a <strong>de</strong>composition of<br />

divergence-free vector-fields in so<strong>le</strong>noids of Smirnov [18], one has in fact<br />

uDn−1(R n ) = sup<br />

γ ∈C 1 (S 1 ;R n )<br />

<br />

1 <br />

<br />

˙γ u<br />

L1 (S1 )<br />

γ(t) <br />

<br />

˙γ(t)dt<br />

.<br />

An open prob<strong>le</strong>m is whether restricting the curves on the right-hand si<strong>de</strong> to be contained in<br />

k-dimensional p<strong>la</strong>nes, to triang<strong>le</strong>s or to circ<strong>le</strong>s would yield an equiva<strong>le</strong>nt norm. The restriction<br />

to curves contained in k-dimensional p<strong>la</strong>nes is equiva<strong>le</strong>nt to requiring ϕ in (1.2) to have a range<br />

whose dimensions is at most k, see Section 6.4.<br />

If s 1, sp = n, and u ∈ W s,p (R n ), then u+ ∈ W s,p (R n ). This property also holds in<br />

BMO(R n ). We do not know whether it holds for Dn−1(R n ). The question whether, for a given<br />

ϕ : R → R one has ϕ(u) ∈ Dn−1(R n ) whenever u ∈ Dn−1(R n ) remains open when ϕ is not affine.<br />

1.2. Integrals with curl-free vector-fields<br />

When s = 1 and p = n = 2, the inequality (1.1) is in fact a <strong>du</strong>al statement of the Sobo<strong>le</strong>v–<br />

Nirenberg embedding<br />

S 1<br />

g L n/(n−1) (R n ) CDg L 1 (R n ) . (1.4)<br />

When s = 1 and p = n>2, the estimate (1.1) is stronger than the embedding (1.4). If n = 3,<br />

(1.4) yields by <strong>du</strong>ality that, for every ϕ ∈ D(R3 ; R3 ) and u ∈ W1,3 (R3 ), if curl ϕ = 0 in the sense<br />

of distributions,<br />

<br />

<br />

<br />

uϕ dx<br />

CϕL1 (R3 ) uWn (R3 ) . (1.5)<br />

R 3<br />

For u ∈ W s,p (R 3 ) with sp = 3, this inequality can be <strong>de</strong><strong>du</strong>ced from (1.1) recalling that, for every<br />

e ∈ R 3<br />

div(ϕ × e) = (curl ϕ) · e. (1.6)<br />

In R 3 , one can now investigate the re<strong>la</strong>tionship between (1.1), (1.5), and the embedding of<br />

W s,p (R n ) in the spaces BMO(R n ) and VMO(R n ). We <strong>de</strong>fine therefore the <strong>semi</strong>norm<br />

and the vector space<br />

u D1(R 3 )<br />

= sup<br />

ϕ∈D(R 3 ;R 3 )<br />

curl ϕ=0<br />

ϕ L 1 (R 3 ) 1<br />

<br />

<br />

<br />

uϕ dx<br />

<br />

3<br />

D1 R = u ∈ D ′ R 3 : uD1(R3 ) < ∞ .<br />

R 3


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 493<br />

Whi<strong>le</strong> VMO(R3 ) ⊂ D1(R3 ), one has the following continuous embeddings:<br />

3<br />

D2 R 3<br />

⊂ D1 R ⊂ BMO R 3 .<br />

The first embedding is a consequence of (1.6), and the second of the <strong>du</strong>ality between BMO(R3 )<br />

and the Hardy space H1 (R3 ), and of a <strong>de</strong>composition of every function in H1 (R3 ) as a sum of<br />

some components of curl-free vector-fields.<br />

If u ∈ D1(R2 ), its extension U(x,y) = u(x) to R3 is in D1(R3 ).ItwouldbeinD1(R3 )<br />

if and only if U was boun<strong>de</strong>d. On the other hand, if u ∈ D2(R3 ) is continuous, one has<br />

the trace inequality u| R2D1(R 2 ) CuD2(R3 ) . The prob<strong>le</strong>m whether the trace inequalities<br />

u| R2BMO(R2 ) CuD1(R3 ) and u|RBMO(R) CuD2(R3 ) hold is open.<br />

The <strong>semi</strong>norm ·D1(R3 ) can also be characterized geometrically: by the co-area formu<strong>la</strong>, for<br />

every u ∈ C(R3 ),<br />

uD1(R3 1<br />

) = sup<br />

Ω H2 <br />

<br />

<br />

(∂Ω) u(y)ν(y)dH 2 <br />

<br />

(y) <br />

,<br />

where the supremum is taken over boun<strong>de</strong>d domains Ω ⊂ R 3 with a smooth connected boundary,<br />

ν(y) is the unit exterior normal vector to the boundary at y ∈ ∂Ω, and H 2 is the two-dimensional<br />

Hausdorff mea<strong>sur</strong>e.<br />

1.3. Integrals along differential forms<br />

In higher dimensions, the generalization of (1.1) corresponding to (1.5) in R3 is expressed<br />

with differential forms: if 1 k n − 1, then, for every compactly supported smooth<br />

k-differential form ϕ ∈ D(Rn ; ΛkRn ) and for every u ∈ Ws,p (Rn ) with p 1 and sp = n, if<br />

dϕ = 0, then<br />

<br />

<br />

<br />

<br />

<br />

<br />

uϕ dx<br />

Cs,pϕL1 (Rn ) uWs,p (Rn ). (1.7)<br />

R n<br />

The previous <strong>de</strong>finitions of Dk(Rn ) are generalized as follows. For 1 k n − 1, we <strong>de</strong>fine<br />

the <strong>semi</strong>norm<br />

<br />

<br />

<br />

uϕ dx<br />

<br />

and the vector space<br />

∂Ω<br />

uDk(R n ) = sup<br />

ϕ∈D(R n ;Λ k R n )<br />

dϕ=0<br />

ϕ L 1 (R n ) 1<br />

R n<br />

n<br />

Dk R = u ∈ D ′ R n : uDk(Rn ) < ∞ .<br />

By (1.7), Ws,p (Rn ) ⊂ Dk(Rn ). These spaces Dk(Rn ) also contain other functions, such as<br />

log( k+1 i=1 x2 i ).<br />

The spaces Dk(Rn ) contain neither BMO(Rn ) nor VMO(Rn ). Our main result is that Dk(Rn )<br />

is embed<strong>de</strong>d in BMO(Rn ). We first show that Dk(Rn ) is embed<strong>de</strong>d in D1(Rn ), then we prove


494 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

that D1(R n ) is embed<strong>de</strong>d in BMO(R n ), in the same fashion as the embedding of D1(R 3 ) in<br />

BMO(R 3 ) outlined above.<br />

The other known properties of the spaces Dk(R n ) can be summarized as follows. The <strong>semi</strong>norm<br />

·Dk(R n ) is a norm mo<strong>du</strong>lo constants and the space Dk(R n ) mo<strong>du</strong>lo constants is a Banach<br />

space. The spaces Dk(R n ) are all different and are <strong>de</strong>creasing with respect to k. The spaces<br />

Dk(R n ) also have a trace property. If u ∈ DK(R N ) is a limit of continuous functions, then it<br />

has a well-<strong>de</strong>fined trace in Dk(R n ) if N − K = n − k. On the other hand, the function spaces<br />

Dk(R n ) do not have better integrability properties then the exponential integrability of functions<br />

in BMO(R n ).<br />

1.4. Organization of the paper<br />

We <strong>de</strong>fine in Section 2 the spaces Dk(R n ) for 1 k n by <strong>du</strong>ality on closed smooth forms,<br />

and characterize them by <strong>du</strong>ality on exact forms (Proposition 2.6). The space Dn(R n ) is in fact<br />

L ∞ (R n )/R (Proposition 2.9). We characterize geometrically the <strong>semi</strong>norm for continuous functions<br />

in the cases k = 1 (Proposition 2.10) and k = n − 1 (Proposition 2.11). For 1


2.2. Definition<br />

J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 495<br />

The spaces Dk(R n ) are <strong>de</strong>fined in terms of appropriate test function spaces.<br />

Definition 2.1. For 1 k n, <strong>de</strong>fine<br />

n k n<br />

D# R ; Λ R <br />

= ϕ ∈ D R n ; Λ k R n <br />

: dϕ = 0 and<br />

L 1 n k n<br />

# R ; Λ R <br />

= ϕ ∈ L 1 R n ; Λ k R n : dϕ = 0 and<br />

Remark 2.2. The restriction 1 k n, is justified by the fact that<br />

when k = 0ork>n.<br />

L 1 n k n<br />

# R ; Λ R n k n<br />

= D# R ; Λ R ={0}<br />

R n<br />

<br />

R n<br />

<br />

ϕdx= 0 ,<br />

<br />

ϕdx= 0 .<br />

Remark 2.3. If 1 k n − 1, ϕ ∈ L 1 (R n ; Λ k R n ) and dϕ = 0, then <br />

R n ϕ = 0, whi<strong>le</strong> for every<br />

ϕ ∈ L 1 (R n ; Λ n R n ), dϕ = 0. Therefore, for a given 1 k 1, only one condition in the <strong>de</strong>finition<br />

is essential.<br />

Definition 2.4. For 1 k n, and u ∈ D ′ (R n ),<strong>le</strong>t<br />

and <strong>de</strong>fine<br />

uDk(R n ) = sup<br />

ϕ∈D#(R n ;Λ k R n )<br />

ϕ L 1 (R n ) 1<br />

<br />

<br />

<br />

uϕ dx<br />

,<br />

R n<br />

n<br />

Dk R = u ∈ D ′ R n : uDk(Rn ) < ∞ .<br />

The integral appearing in the <strong>de</strong>finition of ·Dk(R n ) should be un<strong>de</strong>rstood as a <strong>du</strong>ality pro<strong>du</strong>ct<br />

between D ′ (R n ) and D(R n ; Λ k R), which takes its values in Λ k R n , whi<strong>le</strong> |·|is the standard<br />

Eucli<strong>de</strong>an norm of this integral. The set Dk(R n ) is a vector space; the function ·Dk(R n ) is a<br />

<strong>semi</strong>norm on Dk(R n ), and vanishes for constant distributions.<br />

Remark 2.5. In fact, by Theorem 5.3, if uDk(R n ) = 0, then uBMO(R n ) = 0, whence u is<br />

constant.<br />

The <strong>semi</strong>norm ·Dk(R n ) can also be computed by consi<strong>de</strong>ring exterior differentials of compactly<br />

supported smooth forms in p<strong>la</strong>ce of closed compactly supported smooth forms.<br />

Proposition 2.6. Let 1 k n. For every u ∈ D ′ (Rn ),<br />

<br />

<br />

<br />

udψdx<br />

.<br />

uDk(R n ) = sup<br />

ψ∈D(R n ;Λ k−1 R n )<br />

dψ L 1 (R n ) 1<br />

R n


496 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

Proof. This follows from Theorems A.5 and A.8. ✷<br />

Theorems A.5 and A.8 also allow to extend u by <strong>de</strong>nsity to a linear operator from<br />

L 1 # (Rn ,Λ k R n ) to Λ k R n .<br />

Proposition 2.7. If u ∈ Dk(R n ), then 〈u, ϕ〉∈Λ k R n is well <strong>de</strong>fined for every ϕ ∈ L 1 # (Rn ,Λ k R n ).<br />

We shall also consi<strong>de</strong>r the subspace generated by continuous functions.<br />

Definition 2.8. The space Vk(R n ) is the clo<strong>sur</strong>e of the set of boun<strong>de</strong>d continuous functions in<br />

Dk(R n ).<br />

2.3. Characterization of Dn(R n )<br />

The space Dn(R n ) is well known; it is L ∞ (R n )/R.<br />

Proposition 2.9. The spaces Dn(R n ) and L ∞ (R n )/R are isometrically isomorphic.<br />

Proof. If u ∈ L∞ (Rn ), then for every ϕ ∈ D#(Rn ; ΛnRn ) and λ ∈ R,<br />

<br />

<br />

<br />

uϕ dx<br />

=<br />

<br />

<br />

<br />

<br />

<br />

(u − λ)ϕ dx<br />

u − λL∞ (Rn )ϕL1 (Rn ) .<br />

R<br />

R n<br />

Conversely, if u ∈ Dn(R n ), then<br />

<br />

ϕ ↦→ ℓ(ϕ) =<br />

R n<br />

uϕ dx<br />

is a linear continuous mapping from L1 # (Rn ,ΛnRn ) to ΛnRn ∼ = R. By the Hahn–Banach theorem<br />

there is an extension ¯ℓ to L1 (Rn ,ΛnRn ). This extension is represented as ¯ℓ(ϕ) = <br />

Rn ūϕ dx with<br />

ūL∞ (Rn ) uDn(Rn ). Since <br />

Rn ūϕ dx = <br />

Rn uϕ dx for every ϕ ∈ D#(Rn ,ΛnRn ), there is<br />

λ ∈ R such that u − λ =ū. ✷<br />

2.4. Geometric characterization of V1(R n )<br />

The <strong>de</strong>finition of Dk(R n ), and hence that of Vk(R n ), rely on compactly supported closed<br />

smooth forms (Definition 2.4), or equiva<strong>le</strong>ntly on compactly supported exact forms (Proposition<br />

2.6). Compactly supported smooth forms en<strong>sur</strong>e that the <strong>de</strong>finition makes sense for distributions.<br />

There is a more geometrical characterization for continuous functions, which extends by<br />

<strong>de</strong>nsity to V1(R n ) and Vn−1(R n ).<br />

Proposition 2.10. For every u ∈ C(R n ),<br />

uD1(R n ) = sup<br />

Ω<br />

1<br />

Hn−1 <br />

<br />

<br />

(∂Ω) <br />

∂Ω<br />

u(y)ν(y)dH n−1 <br />

<br />

(y) <br />

, (2.1)


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 497<br />

where the supremum is taken over boun<strong>de</strong>d domains Ω with a smooth connected boundary, and<br />

ν(y) is the unit exterior normal vector to the boundary at y ∈ ∂Ω.<br />

Proof. Let Ω be a boun<strong>de</strong>d domains with a smooth connected boundary. Let ρ ∈ D(B(0, 1)) be<br />

such that ρ 0 and <br />

R n ρdx= 1, <strong>le</strong>t ρε(x) = ρ(x/ε)/ε n , and <strong>de</strong>fine<br />

Since ϕε1 1 and dϕε = 0,<br />

1<br />

ϕε(x) =<br />

Hn−1 <br />

(∂Ω)<br />

R n<br />

∂Ω<br />

ρε(y − x)ν(y)dH n−1 (y).<br />

<br />

<br />

<br />

uϕε dx<br />

uD1(Rn ).<br />

Since u is continuous, ρε ∗ u → u as ε → 0 uniformly on every compact subset of Rn , and<br />

<br />

<br />

<br />

<br />

<br />

<br />

uϕε dx<br />

=<br />

1<br />

Hn−1 <br />

<br />

<br />

(∂Ω) (ρε ∗ u)(y)ν(y) dH n−1 <br />

<br />

(y) <br />

<br />

as ε → 0. Therefore<br />

R n<br />

∂Ω<br />

1<br />

→<br />

Hn−1 <br />

<br />

<br />

(∂Ω) u(y)ν(y)dH n−1 <br />

<br />

(y) <br />

<br />

1<br />

Hn−1 <br />

<br />

<br />

(∂Ω) <br />

∂Ω<br />

∂Ω<br />

u(y)ν(y)dH n−1 <br />

<br />

(y) <br />

uD1(Rn ).<br />

Conversely, <strong>le</strong>t A <strong>de</strong>note the right-hand si<strong>de</strong> of (2.1). First note that for every boun<strong>de</strong>d open<br />

set Ω with a smooth boundary that is not necessarily connected,<br />

<br />

<br />

<br />

u(y)ν(y)dH n−1 <br />

<br />

(y) <br />

AHn−1 (∂Ω).<br />

∂Ω<br />

By Proposition 2.6, we need to evaluate, for ϕ ∈ D(Rn ,Λ0R), <br />

<br />

<br />

<br />

<br />

<br />

u∇ϕdx<br />

=<br />

<br />

<br />

<br />

<br />

<br />

<br />

ϕ∇udx<br />

.<br />

One has<br />

<br />

R n<br />

<br />

ϕ∇udx =<br />

∞<br />

<br />

R n<br />

0 {x∈Rn : ϕ(x)>s}<br />

R n<br />

∇u(x)dxds −<br />

0<br />

<br />

−∞ {x∈Rn : ϕ(x)


498 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

For every s>0, the set {x ∈ Rn : ϕ(x) > s} is open and boun<strong>de</strong>d. Moreover, by Sard’s <strong>le</strong>mma,<br />

for almost every s>0, for every y ∈ ϕ−1 ({s}), ∇ϕ(y) = 0. Hence ∂{x ∈ Rn : ϕ>s} is smooth<br />

and<br />

<br />

<br />

<br />

<br />

<br />

∇u(x) dx<br />

=<br />

<br />

<br />

<br />

<br />

u(y)ν(y)dH n−1 <br />

<br />

(y) <br />

<br />

{x∈R n : ϕ>s}<br />

∂{x∈R n : ϕ(x)>s}<br />

AH n−1 ∂ x ∈ R n : ϕ(x) > s .<br />

A simi<strong>la</strong>r reasoning for s s + H n−1 ∂ x ∈ R n : ϕ(x) < −s ds<br />

<br />

= A |∇ϕ| dx. ✷<br />

2.5. Geometric characterization of Vn−1(R n )<br />

Proposition 2.11. If n 2, for every u ∈ C(R n ),<br />

uDn−1(R n ) = sup<br />

γ ∈C 1 (S 1 ;R n )<br />

<br />

1 <br />

<br />

˙γ u<br />

L1 (S1 )<br />

γ(t) <br />

<br />

˙γ(t)dt<br />

.<br />

The proof repeats the argument of Bourgain and Brezis for the equiva<strong>le</strong>nce between the inequality<br />

(1.7) and<br />

<br />

u γ(t) ˙γ(t)dtCs,p˙γ L1 (S1 ) uWs,p (Rn ),<br />

for every u ∈ W s,p (R n ) with sp = n [3].<br />

Proof. First note that<br />

S 1<br />

uDn−1(R n ) = sup<br />

f ∈D(R n ;R n )<br />

div f =0<br />

f L 1 (R n ;R n ) 1<br />

S 1<br />

<br />

<br />

<br />

uf d x<br />

.<br />

Let ρ ∈ D(B(0, 1)) be such that ρ 0 and <br />

R n ρdx= 1, and <strong>le</strong>t ρε(x) = ρ(x/ε)/ε n . Define<br />

R n<br />

<br />

1<br />

fε(x) =<br />

ρ<br />

˙γ L1 (S1 )<br />

γ(t)−x ˙γ(t)dt.<br />

S 1


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 499<br />

One has fεL1 (Rn ) 1 and div fε = 0, therefore<br />

<br />

<br />

<br />

ufε dx<br />

uDn−1(Rn ).<br />

R n<br />

The proof continues as for the first part of Proposition 2.10.<br />

The converse inequality comes from on a result of Smirnov [18], which states that for<br />

every R>0 and for every f ∈ D(B(0,R); Rn ) there exists (γ ℓ m )1m,ℓ in C1 (S1 ; B(0,R)) and<br />

(λℓ m )1m,ℓ in R such that for every m 1,<br />

<br />

λ ℓ <br />

<br />

m<br />

˙γ ℓ <br />

<br />

m L1 (S1 f ) L1 (Rn ) ,<br />

and for every u ∈ C(B(0,R))<br />

as m →∞. ✷<br />

ℓ1<br />

∞<br />

ℓ=1<br />

λ ℓ m<br />

2.6. Geometric characterization of Vk(R n )<br />

<br />

S 1<br />

u γ ℓ m (t) ˙γ ℓ <br />

m dt →<br />

R n<br />

uf d x<br />

The characterization of the <strong>semi</strong>norm ·Dk(R n ) of Proposition 2.10, relies essentially on the<br />

fact that the <strong>semi</strong>norm could be evaluated by consi<strong>de</strong>ring differential of sca<strong>la</strong>r functions, whi<strong>le</strong><br />

in Proposition 2.11 it relied on the <strong>de</strong>composition result of Smirnov. Those facts do not hold<br />

anymore for 1


500 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

For every finite col<strong>le</strong>ction (xi)1ik in R n , the current ❏x0,...,xk❑ ∈ Dk(R n ) is <strong>de</strong>fined by<br />

<br />

❏x0,...,xk❑,ϕ <br />

=<br />

Sk<br />

<br />

ϕ<br />

k<br />

x0 + λixi<br />

i=1<br />

<br />

<br />

,x1 − x0 ∧···∧xk − x0 dλ.<br />

Definition 2.13. A current T ∈ Dk(R n ) is a real polyhedral chain if there is (x i j )1im,1jk in<br />

R n and (μi)1im in R such that<br />

T =<br />

m<br />

i=1<br />

i<br />

μi x0 ,...,x i ②<br />

k .<br />

The set of k-dimensional real polyhedral chains is <strong>de</strong>noted by Pk(R n ). Every real polyhedral<br />

chain has a compact support and a finite mass. Hence, 〈T,u〉 is well <strong>de</strong>fined when u : R n → Λ k R n<br />

is continuous. Moreover, if u : R n → R is continuous then 〈T,u〉∈(Λ k R n ) ′ ∼ = Λ k R n is naturally<br />

<strong>de</strong>fined.<br />

Definition 2.14. If u : R n → R is continuous, <strong>le</strong>t<br />

u ˜Vk(R n )<br />

= sup<br />

P ∈Pn−k(R n )<br />

∂P=0<br />

M(P )1<br />

〈P,u〉.<br />

The <strong>semi</strong>norm u ˜Vk(R n ) mea<strong>sur</strong>es the oscil<strong>la</strong>tion of the function u through its integral on<br />

k-dimensional real polyhedral chains without boundary.<br />

Theorem 2.15. For every n 1 and 1 k n − 1, there exists c>0 such that for every u ∈<br />

C(R n ),<br />

Proof. Given P in Pn−k(R n ),<strong>le</strong>t<br />

u ˜Vk(R n ) uDk(R n ) cu ˜Vk(R n ) .<br />

ϕε(x) = P,ρε(·−x) ,<br />

where ρε = ρ(·/ε)/ε n with ρ ∈ D(R n ), ρ 0 and <br />

R n ρdx= 1 and where ∗ <strong>de</strong>notes the Hodge<br />

<strong>du</strong>ality between Λ k R n and Λ n−k R n . One checks that dϕε = 0, ϕε L 1 (R n ) M(P ) and<br />

<br />

R n<br />

Since ρε ∗ u → u uniformly as ε → 0,<br />

uϕε dx =∗〈P,ρε ∗ u〉.<br />

uVk(R n ) uDk(R n ).


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 501<br />

The converse inequality is based on the <strong>de</strong>formation Theorem for currents [10,17]. It states<br />

that given T ∈ Dk(R n ) with M(T ) + M(∂T ) < ∞, for every ε>0, there exists P ∈ Pk(R n ),<br />

S ∈ Dk(R n ) and R ∈ Dk+1(R n ) such that<br />

with<br />

and<br />

T − P = ∂R + S,<br />

M(P ) cM(T ), M(∂P ) cM(∂T ),<br />

M(R) cεM(T ), M(S) cεM(∂T ),<br />

supp P ∪ supp R ⊂ x ∈ R n :dist(x, supp T)0 in Pk(R n ) with ∂Pε = 0 and supp Pε ⊂ U such that for every<br />

u ∈ C(R n ; R n ),<br />

and<br />

M(Pε) cM(T ), (2.2)<br />

〈Pε,u〉→〈T,u〉 (2.3)<br />

as ε → 0.<br />

Now given ϕ ∈ D#(Rn ; ΛkRn ), consi<strong>de</strong>r T ∈ Dn−k(Rn ) <strong>de</strong>fined by<br />

<br />

〈T,v〉= ϕ ∧ vdx.<br />

R n<br />

Since T has compact support, ∂T = 0 and M(T ) ϕ L 1 (R n ) , there is a sequence (Pε)ε>0 in<br />

Pn−k(R n ) such that ∂Pε = 0, supp Pε ⊂ U, (2.2) and (2.3), where U is a fixed open boun<strong>de</strong>d set<br />

such that supp T ⊂ U. ✷<br />

3. Basic properties of Dk(R n )<br />

3.1. Mutual injections<br />

The col<strong>le</strong>ction of spaces Dk(R n ) is a <strong>de</strong>creasing sequence of spaces.<br />

Theorem 3.1. Let k ℓ.Ifu ∈ Dℓ(R n ), then u ∈ Dk(R n ), and<br />

where C does not <strong>de</strong>pend on u.<br />

uDk(R n ) CuDℓ(R n ),


502 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

Proof. Let ϕ ∈ D#(Rn ; ΛkRn ).Ifα ∈ Λℓ−kRn , then α ∧ ϕ ∈ D#(Rn ; ΛℓRn ). Therefore<br />

<br />

<br />

<br />

<br />

uα∧ ϕdx<br />

uDℓ(Rn )α ∧ ϕL1 (Rn ) uDℓ(Rn )|α|ϕL1 (Rn ) .<br />

R n<br />

Taking the supremum over α ∈ Λ ℓ−k R n with |α| 1 <strong>le</strong>ads to the conclusion. ✷<br />

3.2. Extension theory<br />

If n 0 are in<strong>de</strong>pen<strong>de</strong>nt of u and U.<br />

R n<br />

R N−n<br />

ϕ(x,y)dy dx.<br />

cU Dk(R N ) uDk(R n ) CU Dk(R N ) ,<br />

Proof. By in<strong>du</strong>ction, it is sufficient to consi<strong>de</strong>r the case N = n + 1.<br />

First <strong>le</strong>t us estimate U Dk(R N ) . Consi<strong>de</strong>r Φ ∈ D#(R N ; Λ k R N ). It can be written as<br />

Φ = Φ0 + Φ1 ∧ ωN,<br />

where Φ0 ∈ D#(R N ; Λ k R n ) and Φ1 ∈ D#(R N ; Λ k−1 R n ). Define<br />

<br />

ϕ(x) =<br />

For m = 1, 2,<br />

and<br />

R<br />

<br />

Φ(x,t)dt, ϕ0(x) =<br />

dϕm(x) =<br />

<br />

R n<br />

n<br />

i=1<br />

R<br />

ωi ∧ ∂ϕm<br />

∂xi<br />

<br />

ϕm dx =<br />

<br />

Φ0(x, t) dt, ϕ1(x) =<br />

<br />

=<br />

<br />

Rn R<br />

R<br />

N<br />

i=1<br />

ωi ∧ ∂Φm<br />

∂xi<br />

Φm dt dx = 0.<br />

R<br />

dt = 0<br />

Φ1(x, t) dt.


Since ϕ = ϕ0 + ϕ1 ∧ ωN, one has<br />

<br />

UΦdz=<br />

and therefore,<br />

<br />

<br />

<br />

<br />

R N<br />

J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 503<br />

R N<br />

R n<br />

<br />

uϕ dx =<br />

R n<br />

<br />

uϕ0 dx +<br />

R n<br />

uϕ1 dx ∧ ωN<br />

<br />

<br />

UΦdz<br />

uDk(Rn )ϕ0L1 (Rn ) +uDk−1(Rn )ϕ1L1 (Rn ) |ωN |.<br />

When k>1, the conclusion comes from Theorem 3.1 and from the inequality<br />

ϕ0 L 1 (R n ) +ϕ1 L 1 (R n ) Φ0 L 1 (R n ) +Φ1 ∧ ωN L 1 (R n ) Φ0 + Φ1 ∧ ωN L 1 (R n ) .<br />

The <strong>la</strong>st inequality comes from the fact that Φ0(x) and Φ1(x) ∧ ωN are orthogonal for every<br />

x ∈ RN . When k = 1, one has ϕ1 = 0, and the conclusion comes simi<strong>la</strong>rly.<br />

Conversely, <strong>le</strong>t us now estimate uDk(Rn ) by Proposition 2.6. Let ψ ∈ D(Rn ; Λk−1Rn ). Consi<strong>de</strong>r<br />

a family (ηλ)λ>0 in D(R) such that ηλ 0, <br />

R ηλ dt = 1 and <br />

R |η′ λ | dt → 0asλ→∞, and <strong>le</strong>t Ψλ(x, t) = ηλ(t)ψ(x). For every λ>0,<br />

<br />

R N<br />

Therefore,<br />

<br />

<br />

<br />

udψdx<br />

=<br />

<br />

<br />

<br />

<br />

R n<br />

<br />

UdΨλdz =<br />

R N<br />

R N<br />

Letting λ →∞yields the conclusion. ✷<br />

<br />

U(dηλ∧ ψ + ηλdψ)dz = 0 +<br />

R n<br />

udψ dx.<br />

<br />

<br />

UdΨλdz UDk(R N <br />

) ψL1 (Rn ) +dψL1 (Rn ) η ′ λL1 <br />

(R) .<br />

Remark 3.3. When kk(see Proposition 4.6). On the<br />

other hand, the extension of a function in Dn(R n ) lies in DN(R N ) by Proposition 2.9. In view of<br />

the trace theory of the next section, one could won<strong>de</strong>r whether when 1 kk.<br />

3.3. Trace theory<br />

The restriction of continuous functions from R N to R n can be exten<strong>de</strong>d to a continuous operator<br />

from VK(R n ) to Vk(R n ) when N − K = n − k.<br />

Theorem 3.4. Let n N, 1 K N and k = K − (N − n). Let U ∈ Vk(R N ) be continuous.<br />

Define for x ∈ R n ,<br />

u(x) = U(x,0).


504 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

Then u ∈ Vk(R n ), and<br />

uDk(R n ) U DK(R N ) .<br />

Proof. By in<strong>du</strong>ction, we can assume N = n + 1. Let ϕ ∈ D#(Rn ; ΛkRn ),<strong>le</strong>tρ∈D(R) such<br />

that ρ 0 and <br />

R ρdt = 1 and <strong>le</strong>t ρε(t) = ρ(t/ε)/ε.LetΦε(x, t) = ρε(t)ψ(x) ∧ ωN . Since<br />

Φε ∈ D#(RN ; ΛKRN ) and u is continuous,<br />

<br />

<br />

<br />

<br />

<br />

<br />

uϕ dx<br />

= lim <br />

<br />

<br />

<br />

uΦε dt dx<br />

<br />

R n<br />

ε→0<br />

Rn R<br />

4. Examp<strong>le</strong>s of functions in Dk(R n )<br />

4.1. Sobo<strong>le</strong>v spaces<br />

UDK (RN ) lim<br />

ε→0 ΦεL1 (RN ) UDK (RN ) ϕL1 (Rn ) . ✷<br />

The first c<strong>la</strong>ss of functions in the space Dk(R n ) are functions in critical Sobo<strong>le</strong>v spaces, which<br />

motivated the <strong>de</strong>finition.<br />

Theorem 4.1. (Bourgain and Brezis [3]) If u ∈ W s,p (R n ), p>1 and sp = n, then for every<br />

1 k n − 1, u ∈ Dk(R n ), and<br />

uDk(R n ) Ck,s,puW s,p (R n ).<br />

The <strong>semi</strong>norm on the right-hand si<strong>de</strong> is the Sobo<strong>le</strong>v <strong>semi</strong>-norm. For 0


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 505<br />

4.2. Locally Lipschitz functions in R n \{0}<br />

The space Dn−1(R n ) is <strong>la</strong>rger than critical Sobo<strong>le</strong>v spaces. There is a simp<strong>le</strong> condition for<br />

locally Lipschitz functions in R n \{0} to be in Dn−1(R n ), which is satisfied e.g. by the function<br />

log |x|.<br />

Proposition 4.3. Let n 2 and u ∈ W 1,1<br />

loc (Rn \{0}).If|x|∇u ∈ L ∞ (R n ), then u ∈ Dn−1(R n ) and<br />

uDn−1(R n ) |x|∇u L ∞ (R n ) .<br />

Remark 4.4. In general, u/∈ Dn(R n ) as shows the function u(x) = log |x|.<br />

Proof. Let f ∈ D(Rn \{0}; Rn ) be such that div f = 0. Lemma 4.5 yields<br />

<br />

<br />

<br />

<br />

<br />

<br />

uf d x<br />

=<br />

<br />

<br />

<br />

<br />

<br />

<br />

x(∇u · f)dx<br />

<br />

∇u|x| L<br />

∞ (Rn f ) L1 (Rn ) .<br />

R n<br />

R n<br />

Therefore, for every ϕ ∈ D#(Rn \{0}; Rn ),<br />

<br />

<br />

<br />

uϕ dx<br />

∇u|x| <br />

L∞ (Rn ϕ ) L1 (Rn ) .<br />

Since {0} has vanishing n-capacity (Lemma A.2),<br />

R n<br />

<br />

n n−1 n n−1<br />

D# R \{0}; Λ R is <strong>de</strong>nse in D# R ; Λ R<br />

(Theorem A.5). This conclu<strong>de</strong>s the proof. ✷<br />

Lemma 4.5. Let u ∈ W 1,1<br />

loc (Rn \{0}) and f ∈ D(Rn \{0}; Rn ).Ifdiv f = 0, then<br />

<br />

<br />

uf d x =− x(f ·∇u) dx.<br />

Proof. By integration by parts,<br />

<br />

<br />

x(f ·∇u) dx =−<br />

R n<br />

R n<br />

R n<br />

R n<br />

<br />

x(div f)udx−<br />

The conclusion comes from the assumption div f = 0. ✷<br />

4.3. Examp<strong>le</strong>s of functions in Dk(R n )<br />

R n<br />

uf d x.<br />

Proposition 4.3 and Theorem 3.2 yield examp<strong>le</strong>s of functions in the spaces Dk(R n ) showing<br />

that these spaces are distinct.


506 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

Proposition 4.6. If 1 ℓ n, then<br />

if and only if 1 k


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 507<br />

Corol<strong>la</strong>ry 4.9. (Bourgain and Brezis [3]) Let f ∈ L1 (R2 ; R2 ).Ifdiv f = 0 in the sense of distributions,<br />

then<br />

log 1<br />

|x| ∗ f ∈ L∞R 2 ; R 2 .<br />

Other interesting examp<strong>le</strong>s can be obtained in a simi<strong>la</strong>r way:<br />

Proposition 4.10. If 1 ℓ n and 0


508 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

and<br />

<br />

<br />

<br />

∂<br />

<br />

2 <br />

<br />

(Kn ∗ f) <br />

∂xi∂xj<br />

L 1 (R n )<br />

Cf H 1 (R n ) .<br />

Finally, D(R n ) ∩ H 1 (R n ) is <strong>de</strong>nse in H 1 (R n ).<br />

The space of functions with vanishing mean oscil<strong>la</strong>tions VMO(R n ; R n ) is the closed subspace<br />

of VMO(R n ) that is characterized by<br />

lim<br />

sup<br />

ε→0 Ln (B)ε<br />

1<br />

Ln (B) 2<br />

<br />

B<br />

<br />

B<br />

<br />

u(x) − u(y) dxdy = 0,<br />

where the supremum is taken over balls B ⊂ R n . The critical Sobo<strong>le</strong>v spaces W s,p (R n ) with<br />

sp = n are embed<strong>de</strong>d in VMO(R n ).<br />

Going back to the examp<strong>le</strong>s of the preceding section, for every ℓ 1,<br />

log<br />

ℓ<br />

i=1<br />

|xi| 2<br />

<br />

∈ BMO R n ,<br />

but it does not belong to VMO(Rn ), whi<strong>le</strong> for 0


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 509<br />

Proof. By Theorem 3.1, we can assume k = 1. The <strong>semi</strong>norm of u in BMO(R n ) will be estimated<br />

by <strong>du</strong>ality with the Hardy space H 1 (R n ).<br />

Let f ∈ D(R n ) ∩ H 1 (R n ).Let<br />

ϕi =<br />

Note ϕj ∈ L 1 # (Rn ; Λ 1 R n ). Moreover,<br />

R n<br />

i=1<br />

R n<br />

n ∂2 (KN ∗ f)ωj .<br />

∂xi∂xj<br />

j=1<br />

<br />

<br />

<br />

fudx<br />

<br />

n<br />

<br />

<br />

<br />

∂ 2 f/∂x 2 i udx<br />

<br />

<br />

<br />

<br />

uD1(R n )<br />

n<br />

<br />

<br />

<br />

<br />

i=1<br />

R n<br />

ϕi udx<br />

n<br />

ϕiL1 (Rn ) CuD1(Rn )f H1 (Rn ) . (5.1)<br />

i=1<br />

(Note that Proposition 2.7 about the well-<strong>de</strong>finiteness of the <strong>du</strong>ality pro<strong>du</strong>ct between D1(R n ) and<br />

L 1 # (Rn ; Λ 1 R n ) was used.) ✷<br />

Remark 5.4. A simi<strong>la</strong>r argument shows that for 1


510 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

5.4. Integrability of functions in Dk(Rn )<br />

If u ∈ BMO(Rn ), uBMO(Rn ) 1 and <br />

B(0,1) udx = 0, the John–Nirenberg theorem states<br />

that<br />

<br />

exp μ|u| dx c, (5.2)<br />

B(0,1)<br />

where μ>0 and c>0 can be chosen in<strong>de</strong>pen<strong>de</strong>ntly of u. Since the spaces Dk(Rn ) are embed<strong>de</strong>d<br />

in BMO(Rn ), this might be improved on Dk(Rn ).<br />

On the other hand, if sp = n,0


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 511<br />

The improvement of Dk(R n ) on BMO(R n ) should thus not be seen as an improvement of the<br />

integrand, but as an improvement on the domains of integration: by the trace Theorem 3.4, the<br />

embedding Theorem 5.3, and the John–Nirenberg inequality, functions in Vk(R n ) are exponentially<br />

integrab<strong>le</strong> on (n − k + 1)-dimensional subspaces.<br />

6. Further prob<strong>le</strong>ms<br />

6.1. Traces of V1(R n ) on VMO(R n−1 )<br />

By Theorems 3.4 and 5.3, functions in Vk(R n ) have VMO traces on (n − k + 1)-dimensional<br />

spaces. The dimension n − k seems more natural: functions in Vn(R n ) are continuous, and hence<br />

have traces on 0-dimensional spaces, i.e. points. If there was such a trace inequality, one could<br />

<strong>de</strong>fine D0(R n ) = BMO(R n ). This notation would be consistent with the mutual injection Theorem<br />

3.1, the extension Theorem 3.2 and the examp<strong>le</strong>s of Proposition 4.6. It would then be nice<br />

to have a <strong>de</strong>finition of Dk(R n ) which encompasses the case k = 0. The two-dimensional case<br />

would already solve the prob<strong>le</strong>m of traces of Vn−1(R n ) on lines.<br />

6.2. Geometric characterizations<br />

By Propositions 2.10 and 2.11, the spaces V1(R n ) and Vn−1(R n ) can be <strong>de</strong>fined by oscil<strong>la</strong>tions<br />

respectively along boundaries of boun<strong>de</strong>d domains and along closed curves. Further<br />

refinements would restrict the set of domains and of curves. The most striking result would be if<br />

the oscil<strong>la</strong>tion could be simply evaluated respectively on spheres and on circ<strong>le</strong>s.<br />

The spaces Vk(R n ) for 1


512 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

6.4. Simi<strong>la</strong>r spaces Ek(R n )<br />

It would have been possib<strong>le</strong> to <strong>de</strong>fine another family of spaces with properties simi<strong>la</strong>r to<br />

Dk(Rn ).For1k n,<strong>le</strong>t<br />

n n<br />

R ; R = ϕ ∈ D R n ; R n :divϕ = 0<br />

D#,k<br />

and the dimension of the range of ϕ is at most k .<br />

(The set D#,k(Rn ; Rn ) is not a vector space when k


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 513<br />

embed<strong>de</strong>d respectively in bmoz( ¯Ω) (functions whose extension by 0 to R n is in BMO(R n )) and<br />

the second in the <strong>la</strong>rger space bmor( ¯Ω) (restrictions of functions in BMO(R n ) to Ω) (see [8] for<br />

the <strong>de</strong>finitions).<br />

Acknow<strong>le</strong>dgments<br />

The author thanks Haïm Brezis who suggested the prob<strong>le</strong>m, and who encouraged and discussed<br />

the progress of this work. He also thanks Thierry De Pauw for discussions, and acknow<strong>le</strong>dges<br />

the hospitality of the Laboratoire Jacques-Louis Lions, Université Pierre et Marie Curie in<br />

Paris, at which this work was initiated.<br />

Appendix A. Density of compactly supported forms<br />

A.1. The clo<strong>sur</strong>e of closed k-forms<br />

This appendix is <strong>de</strong>voted to the study of <strong>de</strong>nse sets in the space L 1 (R n ; Λ k R n ) of summab<strong>le</strong><br />

closed forms.<br />

Definition A.1. Let 1 p0, the s-dimensional Hausdorff<br />

mea<strong>sur</strong>e of Σ vanishes [9].<br />

In a simi<strong>la</strong>r way, one can prove<br />

Lemma A.4. There exists a sequence (ζm)m in D(Rn ) such that 0 ζm 1, ζm → 1 almost<br />

everywhere and<br />

<br />

|∇ζm| n dx → 0<br />

as m →∞.<br />

R n


514 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

Theorem A.5. If Σ ⊂ R n is compact and has vanishing n-capacity, then d(D(R n \Σ; Λ k−1 R n ))<br />

is <strong>de</strong>nse in L 1 # (Rn ; Λ k R n ).<br />

The proof makes use of a result of Bourgain and Brezis.<br />

Theorem A.6. (Bourgain and Brezis [3]) Let 1 k n − 1. For every ϕ ∈ L1 # (Rn ; ΛkRn ), there<br />

exists ψ ∈ Ln/(n−1) (Rn ; Λk−1Rn ) such that<br />

<br />

dψ = ϕ,<br />

δψ = 0.<br />

Here δ <strong>de</strong>notes the codifferential, i.e. the adjoint of d with respect to Hodge star. This result<br />

is based on inequality (1.7). When k = 1, the meaning<strong>le</strong>ss condition δψ = 0 is dropped and this<br />

is equiva<strong>le</strong>nt with the Nirenberg–Sobo<strong>le</strong>v embedding.<br />

Proof of Theorem A.5. Since the exterior differential d commutes with trans<strong>la</strong>tions, by c<strong>la</strong>ssical<br />

smoothing arguments, (C ∞ ∩ L 1 # )(Rn ; Λ k R n ) is <strong>de</strong>nse in L 1 # (Rn ; Λ k R n ).<br />

Let ϕ ∈ (C ∞ ∩ L 1 # )(Rn ; Λ k R n ) and <strong>le</strong>t Σ ⊂ Ω ⊂ R n be open and boun<strong>de</strong>d. Since Σ has<br />

vanishing capacity, there is a sequence (ηm)m1 in D(Ω) such that 0 ηm 1, ηm = 1ona<br />

neighborhood of Σ and ∇ηmL n (R n ) → 0asn →∞. Moreover, by Poincaré’s inequality, up to<br />

a subsequence, ηm → 0 almost everywhere.<br />

Consi<strong>de</strong>r now the sequence<br />

ψm = (1 − ηm)ζmψ,<br />

where ζm is given by Lemma A.4. By <strong>de</strong>finition, ψm ∈ D(R n ; Λ k−1 R n ). We c<strong>la</strong>im that dψm → ϕ<br />

in L 1 (R n ; Λ k R n ).<br />

In fact,<br />

By Höl<strong>de</strong>r’s inequality,<br />

dψm =−ζm dηm ∧ ψ + (1 − ηm)dζm ∧ ψ + (1 − ηm)ζm ϕ. (A.1)<br />

−ζm dηm ∧ ψ L 1 (R n ) ζmL ∞ (R n )dηmL n (R n )ψ L n/(n−1) (R n ) .<br />

Since ∇ηmL n (R n ) → 0 and ψ L n/(n−1) (R n ) < ∞, the first term in (A.1) tends to zero. A simi<strong>la</strong>r<br />

reasoning holds for the second term, and the <strong>la</strong>st term converges to ϕ as m →∞by Lebesgue’s<br />

dominated convergence theorem. ✷<br />

Corol<strong>la</strong>ry A.7. The set D#(R n ; Λ k R n ) is <strong>de</strong>nse in L 1 # (Rn ; Λ k R n ).<br />

A.2. The clo<strong>sur</strong>e of exact n-forms<br />

Theorem A.6 fails when k = n, and therefore the proof Theorem A.5 fails in this case, but<br />

there is in fact a stronger result.<br />

Theorem A.8. The set d(D(R n ; Λ n−1 R n )) is <strong>de</strong>nse in D#(R n ; Λ n R n ).


J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516 515<br />

Remark A.9. The <strong>de</strong>nsity is with respect to the usual topology on the space of test functions [16].<br />

Proof of Theorem A.8. Let ϕ ∈ D#(R n ; Λ n R n ). Therefore ϕ = fω1 ∧ ··· ∧ ωn, with f ∈<br />

D(R n ) and <br />

R n fdx= 0. Let (ρε)ε>0 be a sequence of mollifiers. Define gε ∈ D(R n ; R n ) by<br />

Next, note<br />

2<br />

gε(z) =<br />

f L1 (Rn )<br />

2<br />

div gε(z) =<br />

f L1 (Rn )<br />

2<br />

=<br />

f L1 (Rn )<br />

<br />

R n ×R n<br />

<br />

R n ×R n<br />

<br />

R n ×R n<br />

<br />

(x − y)<br />

1<br />

0<br />

0<br />

1<br />

<br />

ρε z − tx − (1 − t)y f+(x)f−(y) dt dx dy.<br />

<br />

(x − y) ·∇ρε z − tx − (1 − t)y f+(x)f−(y) dt dx dy<br />

ρε(z − x) − ρε(z − y) f+(x)f−(y) dx dy<br />

= (ρε ∗ f )(z). (A.2)<br />

Therefore div gε → f in D(R n ) as ε → 0. Letting<br />

one conclu<strong>de</strong>s<br />

as ε → 0. ✷<br />

ψε =<br />

n<br />

g i ε (−1)i+1ω1 ∧···∧ωi ∧···∧ωn,<br />

i=1<br />

dψε = div gε ω1 ∧···∧ωn → fω1 ∧···∧ωn = ϕ<br />

Remark A.10. The construction (A.2) is inspired from the construction of a non-optimal mass<br />

disp<strong>la</strong>cement p<strong>la</strong>n in the Monge–Kantorovich mass disp<strong>la</strong>cement prob<strong>le</strong>m [11].<br />

References<br />

[1] R.A. Adams, Sobo<strong>le</strong>v Spaces, Pure Appl. Math., vol. 65, Aca<strong>de</strong>mic Press, New York, 1975.<br />

[2] F. Bethuel, G. Or<strong>la</strong>ndi, D. Smets, Approximations with vorticity bounds for the Ginzburg–Landau functional, Commun.<br />

Contemp. Math. 6 (5) (2004) 803–832.<br />

[3] J. Bourgain, H. Brezis, New estimates for the Lap<strong>la</strong>cian, the div–curl, and re<strong>la</strong>ted Hodge systems, C. R. Math. Acad.<br />

Sci. Paris 338 (7) (2004) 539–543.<br />

[4] J. Bourgain, H. Brezis, P. Mironescu, H 1/2 maps with values into the circ<strong>le</strong>: Minimal connections, lifting, and the<br />

Ginzburg–Landau equation, Publ. Math. Inst. Hautes Étu<strong>de</strong>s Sci. 99 (2004) 1–115.<br />

[5] H. Brezis, Analyse fonctionnel<strong>le</strong>, Col<strong>le</strong>ct. Math. Appl. Maîtrise, Masson, Paris, 1983.<br />

[6] H. Brezis, L. Nirenberg, Degree theory and BMO. I. Compact manifolds without boundaries, Se<strong>le</strong>cta Math.<br />

(N.S.) 1 (2) (1995) 197–263.<br />

[7] H. Brezis, J. Van Schaftingen, L 1 estimates on domains, in preparation.<br />

[8] D.-C. Chang, G. Dafni, E.M. Stein, Hardy spaces, BMO, and boundary value prob<strong>le</strong>ms for the Lap<strong>la</strong>cian on a<br />

smooth domain in R n , Trans. Amer. Math. Soc. 351 (4) (1999) 1605–1661.


516 J. Van Schaftingen / Journal of Functional Analysis 236 (2006) 490–516<br />

[9] L.C. Evans, R.F. Gariepy, Mea<strong>sur</strong>e theory and fine properties of functions, Stud. Adv. Math., CRC Press, Boca<br />

Raton, FL, 1992.<br />

[10] H. Fe<strong>de</strong>rer, Geometric Mea<strong>sur</strong>e Theory, Springer-Ver<strong>la</strong>g, New York, 1969.<br />

[11] L. Kantorovitch, G. Akilov, Analyse fonctionnel<strong>le</strong>. Tome 1. Opérateurs et fonctionnel<strong>le</strong>s linéaires, Mir, Moscow,<br />

1981.<br />

[12] C.J. Neugebauer, Lipschitz spaces and exponentially integrab<strong>le</strong> functions, Indiana Univ. Math. J. 23 (1973–1974)<br />

103–106.<br />

[13] H.M. Reimann, Functions of boun<strong>de</strong>d mean oscil<strong>la</strong>tion and quasiconformal mappings, Comment. Math. Helv. 49<br />

(1974) 260–276.<br />

[14] T. Runst, W. Sickel, Sobo<strong>le</strong>v spaces of Fractional Or<strong>de</strong>r, Nemytskij Operators, and Nonlinear Partial Differential<br />

Equations, <strong>de</strong> Gruyter Ser. Nonlinear Anal. Appl., vol. 3, <strong>de</strong> Gruyter, Berlin, 1996.<br />

[15] D. Sarason, Functions of vanishing mean oscil<strong>la</strong>tion, Trans. Amer. Math. Soc. 207 (1975) 391–405.<br />

[16] L. Schwartz, Théorie <strong>de</strong>s distributions, vols. IX–X, Publ. Inst. Math. Univ. <strong>de</strong> Strasbourg, Hermann, Paris, 1966.<br />

[17] L. Simon, Lectures on Geometric Mea<strong>sur</strong>e Theory, Proc. Centre Math. Anal. Austral. Nat. Univ., vol. 3, Australian<br />

National University, Centre for Mathematical Analysis, Canberra, 1983.<br />

[18] S.K. Smirnov, Decomposition of so<strong>le</strong>noidal vector charges into e<strong>le</strong>mentary so<strong>le</strong>noids, and the structure of normal<br />

one-dimensional flows, Algebra i Analiz 5 (4) (1993) 206–238.<br />

[19] E.M. Stein, Singu<strong>la</strong>r Integrals and Differentiability Properties of Functions, Princeton Math. Ser., vol. 30, Princeton<br />

University Press, Princeton, NJ, 1970.<br />

[20] E.M. Stein, Harmonic Analysis: Real-Variab<strong>le</strong> Methods, Orthogonality, and Oscil<strong>la</strong>tory Integrals, Princeton Math.<br />

Ser., vol. 43, Princeton University Press, Princeton, NJ, 1993.<br />

[21] H. Triebel, Interpo<strong>la</strong>tion Theory, Function Spaces, Differential Operators, North-Hol<strong>la</strong>nd Math. Library, vol. 18,<br />

North-Hol<strong>la</strong>nd, Amsterdam, 1978.<br />

[22] J. Van Schaftingen, A simp<strong>le</strong> proof of an inequality of Bourgain, Brezis and Mironescu, C. R. Math. Acad. Sci.<br />

Paris 338 (1) (2004) 23–26.<br />

[23] J. Van Schaftingen, Estimates for L 1 -vector fields, C. R. Math. Acad. Sci. Paris 339 (3) (2004) 181–186.<br />

[24] W.P. Ziemer, Weakly Differentiab<strong>le</strong> Functions: Sobo<strong>le</strong>v Spaces and Functions of Boun<strong>de</strong>d Variation, Grad. Texts in<br />

Math., vol. 120, Springer-Ver<strong>la</strong>g, New York, 1989.


Journal of Functional Analysis 236 (2006) 517–545<br />

A characteristic operator function<br />

for the c<strong>la</strong>ss of n-hypercontractions ✩<br />

An<strong>de</strong>rs Olofsson<br />

Falugatan 22 1tr, SE-113 32 Stockholm, Swe<strong>de</strong>n<br />

Received 5 December 2005; accepted 10 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 18 April 2006<br />

Communicated by G. Pisier<br />

www.elsevier.com/locate/jfa<br />

Abstract<br />

We consi<strong>de</strong>r a c<strong>la</strong>ss of boun<strong>de</strong>d linear operators on Hilbert space cal<strong>le</strong>d n-hypercontractions which re<strong>la</strong>tes<br />

naturally to adjoint shift operators on certain vector-valued standard weighted Bergman spaces on the<br />

unit disc. In the context of n-hypercontractions in the c<strong>la</strong>ss C0· we intro<strong>du</strong>ce a counterpart to the so-cal<strong>le</strong>d<br />

characteristic operator function for a contraction operator. This generalized characteristic operator function<br />

Wn,T is an operator-valued analytic function in the unit disc whose values are operators between two<br />

Hilbert spaces of <strong>de</strong>fect type. Using an operator-valued function of the form Wn,T , we parametrize the wan<strong>de</strong>ring<br />

subspace for a general shift invariant subspace of the corresponding vector-valued standard weighted<br />

Bergman space. The operator-valued analytic function Wn,T is shown to act as a contractive multiplier from<br />

the Hardy space into the associated standard weighted Bergman space.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Characteristic operator function; n-Hypercontraction; Wan<strong>de</strong>ring subspace; Standard weighted Bergman<br />

space; Repro<strong>du</strong>cing kernel function<br />

0. Intro<strong>du</strong>ction<br />

Let us first <strong>de</strong>scribe a c<strong>la</strong>ss of vector-valued standard weighted Bergman spaces that will p<strong>la</strong>y<br />

an important ro<strong>le</strong> in this paper. Let n 1 be an integer and <strong>le</strong>t E be a general not necessarily<br />

✩ Research supported by the M.E.N.R.T. (France) and the G.S. Magnuson’s Fund of the Royal Swedish Aca<strong>de</strong>my of<br />

Sciences.<br />

E-mail address: ao@math.kth.se.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.004


518 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

separab<strong>le</strong> Hilbert space. We <strong>de</strong>note by An(E) the Hilbert space of all E-valued analytic functions<br />

f(z)= <br />

akz k , z∈D, (0.1)<br />

in the unit disc D with finite norm<br />

k0<br />

f 2 <br />

An =<br />

k0<br />

ak 2 μn;k,<br />

where μn;k = 1/ k+n−1 k for k 0. Here the Taylor coefficients ak in (0.1) are e<strong>le</strong>ments in E.<br />

The weight sequence {μn;k}k0 is naturally i<strong>de</strong>ntified as a sequence of moments of a certain<br />

radial mea<strong>sur</strong>e dμn on the closed unit disc in the sense that<br />

<br />

μn;k = |z| 2k <br />

dμn(z) = 1<br />

<br />

k + n − 1<br />

, k0. k<br />

For n 2 the mea<strong>sur</strong>e dμn is given by<br />

¯D<br />

dμn(z) = (n − 1) 1 −|z| 2 n−2 dA(z), z ∈ D,<br />

where dμ2(z) = dA(z) = dxdy/π, z = x + iy, is the usual p<strong>la</strong>nar Lebesgue area mea<strong>sur</strong>e normalized<br />

so that the unit disc D is of unit area. The mea<strong>sur</strong>e dμ1 is the normalized Lebesgue arc<br />

<strong>le</strong>ngth mea<strong>sur</strong>e on the unit circ<strong>le</strong> T = ∂D. The norm of An(E) can also be expressed as<br />

f 2 An<br />

<br />

= lim<br />

r→1<br />

¯D<br />

<br />

f(rz) 2 dμn(z), f ∈ An(E).<br />

The shift operator Sn on the space An(E) is <strong>de</strong>fined by<br />

(Snf )(z) = zf (z) = <br />

ak−1z k , z∈D, (0.2)<br />

k1<br />

for f ∈ An(E) given by (0.1). It is easy to see that the shift operator Sn is boun<strong>de</strong>d on An(E) of<br />

norm equal to 1 (the weight sequence {μn;k}k0 is <strong>de</strong>creasing and the ratio μn;k+1/μn;k tends<br />

to 1 as k →∞). The adjoint operator S∗ n of Sn has the form<br />

∗<br />

Snf (z) = μn;k+1<br />

ak+1z k , z∈D, (0.3)<br />

k0<br />

μn;k<br />

where the function f ∈ An(E) is given by (0.1).<br />

Let I be a shift invariant subspace of An(E). By this we mean that I is a closed subspace of<br />

An(E) which is invariant un<strong>de</strong>r the shift operator Sn in the sense that Sn(I) ⊂ I. The wan<strong>de</strong>ring<br />

subspace EI for I is the subspace<br />

EI = I ⊖ Sn(I)


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 519<br />

of I. The subspace EI has the property that EI ⊥ S k n (EI) for k 1, which is often used as<br />

the <strong>de</strong>fining property for a wan<strong>de</strong>ring subspace. The notion of a wan<strong>de</strong>ring subspace is often<br />

accredited to Halmos [12] and was used as an important concept in his <strong>de</strong>scription of the shift<br />

invariant subspaces of the Hardy space A1(E) using operator-valued inner functions.<br />

In the genuine Bergman space case n 2 it is known that the wan<strong>de</strong>ring subspace EI for a<br />

shift invariant subspace I of An(E) can have dimension equal to any positive integer or +∞ even<br />

in the case E = C of sca<strong>la</strong>r-valued functions. This was first proved by Apostol et al. [5] using<br />

<strong>du</strong>al algebras, and <strong>la</strong>ter more explicit constructions have been found by He<strong>de</strong>nmalm et al. [19]<br />

and others.<br />

In <strong>la</strong>ter <strong>de</strong>velopments the notion of a wan<strong>de</strong>ring subspace has proved to be a useful concept<br />

to study shift invariant subspaces in a Bergman space context. In the sca<strong>la</strong>r case of invariant<br />

subspaces generated by zero sets He<strong>de</strong>nmalm [13–15] has shown that functions in the wan<strong>de</strong>ring<br />

subspace also cal<strong>le</strong>d Bergman inner functions can be used to divi<strong>de</strong> out zeroes of functions in the<br />

subspace for n = 2, 3. For n = 1 we are in the Hardy space context, and for n 4 such a theorem<br />

fails (see [18]).<br />

A re<strong>la</strong>ted question which has attracted much attention is to what extent the wan<strong>de</strong>ring subspace<br />

EI generates the who<strong>le</strong> invariant subspace I in the sense that<br />

I =[EI]= <br />

S k n (EI); (0.4)<br />

k0<br />

we use [F] to <strong>de</strong>note the smal<strong>le</strong>st (closed) shift invariant subspace containing the set F. In our<br />

context of the Bergman spaces An(E) the approximation re<strong>la</strong>tion (0.4) is known to hold true for a<br />

general shift invariant subspace I of An(E) for the values n = 1, 2, 3, and is most probably false<br />

in general for n 4 (see [17,18]). The case n = 1 here is the Hardy space case mentioned earlier<br />

where a parametrization of the shift invariant subspaces is avai<strong>la</strong>b<strong>le</strong>. In the case of an unweighted<br />

Bergman space (n = 2) the approximation re<strong>la</strong>tion (0.4) for a general shift invariant subspace was<br />

first established by A<strong>le</strong>man et al. [3] using function theoretic properties of the biharmonic Green<br />

function for the unit disc; see also [16, Section 3.6] and [20]. Later Shimorin [24] found a more<br />

general result which applies to a more general c<strong>la</strong>ss of pure operators satisfying a certain operator<br />

inequality satisfied by the Bergman shift operator S2. The case n = 3 is <strong>du</strong>e to Shimorin [25]. In<br />

the case n = 2 some summability results stronger than (0.4) are known to hold true (see [21]).<br />

Despite all the <strong>de</strong>velopments indicated above there are few explicit examp<strong>le</strong>s known of<br />

Bergman inner functions or, what is the same, wan<strong>de</strong>ring subspaces in the Bergman spaces.<br />

It is the purpose of the present paper to give a parametrization of the wan<strong>de</strong>ring subspace for a<br />

general shift invariant subspace in the context of the vector-valued standard weighted Bergman<br />

spaces An(E) <strong>de</strong>scribed above. This parametrization is done in terms of certain operator theoretic<br />

quantities known as <strong>de</strong>fect spaces, and some explicit formu<strong>la</strong>s are obtained in the process. Let us<br />

now proceed to <strong>de</strong>scribe the content of the present paper.<br />

By a Hilbert space we mean a general not necessarily separab<strong>le</strong> comp<strong>le</strong>x Hilbert space. We<br />

<strong>de</strong>note by L(H) the space of all boun<strong>de</strong>d linear operators on a Hilbert space H. Letn 1bean<br />

integer. An operator T ∈ L(H) is cal<strong>le</strong>d an n-hypercontraction if the operator inequality<br />

m<br />

k=0<br />

(−1) k<br />

<br />

m<br />

T<br />

k<br />

∗k T k 0 inL(H)


520 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

holds for every 1 m n. In this terminology a 1-hypercontraction is a contraction, and for<br />

n 2 the c<strong>la</strong>ss of n-hypercontractions <strong>de</strong>fines a more restricted c<strong>la</strong>ss of operators.<br />

The c<strong>la</strong>ss of n-hypercontractions was first intro<strong>du</strong>ced by Ag<strong>le</strong>r [1,2]. A principal result<br />

of [2] concerns the <strong>de</strong>scription of a general n-hypercontraction. An operator T ∈ L(H) is an<br />

n-hypercontraction if and only if it is unitarily equiva<strong>le</strong>nt to the restriction to an invariant subspace<br />

of an operator of the form S∗ n ⊕ U, where Sn is the shift operator on a Bergman space<br />

An(E) and U is an isometry. A special case of this result is the well-known fact that an operator<br />

T ∈ L(H) is a contraction if and only if it is part of an operator of the form S∗ 1 ⊕ U, where S1<br />

is the shift operator on a Hardy space A1(E) and U is an isometry, which is often accredited to<br />

Rota, <strong>de</strong> Branges, Rovnyak, Sz.-Nagy and Foias (see [26, Section I.10.1]).<br />

Let us recall that an operator T ∈ L(H) is said to belong to the c<strong>la</strong>ss C0· if limk→∞ T k = 0<br />

in the strong operator topology meaning that limk→∞ T kx = 0inH for every x ∈ H (see [26,<br />

Section II.4]). For operators from the c<strong>la</strong>ss C0· the isometric term U in the <strong>de</strong>scription in the<br />

previous paragraph vanishes and one has that an operator T ∈ L(H) is an n-hypercontraction<br />

such that limk→∞ T k = 0 in the strong operator topology if and only if it is a restriction of the<br />

adjoint shift operator S∗ n to an invariant subspace.<br />

We shall need some more notations re<strong>la</strong>ted to an n-hypercontraction T ∈ L(H). We consi<strong>de</strong>r<br />

the <strong>de</strong>fect operators<br />

Dm,T =<br />

m<br />

k=0<br />

(−1) k<br />

<br />

m<br />

T<br />

k<br />

∗k T k<br />

1/2 in L(H)<br />

for 1 m n, where the positive square root is used. The <strong>de</strong>fect space Dm,T is <strong>de</strong>fined as<br />

the clo<strong>sur</strong>e in H of the range of the operator Dm,T , that is, Dm,T = Dm,T (H). Forn = 1 and<br />

T ∈ L(H) a contraction operator we write also<br />

DT = D1,T = (I − T ∗ T) 1/2<br />

in L(H)<br />

and DT = D1,T = DT (H) for the <strong>de</strong>fect operator and the associated <strong>de</strong>fect space.<br />

In recent work [22] we have revisited the operator mo<strong>de</strong>l theory for the c<strong>la</strong>ss of n-<br />

hypercontractions. It turns out that there is a canonical way to mo<strong>de</strong>l an n-hypercontraction<br />

T ∈ L(H) as part of an operator of the form S∗ n ⊕ U, where U is an isometry. For x ∈ H we<br />

consi<strong>de</strong>r the Dn,T -valued analytic function Vnx <strong>de</strong>fined by the formu<strong>la</strong><br />

(Vnx)(z) = Dn,T (I − zT ) −n x = <br />

<br />

k + n − 1 Dn,T T<br />

k<br />

k x z k , z∈D. (0.5)<br />

k0<br />

It turns out that if T ∈ L(H) is an n-hypercontraction such that limk→∞ T k = 0 in the strong<br />

operator topology, then the map Vn : x ↦→ Vnx given by (0.5) is an isometry<br />

Vn : H → An(Dn,T )<br />

of H into An(Dn,T ) satisfying the intertwining re<strong>la</strong>tion<br />

VnT = S ∗ n Vn.


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 521<br />

In this way an n-hypercontraction T ∈ L(H) in the c<strong>la</strong>ss C0· is naturally mo<strong>de</strong><strong>le</strong>d as part of the<br />

adjoint shift operator S ∗ n on the Bergman space An(Dn,T ). Full <strong>de</strong>tails of this construction can<br />

be found in [22, Sections 6 and 7].<br />

We mention that construction of operator mo<strong>de</strong>ls of this type is a topic of current interest in<br />

multi-variab<strong>le</strong> operator theory with recent contributions by Ambrozie et al. [4] and Arazy and<br />

Engliš [6]. Operator mo<strong>de</strong>ls of this type also form an integral part in recent work on constrained<br />

von Neumann inequalities by Ba<strong>de</strong>a and Cassier [8].<br />

In this paper we shall consi<strong>de</strong>r in some more <strong>de</strong>tail the subspace<br />

In,T = An(Dn,T ) ⊖ Vn(H)<br />

of An(Dn,T ). Since the range Vn(H) is invariant for S ∗ n , its orthogonal comp<strong>le</strong>ment In,T is<br />

invariant for the shift operator Sn. In other words, the space In,T is a shift invariant subspace of<br />

An(Dn,T ). The wan<strong>de</strong>ring subspace En,T for In,T is the subspace<br />

En,T = In,T ⊖ Sn(In,T )<br />

of In,T . To present our parametrization of the wan<strong>de</strong>ring subspace En,T for In,T we need some<br />

more notations.<br />

Let T ∈ L(H) be an n-hypercontraction. We <strong>de</strong>note by Hn the space H equipped with the<br />

equiva<strong>le</strong>nt norm<br />

x 2 n =<br />

n−1<br />

k=0<br />

(−1) k<br />

(see Lemma 3.1). It turns out that the operator<br />

TT ∗<br />

<br />

n T n−1<br />

k<br />

x2 2<br />

=x + Dk,T x<br />

k + 1<br />

2 , x ∈ H (0.6)<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

k=1<br />

in L(H)<br />

is self-adjoint in L(Hn) and has its spectrum contained in the closed unit interval [0, 1] (see<br />

Lemma 3.3). We <strong>de</strong>note by Qn,T the operator<br />

Qn,T =<br />

<br />

I − TT ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

1/2 in L(H),<br />

where the positive square root is computed in L(Hn).ByD∗ n,T we <strong>de</strong>note the clo<strong>sur</strong>e in H of the<br />

range of this operator Qn,T , and we equip this space D∗ n,T with the norm ·n <strong>de</strong>fined by (0.6).<br />

It turns out that we have the equality<br />

T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

in L(H) (see Lemma 3.4).<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

Qn,T = Dn,T T ∗<br />

<br />

n−1<br />

k=0<br />

(−1) k<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

(0.7)


522 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

In the case n = 1 of a contraction operator T ∈ L(H) the notions in the previous paragraph<br />

specialize to well-known objects. The norm ·1 <strong>de</strong>fined by (0.6) coinci<strong>de</strong>s with the usual norm<br />

of H. The operator Q1,T is the <strong>de</strong>fect operator for the adjoint operator T ∗ , that is, Q1,T = DT ∗,<br />

and the space D ∗ 1,T is the <strong>de</strong>fect space D∗ 1,T = DT ∗ for T ∗ . The equality (0.7) re<strong>du</strong>ces to the<br />

well-known formu<strong>la</strong> T ∗ DT ∗ = DT T ∗ for <strong>de</strong>fect operators.<br />

For an n-hypercontraction T ∈ L(H) we consi<strong>de</strong>r the operator-valued analytic function Wn,T<br />

in the unit disc D <strong>de</strong>fined by the formu<strong>la</strong><br />

Wn,T (z) =<br />

z ∈ D.<br />

<br />

−T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

n<br />

+ zDn,T (I − zT )<br />

k=1<br />

−k<br />

<br />

D<br />

Qn,T ,<br />

∗<br />

n,T<br />

Notice that by (0.7) the values Wn,T (z) attained by this function Wn,T are operators in<br />

L(D∗ n,T , Dn,T ), that is, boun<strong>de</strong>d linear operators from D∗ n,T into Dn,T .<br />

We remark that in the case n = 1 of a contraction operator T ∈ L(H) we get the so-cal<strong>le</strong>d<br />

characteristic operator function<br />

WT (z) = W1,T (z) = −T ∗ + zDT (I − zT ) −1 DT ∗<br />

<br />

DT , z∈D, ∗<br />

whose values are operators in L(DT ∗, DT ) which has been studied by Sz.-Nagy and Foias<br />

(see [26, Chapter VI]).<br />

We have the following <strong>de</strong>scription of the wan<strong>de</strong>ring subspace En,T . A function f in An(Dn,T )<br />

belongs to the wan<strong>de</strong>ring subspace En,T for In,T if and only if it has the form<br />

f(z)= Wn,T (z)x, z ∈ D, (0.8)<br />

for some e<strong>le</strong>ment x ∈ D∗ n,T . Furthermore, we have the norm equality<br />

f 2 An =x2 n , x ∈ D∗ n,T ,<br />

when f is given by (0.8). We recall that the norm ·n is <strong>de</strong>fined by (0.6). This parametrization<br />

of the wan<strong>de</strong>ring subspace En,T for In,T is the result of Theorem 3.3 in this paper. The proof of<br />

Theorem 3.3 proceeds in several steps and takes up Sections 2 and 3 in the paper.<br />

It turns out that the operator-valued analytic function Wn,T has a certain multiplier property<br />

in that it acts as a contractive multiplier from the Hardy space A1(D∗ n,T ) into the Bergman space<br />

An(Dn,T ) (see Theorem 4.1). This contractive multiplier property <strong>le</strong>ads in turn to an estimate<br />

Wn,T (z)Wn,T (z) ∗ <br />

1<br />

(1 −|z| 2 ) n−1 IDn,T in L(Dn,T ), z ∈ D (0.9)<br />

(see Theorem 4.2).<br />

In the case n = 1 of a contraction operator T ∈ L(H) in the c<strong>la</strong>ss C0· it is known that the<br />

characteristic operator function WT is an isometric multiplier from the Hardy space A1(DT ∗) into<br />

the Hardy space A1(DT ) with range equal to I1,T (see Corol<strong>la</strong>ry 4.1). Here the inequality (0.9)<br />

says that the characteristic operator function WT attains contractive values (see Corol<strong>la</strong>ry 4.2).


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 523<br />

We mention also that the contractive multiplier property of Wn,T and the inequality (0.9)<br />

generalize to a vector-valued context known properties of Bergman inner functions going back<br />

to He<strong>de</strong>nmalm [13–15] for n = 2, 3.<br />

As we have indicated above the previous consi<strong>de</strong>rations apply also to general shift invariant<br />

subspaces in the Bergman spaces An(E). LetI be a shift invariant subspace of An(E). Nowthe<br />

orthogonal comp<strong>le</strong>ment<br />

H = An(E) ⊖ I<br />

of I is invariant for S ∗ n and we set T = S∗ n |H. The operator T ∈ L(H) is an n-hypercontraction in<br />

the c<strong>la</strong>ss C0·. As above we can mo<strong>de</strong>l this operator T by means of the map Vn given by (0.5). Furthermore,<br />

by a uniqueness property of this representation, there exists an isometry ˆVn : Dn,T → E<br />

such that every function f ∈ H admits the representation<br />

f(z)= ˆVnDn,T (I − zT ) −n f, z ∈ D. (0.10)<br />

The isometry ˆVn : Dn,T → E is uniquely <strong>de</strong>termined by (0.10) and is given by<br />

ˆVn : Dn,T f ↦→ f(0) for f ∈ H.<br />

Full <strong>de</strong>tails of this construction can be found in [22, Sections 6 and 7].<br />

We write Ê = ˆVn(Dn,T ) ⊂ E. Themap ˆVn naturally extends to an isometry<br />

ˆVn : An(Dn,T ) → An(E)<br />

of An(Dn,T ) into An(E) with range equal to An(Ê) by setting<br />

<br />

( ˆVnf )(z) = ˆVn f(z) , z∈D, for f ∈ An(Dn,T ).<br />

The shift invariant subspace I now <strong>de</strong>composes as an orthogonal sum<br />

I = An(E ⊖ Ê) ⊕ ˆVn(In,T )<br />

(see Theorem 5.1), and we can i<strong>de</strong>ntify the wan<strong>de</strong>ring subspace EI for I as the orthogonal sum<br />

EI = (E ⊖ Ê) ⊕ ˆVn(En,T ).<br />

By our previous <strong>de</strong>scription of the wan<strong>de</strong>ring subspace En,T for In,T we have that a function f<br />

in An(E) belongs to the wan<strong>de</strong>ring subspace EI for I if and only if it has the form<br />

f(z)= a0 + ˆVnWn,T (z)g, z ∈ D, (0.11)<br />

for some e<strong>le</strong>ments a0 ∈ E ⊖ Ê and g ∈ D ∗ n,T . Furthermore, we have the norm equality f 2 An =<br />

a0 2 +g 2 n for f ∈ An(E) of the form (0.11) (see Theorem 5.2).<br />

As a bypro<strong>du</strong>ct of our consi<strong>de</strong>rations we obtain in an explicit form a parametrization of the<br />

shift invariant subspaces of the Hardy space A1(E) in case of a general not necessarily separab<strong>le</strong>


524 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

Hilbert space E (see Corol<strong>la</strong>ries 4.1 and 5.1). We discuss also some re<strong>la</strong>tions between the in<strong>de</strong>x<br />

of a shift invariant subspace and the <strong>de</strong>fect in<strong>de</strong>xes of the adjoint shift restricted to its orthogonal<br />

comp<strong>le</strong>ment (see Corol<strong>la</strong>ry 5.2 and Proposition 5.1).<br />

We wish to mention that a source of inspiration for the work presented in this artic<strong>le</strong> has been<br />

the <strong>sur</strong>vey paper [9] by Ball and Cohen.<br />

1. Preliminaries<br />

Let us first recall some constructions <strong>de</strong>veloped in the context of so-cal<strong>le</strong>d wan<strong>de</strong>ring subspace<br />

theorems in the papers [23,24]. The reason to inclu<strong>de</strong> this discussion here is that it provi<strong>de</strong>s a<br />

motivation for some of the arguments we shall use in <strong>la</strong>ter sections.<br />

Let T ∈ L(H) be an injective operator. It is easy to see that then the following statements are<br />

equiva<strong>le</strong>nt:<br />

• The operator T has closed range T(H).<br />

• The operator T is boun<strong>de</strong>d from below in the sense that Tx 2 cx 2 for x ∈ H and some<br />

positive constant c.<br />

• The operator T is <strong>le</strong>ft-invertib<strong>le</strong>.<br />

Let now T ∈ L(H) be an operator satisfying any of these conditions. The wan<strong>de</strong>ring subspace<br />

for the operator T ∈ L(H) is the subspace<br />

E = H ⊖ T(H) = ker T ∗<br />

of H. The operator L = (T ∗ T) −1 T ∗ in L(H) is the <strong>le</strong>ft-inverse of T with kernel E:<br />

The operator<br />

LT = I in L(H) and ker L = ker T ∗ = E.<br />

P = I − TL in L(H)<br />

is the orthogonal projection of H onto E. In<strong>de</strong>ed, the operator TL= T(T ∗ T) −1 T ∗ is self-adjoint,<br />

i<strong>de</strong>mpotent and has range equal to T(H).<br />

We shall also have use of the operator<br />

T ′ = L ∗ = T(T ∗ T) −1<br />

in L(H).<br />

The operator T ′ turns out to have some properties <strong>du</strong>al to those of T (see [21,24]). We notice<br />

that<br />

(T ′ ) ∗ T ′ = (T ∗ T) −1 T ∗ T(T ∗ T) −1 = (T ∗ T) −1<br />

in L(H). (1.1)<br />

Let us now specialize to the shift operator Sn on the Bergman space An(E). The formu<strong>la</strong>s<br />

(0.2) and (0.3) make evi<strong>de</strong>nt that the operator S∗ nSn on An(E) acts as<br />

<br />

∗ μn;k+1<br />

Sn Sn (f )(z) = akz<br />

μn;k<br />

k0<br />

k , z∈D,


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 525<br />

where f ∈ An(E) is given by (0.1). We notice that<br />

Snf 2 An = S ∗ <br />

nSnf,f = An<br />

k0<br />

ak 2 μn;k+1,<br />

where f ∈ An(E) is given by (0.1). It is straightforward to verify that the quotient μn;k+1/μn;k is<br />

increasing in k 0. As a result we have that the operator Sn is boun<strong>de</strong>d from below with constant<br />

c = 1/n.<br />

A computation shows that the operator Ln = (S ∗ n Sn) −1 S ∗ n on An(E) acts as<br />

(Lnf )(z) =<br />

f(z)− f(0)<br />

z<br />

= <br />

ak+1z k , z∈D, where f ∈ An(E) is given by (0.1). The operator S ′ n = L∗n = Sn(S∗ nSn) −1 is the weighted shift<br />

operator on An(E) acting as<br />

′<br />

S nf (z) = μn;k−1<br />

ak−1z k , z∈D, (1.2)<br />

k1<br />

μn;k<br />

where f ∈ An(E) is given by (0.1). We notice that<br />

<br />

S ′ nf 2 <br />

= An<br />

k1<br />

μ 2 n;k−1<br />

μn;k<br />

k0<br />

ak−1 2 = <br />

μ 2 n;k<br />

μn;k+1<br />

k0<br />

ak 2 , (1.3)<br />

where f ∈ An(E) is given by (0.1).<br />

Sums involving binomial coefficients will appear in our calcu<strong>la</strong>tions. For the sake of easy<br />

reference we record the following <strong>le</strong>mma.<br />

Lemma 1.1. Let μn;k = 1/ k+n−1 k for n 1 and k 0. Then<br />

min(n−1,k) <br />

(−1) j<br />

j=0<br />

Proof. A computation shows that<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

<br />

1<br />

j + 1 μn;k−j<br />

<br />

n<br />

z<br />

k + 1<br />

k =<br />

= 1<br />

.<br />

μn;k+1<br />

1 − (1 − z)n<br />

.<br />

z<br />

The sum in the <strong>le</strong>mma equals the kth Taylor coefficient of the function<br />

1 − (1 − z) n<br />

z<br />

<br />

<br />

1 1 1<br />

= − 1<br />

(1 − z) n z (1 − z) n<br />

This yields the conclusion of the <strong>le</strong>mma. ✷<br />

= <br />

1<br />

μn;k+1<br />

k0<br />

z k , z∈ D.


526 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

We shall need the following norm equality.<br />

Proposition 1.1. Let f ∈ An(E) be given by (0.1). Then<br />

n−1<br />

(−1) k<br />

k=0<br />

Proof. By(0.3)wehavethat<br />

<br />

n S<br />

k + 1<br />

∗k<br />

n f 2 <br />

= An<br />

k0<br />

μ 2 n;k<br />

μn;k+1<br />

<br />

S ∗j<br />

n f 2 μ<br />

= An<br />

k0<br />

2 n;k+j<br />

ak+j <br />

μn;k<br />

2<br />

ak 2 .<br />

for j 0. Summing these equalities we have by a change of or<strong>de</strong>r of summation that<br />

n−1<br />

(−1) j<br />

j=0<br />

<br />

n S∗j n f<br />

j + 1<br />

2 <br />

= An<br />

k0<br />

= <br />

<br />

min(n−1,k) <br />

(−1) j<br />

μ 2 n;k<br />

j=0<br />

μn;k+1<br />

k0<br />

where the <strong>la</strong>st equality follows by Lemma 1.1. ✷<br />

ak 2 ,<br />

<br />

n 1<br />

ak<br />

j + 1 μn;k−j<br />

2 μ 2 n;k<br />

We remark that the sums in Proposition 1.1 equal S ′ nf 2 by (1.3). By a po<strong>la</strong>rization argu-<br />

An<br />

ment we conclu<strong>de</strong> that<br />

∗ −1 ′ ∗S ′<br />

Sn Sn = S n n =<br />

where the first equality follows by (1.1).<br />

n−1<br />

(−1) k<br />

k=0<br />

2. A first <strong>de</strong>scription of the wan<strong>de</strong>ring subspace<br />

<br />

n<br />

S<br />

k + 1<br />

k nS∗k n in L An(E) , (1.4)<br />

We use the same basic notations as in the intro<strong>du</strong>ction. The operator Vn in L(H,An(Dn,T ))<br />

is <strong>de</strong>fined by (0.5),<br />

In,T = An(Dn,T ) ⊖ Vn(H) and En,T = In,T ⊖ Sn(In,T ).<br />

In this section we shall give a first <strong>de</strong>scription of the wan<strong>de</strong>ring subspace En,T for In,T in Theorem<br />

2.1. First we need a few <strong>le</strong>mmas.<br />

Lemma 2.1. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then the operator<br />

Pn = I − VnV ∗ n<br />

is the orthogonal projection of An(Dn,T ) onto In,T .<br />

in L An(Dn,T )


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 527<br />

Proof. We consi<strong>de</strong>r the operator Qn = VnV ∗ n . Recall that Vn : H → An(Dn,T ) is an isometry<br />

(see [22, Section 7]). It is straightforward to see that the operator Qn is self-adjoint, i<strong>de</strong>mpotent<br />

and has range equal to Vn(H). Accordingly the operator Qn is the orthogonal projection of<br />

An(Dn,T ) onto Vn(H), and Pn = I − Qn is the orthogonal projection of An(Dn,T ) onto In,T =<br />

An(Dn,T ) ⊖ Vn(H). ✷<br />

We next compute the operator V ∗ n .<br />

Lemma 2.2. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then the adjoint operator<br />

V ∗ n : An(Dn,T ) → H acts as<br />

V ∗ n<br />

<br />

f = T ∗k Dn,T ak weakly in H, (2.1)<br />

k0<br />

where f ∈ An(Dn,T ) is given by (0.1).<br />

Proof. For x ∈ H we have that<br />

∗<br />

Vn f,x <br />

<br />

=〈f,Vnx〉An = ak, 1 <br />

Dn,T T<br />

μn;k<br />

k x <br />

μn;k<br />

k0<br />

k0<br />

= <br />

ak,Dn,T T k x <br />

N<br />

= lim<br />

This gives the conclusion of the <strong>le</strong>mma. ✷<br />

N→∞<br />

k=0<br />

T ∗k Dn,T ak,x<br />

We remark that the sum in (2.1) converges in the weak topology in H.<br />

We shall next compute the operator V ∗ n (S∗ n Sn) −1 Vn = V ∗ n (S′ n )∗ S ′ n Vn.<br />

Lemma 2.3. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then<br />

V ∗ ∗ −1Vn<br />

n Sn Sn = V ∗ ′ ∗S ′<br />

n S n nVn =<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

.<br />

in L(H).<br />

Proof. Recall that the operator Vn in L(H,An(Dn,T )) is an isometry such that VnT = S ∗ n Vn<br />

(see [22, Sections 6 and 7]). By formu<strong>la</strong> (1.4) we have that<br />

V ∗ ∗ −1Vn<br />

n Sn Sn = V ∗ ′ ∗S ′<br />

n S n nVn = V ∗ n<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

S<br />

k + 1<br />

k nS∗k <br />

n Vn in L(H).<br />

The intertwining re<strong>la</strong>tion VnT = S∗ nVn gives that VnT k = S∗k n Vn for k 0, and taking adjoints<br />

we see that also T ∗kV ∗ n = V ∗ n Sk n for k 0. Using these intertwining formu<strong>la</strong>s we now have that


528 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

V ∗ ∗<br />

n Sn Sn<br />

<br />

−1Vn<br />

=<br />

n−1<br />

(−1) k<br />

k=0<br />

n−1<br />

=<br />

k=0<br />

(−1) k<br />

<br />

n<br />

T<br />

k + 1<br />

∗k V ∗ k<br />

n VnT<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

in L(H),<br />

where the <strong>la</strong>st equality follows by V ∗ n Vn = I in L(H). This comp<strong>le</strong>tes the proof of the<br />

<strong>le</strong>mma. ✷<br />

We can now give a first <strong>de</strong>scription of the wan<strong>de</strong>ring subspace En,T .<br />

Theorem 2.1. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then a function f in<br />

An(Dn,T ) belongs to the wan<strong>de</strong>ring subspace En,T for In,T if and only if it has the form<br />

f = a0 + S ′ nVnx ∗ −1Vnx<br />

= a0 + Sn Sn Sn<br />

for some e<strong>le</strong>ments a0 ∈ Dn,T and x ∈ H such that<br />

Proof. Notice first that<br />

and simi<strong>la</strong>rly that<br />

Here<br />

Dn,T a0 + T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

In,T = An(Dn,T ) ⊖ Vn(H) = ker V ∗ n ,<br />

En,T = In,T ⊖ Sn(In,T ) = ker(Sn|In,T )∗ .<br />

(Sn|In,T )∗ = PnS ∗ n = I − VnV ∗ ∗<br />

n Sn <br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x = 0. (2.2)<br />

by Lemma 2.1. An e<strong>le</strong>ment f in An(Dn,T ) thus belongs to the wan<strong>de</strong>ring subspace En,T for In,T<br />

if and only if V ∗ n f = 0 and (I − VnV ∗ n )S∗ nf = 0.<br />

We consi<strong>de</strong>r first the equation (I − VnV ∗ n )S∗ nf = 0. This equation can be rewritten as S∗ nf =<br />

VnV ∗ n S∗ n f . We apply the operator (S∗ n Sn) −1 to obtain that<br />

Lnf = S ∗ nSn −1S∗ nf = S ∗ nSn −1VnV ∗ n S∗ nf = S ∗ nSn −1Vnx, where x = V ∗ n S∗ nf ∈ H. We now have that<br />

∗ −1Vnx<br />

f = a0 + SnLnf = a0 + Sn Sn Sn = a0 + S ′ nVnx, (2.3)<br />

where a0 ∈ Dn,T and x ∈ H. Conversely, if f in An(Dn,T ) is of the form (2.3), then<br />

<br />

I − VnV ∗ ∗<br />

n Snf = I − VnV ∗ n Vnx = 0,


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 529<br />

since Vn is an isometry. We have thus shown that a function f in An(Dn,T ) satisfies the equation<br />

(I − VnV ∗ n )S∗ nf = 0 if and only if it has the form (2.3) for some e<strong>le</strong>ments a0 ∈ Dn,T and x ∈ H.<br />

We shall now compute V ∗ n f when f ∈ An(Dn,T ) is of the form (2.3) for some e<strong>le</strong>ments a0 ∈<br />

Dn,T and x ∈ H. Notice that the intertwining re<strong>la</strong>tion VnT = S∗ nVn gives that V ∗ n Sn = T ∗V ∗ n .By<br />

Lemma 2.2 we now have that<br />

V ∗ n f = Dn,T a0 + V ∗ n Sn<br />

∗ −1Vnx<br />

Sn Sn = Dn,T a0 + T ∗ V ∗ ∗ −1Vnx.<br />

n Sn Sn<br />

Recall that the operator V ∗ n (S∗ n Sn) −1 Vn was computed in Lemma 2.3. We conclu<strong>de</strong> that<br />

V ∗ n f = Dn,T a0 + T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x,<br />

where f ∈ An(Dn,T ), a0 ∈ Dn,T and x ∈ H are re<strong>la</strong>ted as in (2.3). This gives the conclusion of<br />

the theorem. ✷<br />

We shall next compute the norm of a function of the form f = a0 + S ′ n Vnx.<br />

Theorem 2.2. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Let f in An(Dn,T ) be of<br />

the form<br />

for some e<strong>le</strong>ments a0 ∈ Dn,T and x ∈ H. Then<br />

Proof. By Lemma 2.3 we have that<br />

f = a0 + S ′ n Vnx<br />

f 2 An =a0 2 n−1<br />

+<br />

k=0<br />

(−1) k<br />

f 2 An =a0 2 + ′<br />

S nVnx 2 An =a0 2 n−1<br />

+<br />

This comp<strong>le</strong>tes the proof of the theorem. ✷<br />

3. Parametrization of the wan<strong>de</strong>ring subspace En,T<br />

<br />

n T<br />

k + 1<br />

k x 2 .<br />

k=0<br />

(−1) k<br />

<br />

n T <br />

k<br />

x2 .<br />

k + 1<br />

In this section we shall solve Eq. (2.2) and give a more refined <strong>de</strong>scription of the wan<strong>de</strong>ring<br />

subspace En,T for In,T using the operator-valued analytic function Wn,T .<br />

We shall need some constructions involving the use of <strong>de</strong>fect operators of contractions between<br />

Hilbert spaces. Let A ∈ L(H, K) be a contraction operator mapping a Hilbert space H<br />

into a Hilbert space K. Associated to this operator A we have the <strong>de</strong>fect operator DA <strong>de</strong>fined by<br />

DA = (I − A ∗ A) 1/2<br />

in L(H),


530 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

where the positive square root is used. Notice that<br />

x 2 =Ax 2 +DAx 2 , x ∈ H, (3.1)<br />

and that this equality (3.1) can be restated saying that I = A∗A + D2 A in L(H). The <strong>de</strong>fect space<br />

DA for A is <strong>de</strong>fined as the clo<strong>sur</strong>e in H of the range of the operator DA, that is, DA = DA(H).<br />

In the same way the adjoint operator A∗ ∈ L(K, H) has an associated <strong>de</strong>fect operator<br />

DA ∗ = (I − AA∗ ) 1/2<br />

in L(K),<br />

and a <strong>de</strong>fect space DA∗ = DA∗(K) contained in K. These operators satisfy the equalities<br />

DA∗A = ADA in L(H, K) and DAA ∗ = A ∗ DA∗ in L(K, H). (3.2)<br />

The verification of the equalities (3.2) uses the functional calculus for self-adjoint operators in<br />

Hilbert space (see [10, Section XXVII.1] for <strong>de</strong>tails).<br />

Using the equalities (3.1) and (3.2) in the previous paragraph one verifies that the block oper-<br />

ator matrix<br />

<br />

A DA<br />

θA =<br />

∗<br />

DA −A∗ <br />

: H ⊕ DA∗ → K ⊕ DA<br />

is a unitary operator. The construction of this unitary operator θA goes back to Halmos [11].<br />

Let us now return to an n-hypercontraction T ∈ L(H). We shall need the following <strong>le</strong>mma.<br />

Lemma 3.1. Let T ∈ L(H) be an n-hypercontraction. Then<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k = I +<br />

Proof. For 1 m n we <strong>de</strong>note by Σm the operator<br />

m−1 <br />

Σm =<br />

k=0<br />

(−1) k<br />

C<strong>le</strong>arly Σ1 = I .Form 2 we have that<br />

Σm − Σm−1 = (−1) m−1 T ∗(m−1) T m−1 m−2 <br />

+<br />

n−1<br />

D<br />

k=1<br />

2 k,T<br />

<br />

m<br />

T<br />

k + 1<br />

∗k T k<br />

k=0<br />

(−1) k<br />

in L(H).<br />

in L(H).<br />

<br />

m m − 1<br />

− T<br />

k + 1 k + 1<br />

∗k T k<br />

in L(H).<br />

We now use the standard formu<strong>la</strong> m m−1 m−1 k+1 = k+1 + k for binomial coefficients to conclu<strong>de</strong><br />

that<br />

m−1 <br />

Σm − Σm−1 =<br />

k=0<br />

(−1) k<br />

<br />

m − 1<br />

T<br />

k<br />

∗k T k = D 2 m−1,T<br />

An easy in<strong>du</strong>ction argument now comp<strong>le</strong>tes the proof of the <strong>le</strong>mma. ✷<br />

in L(H).


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 531<br />

Let T ∈ L(H) be an n-hypercontraction. We <strong>de</strong>note by Hn the Hilbert space H equipped with<br />

the equiva<strong>le</strong>nt norm<br />

x 2 n =<br />

n−1<br />

k=0<br />

(−1) k<br />

<br />

n T<br />

k + 1<br />

k x 2 =x 2 n−1<br />

+ Dk,T x 2 , x ∈ H. (3.3)<br />

The equality of these two expressions for the norm ·n in (3.3) follows by Lemma 3.1. We<br />

<strong>de</strong>note by In the inclusion map of H into Hn <strong>de</strong>fined by Inx = x for x ∈ H.<br />

Lemma 3.2. Let T ∈ L(H) be an n-hypercontraction, and consi<strong>de</strong>r the inclusion map<br />

In : H → Hn <strong>de</strong>fined by Inx = x for x ∈ H. Then the adjoint operator I ∗ n ∈ L(Hn, H) acts<br />

as<br />

I ∗ n x =<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

Proof. For x ∈ Hn and y ∈ H we have that<br />

∗<br />

In x,y n−1<br />

=〈x,Iny〉n =<br />

k=0<br />

(−1) k<br />

This gives the conclusion of the <strong>le</strong>mma. ✷<br />

k=1<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x, x ∈ Hn.<br />

<br />

n T n−1<br />

k k<br />

x,T y =<br />

k + 1<br />

We shall consi<strong>de</strong>r also the operator Tn = InT in L(H, Hn).<br />

k=0<br />

(−1) k<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x,y .<br />

Lemma 3.3. Let T ∈ L(H) be an n-hypercontraction. Then the operator Tn = InT in L(H, Hn)<br />

is a contraction operator with <strong>de</strong>fect operator and <strong>de</strong>fect space given by<br />

The adjoint operator T ∗ n ∈ L(Hn, H) acts as<br />

Proof. Notice first that<br />

DTn = Dn,T in L(H) and DTn = Dn,T .<br />

T ∗ ∗<br />

n x = T<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

I ∗ n In<br />

n−1<br />

= I +<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x, x ∈ Hn.<br />

k=1<br />

D 2 k,T<br />

in L(H)<br />

by Lemmas 3.1 and 3.2. By the standard formu<strong>la</strong> k+1 k k <br />

j = j + j−1 for binomial coefficients<br />

we have the equalities<br />

D 2 k+1,T = D2 k,T − T ∗ D 2 k,T T in L(H)


532 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

for 1 k n − 1. Using these equalities we compute that<br />

T ∗ n Tn = T ∗ I ∗ n InT = T ∗ n−1<br />

T + T ∗ D 2 k,T T = T ∗ n−1<br />

2<br />

T + Dk,T − D 2 <br />

k+1,T<br />

k=1<br />

= T ∗ T + D 2 1,T − D2 n,T = I − D2 n,T<br />

in L(H).<br />

The equality T ∗ n Tn + D2 n,T = I in L(H) shows that the operator Tn ∈ L(H, Hn) is a contraction<br />

with <strong>de</strong>fect operator and <strong>de</strong>fect space as in the <strong>le</strong>mma.<br />

The action of the adjoint operator T ∗ n is evi<strong>de</strong>nt by Lemma 3.2. ✷<br />

We can now refine the <strong>de</strong>scription of the wan<strong>de</strong>ring subspace from Theorem 2.1.<br />

Theorem 3.1. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Let the operators Tn and<br />

In in L(H, Hn) be as above. Then a function f in An(Dn,T ) belongs to the wan<strong>de</strong>ring subspace<br />

En,T for In,T if and only if it has the form<br />

k=1<br />

f =−T ∗ n y + S′ −1<br />

nVnIn DT ∗ n y, y ∈ DT ∗ n .<br />

Furthermore, we have the norm equality that f 2 An =y2 n .<br />

Proof. By Theorem 2.1 we know that f ∈ An(Dn,T ) belongs to the wan<strong>de</strong>ring subspace En,T<br />

for In,T if and only if it has the form<br />

f = a0 + S ′ n Vnx<br />

for some e<strong>le</strong>ments a0 ∈ Dn,T and x ∈ H such that Eq. (2.2) holds. Using Lemma 3.3 we can<br />

rewrite Eq. (2.2) as<br />

using the operators Tn and In. Herea0∈Dn,T = DTn and x ∈ H.<br />

Let us now solve Eq. (3.4). We shall use the unitary operator<br />

and its adjoint operator<br />

θTn =<br />

θ ∗ Tn =<br />

<br />

T ∗<br />

n<br />

Tn DT ∗ n<br />

T ∗ n Inx + DTn a0 = 0 (3.4)<br />

DTn −T ∗ n<br />

DT ∗ n<br />

a0<br />

<br />

: H ⊕ DT ∗ n → Hn ⊕ DTn<br />

DTn<br />

−Tn<br />

<br />

: Hn ⊕ DTn → H ⊕ DT ∗ n .<br />

Assume now that x ∈ H and a0 ∈ DTn satisfies (3.4). Then<br />

θ ∗ <br />

Inx T ∗<br />

Tn = n Inx + DTna0 <br />

0<br />

= ,<br />

y<br />

DT ∗ n Inx − Tna0


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 533<br />

where y = DT ∗ n Inx − Tna0 ∈ DT ∗. Applying the operator θTn to this <strong>la</strong>st equality we find that<br />

n<br />

Inx<br />

a0<br />

<br />

= θTn<br />

<br />

0 DT<br />

=<br />

y<br />

∗ n y<br />

−T ∗ n y<br />

<br />

.<br />

This makes evi<strong>de</strong>nt that every solution x ∈ H and a0 ∈ DTn of Eq. (3.4) is of the form<br />

<br />

x = I −1<br />

n DT ∗ n y,<br />

a0 =−T ∗ n y<br />

for some e<strong>le</strong>ment y ∈ DT ∗ n . Also, if x ∈ H and a0 ∈ DTn are given by (3.5) for some e<strong>le</strong>ment<br />

y ∈ DT ∗, then, by property (3.2) of <strong>de</strong>fect operators, Eq. (3.4) holds. We have thus shown that<br />

n<br />

the solutions of (3.4) are parametrized by (3.5). By (3.5) we now have that<br />

f = a0 + S ′ nVnx =−T ∗ n y + S′ −1<br />

nVnIn DT ∗ n y,<br />

where y ∈ DT ∗ n .<br />

Let us now prove the norm equality that f 2 An =y2 n .Letx ∈ H and a0 ∈ Dn,T be given<br />

by (3.5). By Theorem 2.2 we have that<br />

f 2 An =a0 2 n−1<br />

+<br />

k=0<br />

(−1) k<br />

<br />

n T<br />

k + 1<br />

k x 2 =a0 2 +x 2 n = T ∗ n y 2 +DT ∗ n y2 n =y2 n ,<br />

where the <strong>la</strong>st equality follows by (3.1). This comp<strong>le</strong>tes the proof of the theorem. ✷<br />

Let T ∈ L(H) be an n-hypercontraction. Notice that by Lemma 3.2 the operator TnT ∗ n in<br />

L(Hn) acts as<br />

TnT ∗ n x = TT∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x, x ∈ Hn.<br />

Since the operator Tn ∈ L(H, Hn) is a contraction by Lemma 3.3, the operator<br />

TT ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

in L(H)<br />

is self-adjoint in L(Hn) and has its spectrum contained in the closed unit interval [0, 1]. We<br />

<strong>de</strong>note by Qn,T the operator<br />

Qn,T =<br />

<br />

I − TT ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

1/2 in L(H),<br />

(3.5)


534 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

where the positive square root is computed in L(Hn). We <strong>de</strong>note by D∗ n,T the clo<strong>sur</strong>e in H of<br />

the range of the operator Qn,T , that is, D∗ n,T = Qn,T (H), and we equip this space D∗ n,T with the<br />

Hilbert space structure given by the norm ·n <strong>de</strong>fined by (3.3).<br />

We can now restate Theorem 3.1 using the space D∗ n,T as follows.<br />

Theorem 3.2. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then a function f in<br />

An(Dn,T ) belongs to the wan<strong>de</strong>ring subspace En,T for In,T if and only if it has the form<br />

f =−T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

Furthermore, we have the norm equality that f 2 An =x2 n .<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x + S ′ nVnQn,T x, x ∈ D ∗ n,T . (3.6)<br />

Proof. Recall the action of the adjoint operator T ∗ n ∈ L(Hn, H) given by Lemma 3.3. The map<br />

In : H → Hn naturally i<strong>de</strong>ntifies the space D∗ n,T with the <strong>de</strong>fect space DT ∗. The result is evi<strong>de</strong>nt<br />

n<br />

by Theorem 3.1. ✷<br />

We record also the following <strong>le</strong>mma.<br />

Lemma 3.4. Let T ∈ L(H) be an n-hypercontraction. Then<br />

in L(H).<br />

T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

Qn,T = Dn,T T ∗<br />

<br />

n−1<br />

k=0<br />

(−1) k<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

Proof. By (3.2) we have the formu<strong>la</strong> T ∗ n DT ∗ n = DTnT ∗ n in L(Hn, H). Recall that DTn = Dn,T by<br />

Lemma 3.3, and the action of T ∗ n given by the same <strong>le</strong>mma. This makes evi<strong>de</strong>nt the conclusion<br />

of the <strong>le</strong>mma. ✷<br />

Let T ∈ L(H) be an n-hypercontraction. We recall from the intro<strong>du</strong>ction the <strong>de</strong>finition of the<br />

operator-valued analytic function Wn,T :<br />

Wn,T (z) =<br />

z ∈ D.<br />

<br />

−T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

n<br />

+ zDn,T (I − zT )<br />

k=1<br />

−k<br />

<br />

D<br />

Qn,T ,<br />

∗<br />

n,T<br />

By Lemma 3.4 the values Wn,T (z) attained by this function Wn,T are operators in L(D ∗ n,T , Dn,T ).<br />

Notice that<br />

<br />

1<br />

μn;k+1<br />

k0<br />

z k = 1<br />

<br />

<br />

1<br />

− 1<br />

z (1 − z) n<br />

=<br />

n<br />

k=1<br />

1<br />

, z∈D. (1 − z) k


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 535<br />

The function Wn,T thus has the power series expansion<br />

Wn,T (z) =<br />

<br />

−T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

+ <br />

k1<br />

1<br />

μn;k<br />

<br />

Dn,T T k−1 <br />

D<br />

k<br />

Qn,T z ,<br />

∗<br />

n,T<br />

z ∈ D. (3.7)<br />

We can now parametrize the wan<strong>de</strong>ring subspace En,T for In,T using the function Wn,T as<br />

follows.<br />

Theorem 3.3. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then a function f in<br />

An(Dn,T ) belongs to the wan<strong>de</strong>ring subspace En,T for In,T if and only if it has the form<br />

f(z)= Wn,T (z)x, z ∈ D, (3.8)<br />

for some e<strong>le</strong>ment x ∈ D∗ n,T . Furthermore, we have the norm equality<br />

when f is of the form (3.8).<br />

f 2 An =x2n =x2 n−1<br />

+ Dk,T x 2 , x ∈ D ∗ n,T ,<br />

Proof. Let f ∈ En,T be given by (3.6). By formu<strong>la</strong>s (0.5) and (1.2) we have that<br />

f(z)=−T ∗<br />

<br />

n−1<br />

(−1) k<br />

k=0<br />

k=1<br />

<br />

n<br />

T<br />

k + 1<br />

∗k T k<br />

<br />

x + <br />

k1<br />

1<br />

μn;k<br />

By the power series expansion (3.7) of the function Wn,T we conclu<strong>de</strong> that<br />

f(z)= Wn,T (z)x, z ∈ D.<br />

The conclusion of the theorem is now evi<strong>de</strong>nt by Theorem 3.2. ✷<br />

Dn,T T k−1 Qn,T x z k , z∈ D.<br />

We remark that in the case n = 1 of a contraction operator T ∈ L(H) the L(DT ∗, DT )-valued<br />

analytic function<br />

WT (z) = W1,T (z) = −T ∗ + zDT (I − zT ) −1 DT ∗<br />

<br />

DT , z∈D, ∗<br />

is the characteristic operator function studied by Sz.-Nagy and Foias (see [26, Chapter VI]).<br />

4. Multiplier properties of the function Wn,T<br />

In this section we discuss some multiplier properties of the function Wn,T .Wefirstshow<br />

that the function Wn,T acts as a contractive multiplier from the Hardy space A1(D∗ n,T ) into the<br />

Bergman space An(Dn,T ).


536 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

Theorem 4.1. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then the function Wn,T<br />

acts as a contractive multiplier Wn,T : f ↦→ Wn,T f from the Hardy space A1(D∗ n,T ) into the<br />

Bergman space An(Dn,T ):<br />

Wn,T f 2 An f 2 ∗<br />

A1<br />

, f ∈ A1 Dn,T ;<br />

here the space D ∗ n,T is equipped with the norm ·n given by (3.3).<br />

Proof. Let f ∈ A1(D∗ n,T ) be a polynomial of the form (0.1) with coefficients ak ∈ D∗ n,T .We<br />

write the function Wn,T f as<br />

Wn,T f = <br />

S k nWn,T ak.<br />

k0<br />

Recall that by Theorem 3.3 the e<strong>le</strong>ments Wn,T ak all belong to the wan<strong>de</strong>ring subspace En,T .We<br />

rewrite the sum for Wn,T f as<br />

<br />

<br />

f = Wn,T a0 + Sn S k nWn,T <br />

ak+1 .<br />

Now, since the wan<strong>de</strong>ring subspace En,T for In,T is orthogonal to Sn(In,T ), we have that<br />

k0<br />

Wn,T f 2 An =Wn,T a0 2 An +<br />

<br />

<br />

<br />

=a0 2 n +<br />

<br />

<br />

<br />

Sn<br />

Sn<br />

<br />

k0<br />

<br />

k0<br />

S k n Wn,T ak+1<br />

S k n Wn,T ak+1<br />

where the <strong>la</strong>st equality follows by Theorem 3.3. The fact that the shift operator Sn on An(Dn,T )<br />

is a contraction now gives that<br />

Wn,T f 2 An a0 2 n +<br />

<br />

<br />

<br />

<br />

k0<br />

S k n Wn,T ak+1<br />

We can now iterate this <strong>la</strong>st inequality (4.1) to obtain that<br />

<br />

ak 2 n .<br />

Wn,T f 2 An<br />

k0<br />

<br />

2<br />

<br />

<br />

<br />

<br />

An<br />

2<br />

An<br />

,<br />

<br />

2<br />

An<br />

. (4.1)<br />

Since the space of D∗ n,T -valued polynomials is <strong>de</strong>nse in A1(D∗ n,T ), an approximation argument<br />

now yields the conclusion of the theorem. ✷<br />

Remark 4.1. We remark that the clo<strong>sur</strong>e in An(Dn,T ) of the range of the multiplier<br />

Wn,T : A1(D ∗ n,T ) → An(Dn,T ), that is, the clo<strong>sur</strong>e in An(Dn,T ) of the set of all functions of<br />

the form Wn,T g, where g ∈ A1(D∗ n,T ) is a D∗ n,T -valued polynomial, equals the shift invariant<br />

subspace [En,T ] generated by the wan<strong>de</strong>ring subspace En,T . In particu<strong>la</strong>r, the multiplier Wn,T<br />

maps A1(D ∗ n,T ) into In,T .


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 537<br />

We recall that the space D∗ n,T is equipped with the norm ·n given by (3.3). In particu<strong>la</strong>r,<br />

this means that the norm of A1(D∗ n,T ) is given by<br />

f 2 <br />

A1<br />

=<br />

k0<br />

for f ∈ A1(D∗ n,T ) as in (0.1).<br />

Let us consi<strong>de</strong>r the case n = 1 in some more <strong>de</strong>tail.<br />

ak 2 n<br />

Corol<strong>la</strong>ry 4.1. Let T ∈ L(H) be a contraction in the c<strong>la</strong>ss C0·. Then the characteristic operator<br />

function WT = W1,T is an isometric multiplier WT : f ↦→ WT f from the Hardy space A1(DT ∗)<br />

into the Hardy space A1(DT ) with range equal to I1,T .<br />

Proof. In this case the shift operator S1 on A1(DT ) is an isometry and we have equality<br />

in (4.1). This gives that the multiplier WT maps A1(DT ∗) isometrically into A1(DT ). Bythe<br />

von Neumann–Wold <strong>de</strong>composition of an isometry (see [26, Section I.1]), the range of the multiplier<br />

WT equals I1,T (see Remark 4.1). ✷<br />

We remark that the proof of Theorem 4.1 is mo<strong>de</strong><strong>le</strong>d on an argument of Shimorin [25,<br />

Lemma 2.1].<br />

Remark 4.2. In the sca<strong>la</strong>r case when the <strong>de</strong>fect spaces Dn,T and D∗ n,T are both one-dimensional<br />

and n = 2 the result of Theorem 4.1 is <strong>du</strong>e to He<strong>de</strong>nmalm [13,15]. The case n = 3 goes back to<br />

He<strong>de</strong>nmalm [14].<br />

We next show that the multiplier Wn,T : A1(D ∗ n,T ) → An(Dn,T ) is injective.<br />

Proposition 4.1. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Let the function Wn,T<br />

act as a multiplier from A1(D ∗ n,T ) into An(Dn,T ) as in Theorem 4.1. Denote by L the operator<br />

Then the intertwining re<strong>la</strong>tions<br />

L = (Sn|In,T )∗ Sn|In,T<br />

−1(Sn|In,T )∗<br />

in L(In,T ).<br />

SnWn,T = Wn,T S1 and LWn,T = Wn,T L1<br />

holds. In particu<strong>la</strong>r, the multiplier Wn,T : A1(D ∗ n,T ) → An(Dn,T ) is injective.<br />

Proof. The first intertwining re<strong>la</strong>tion SnWn,T = Wn,T S1 is obvious. Let us verify the second<br />

intertwining re<strong>la</strong>tion LWn,T = Wn,T L1. Recall from Section 1 that the operator L in L(In,T ) is<br />

the <strong>le</strong>ft-inverse of Sn|In,T with kernel ker L = ker(Sn|In,T )∗ = En,T .Let<br />

g(z) = <br />

bkz k , z∈D, (4.2)<br />

k0


538 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

be a D∗ n,T -valued polynomial. The function f = Wn,T g has the form<br />

f = <br />

k0<br />

S k n Wn,T bk<br />

and the e<strong>le</strong>ments Wn,T bk all belong to En,T (see Theorem 3.3). We now have that<br />

Lf = <br />

k1<br />

S k−1<br />

n Wn,T bk = <br />

S k nWn,T bk+1 = Wn,T L1g.<br />

k0<br />

This shows that LWn,T g = Wn,T L1g when is g is a D∗ n,T -valued polynomial. The intertwining<br />

re<strong>la</strong>tion LWn,T = Wn,T L1 now follows by a standard approximation argument.<br />

Let us now prove that the multiplier Wn,T : A1(D∗ n,T ) → An(Dn,T ) is injective. We shall<br />

use the operator P = I − SnL in L(In,T ) which is the orthogonal projection of In,T onto En,T<br />

(see Section 1). Let f = Wn,T g, where g ∈ A1(D∗ n,T ) is given by (4.2). A computation using the<br />

intertwining re<strong>la</strong>tions shows that<br />

PL k f = (I − SnL)L k <br />

Wn,T g = Wn,T I − S1L 1 k<br />

1 L1g = Wn,T bk, k0. By Theorem 3.3 this <strong>la</strong>st equality <strong>de</strong>termines the coefficients bk uniquely. This comp<strong>le</strong>tes the<br />

proof of the proposition. ✷<br />

Recall that the repro<strong>du</strong>cing kernel for a Hilbert space H of E-valued analytic functions in the<br />

unit disc D is the function KH : D × D → L(E) satisfying the conditions that KH(·,ζ)xbelongs to H for every ζ ∈ D and x ∈ E, and that<br />

<br />

f(ζ),x= f,KH(·,ζ)x <br />

H , ζ ∈ D, f∈ H, (4.3)<br />

for x ∈ E. This <strong>la</strong>st property (4.3) is cal<strong>le</strong>d the repro<strong>du</strong>cing property of the kernel function KH.<br />

We remind that the Bergman space An(E) has the repro<strong>du</strong>cing kernel<br />

Kn(z, ζ ) =<br />

where IE <strong>de</strong>notes the i<strong>de</strong>ntity operator on E.<br />

We next compute the repro<strong>du</strong>cing kernel for the space Vn(H).<br />

1<br />

(1 − ¯ζz) n IE, (z,ζ)∈ D × D, (4.4)<br />

Proposition 4.2. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then the repro<strong>du</strong>cing<br />

kernel for the space Vn(H) is the Dn,T -valued function given by<br />

KVn(H)(z, ζ ) = Dn,T (I − zT ) −n (I − ¯ζT ∗ ) −n Dn,T , (z,ζ)∈ D × D.<br />

Proof. Let f = Vnx be a function in Vn(H). Then for y ∈ Dn,T we have that<br />

<br />

f(ζ),y= Dn,T (I − ζT) −n x,y = x,(I − ¯ζT ∗ ) −n Dn,T y .


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 539<br />

The fact that the operator Vn in L(H,An(Dn,T )) is an isometry now gives that<br />

<br />

f(ζ),y= Vnx,Vn(I − ¯ζT ∗ ) −n Dn,T y <br />

An = f,Vn(I − ¯ζT ∗ ) −n Dn,T y <br />

This comp<strong>le</strong>tes the proof of the proposition. ✷<br />

We <strong>de</strong>note by W the range of the multiplier Wn,T : g ↦→ Wn,T g mapping A1(D∗ n,T ) into<br />

An(Dn,T ) by Theorem 4.1, that is, the space W consists of all functions f ∈ An(Dn,T ) of the<br />

form<br />

f(z)= Wn,T (z)g(z), z ∈ D, (4.5)<br />

for some g ∈ A1(D∗ n,T ). Recall that by Proposition 4.1 the function f ∈ W uniquely <strong>de</strong>termines<br />

the function g ∈ A1(D∗ n,T ) by (4.5). We equip the space W with the norm in<strong>du</strong>ced by A1(D∗ n,T ),<br />

that is, we set f 2 W =g2 A1 when f ∈ W and g ∈ A1(D∗ n,T ) are re<strong>la</strong>ted as in (4.5). In this way<br />

the space W becomes a Hilbert space of Dn,T -valued analytic functions in D. We next compute<br />

the repro<strong>du</strong>cing kernel function for the space W.<br />

Lemma 4.1. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Let the Hilbert space<br />

W of Dn,T -valued analytic functions in D be <strong>de</strong>fined as in the previous paragraph. Then the<br />

repro<strong>du</strong>cing kernel for the space W is given by<br />

KW(z, ζ ) = 1<br />

1 − ¯ζz Wn,T (z)Wn,T (ζ ) ∗ , (z,ζ)∈ D × D.<br />

Proof. Let f ∈ W and g ∈ A1(D ∗ n,T ) be as in (4.5). For x ∈ Dn,T we have that<br />

f(ζ),x = Wn,T (ζ )g(ζ ), x = g(ζ),Wn,T (ζ ) ∗ x <br />

n .<br />

By the repro<strong>du</strong>cing property of the kernel function K1 for the space A1(D∗ n,T ) we have that<br />

<br />

f(ζ),x= g,K1(·,ζ)Wn,T (ζ ) ∗ x <br />

Now since the function Wn,T acts as an isometric multiplier of A1(D∗ n,T ) onto the space W we<br />

conclu<strong>de</strong> that<br />

f(ζ),x = Wn,T g,Wn,T K1(·,ζ)Wn,T (ζ ) ∗ x <br />

W = f,Wn,T K1(·,ζ)Wn,T (ζ ) ∗ x <br />

W .<br />

This comp<strong>le</strong>tes the proof of the <strong>le</strong>mma. ✷<br />

The contractive multiplier property from Theorem 4.1 <strong>le</strong>ads to the following result on domination<br />

of repro<strong>du</strong>cing kernel functions.<br />

Theorem 4.2. Let T ∈ L(H) be an n-hypercontraction in the c<strong>la</strong>ss C0·. Then the Dn,T -valued<br />

function<br />

A1 .<br />

An .


540 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

L(z, ζ ) =<br />

1<br />

(1 − ¯ζz) n IDn,T − Dn,T (I − zT ) −n (I − ¯ζT ∗ ) −n Dn,T<br />

− 1<br />

1 − ¯ζz Wn,T (z)Wn,T (ζ ) ∗ , (z,ζ)∈ D × D,<br />

is positive <strong>de</strong>finite on D × D. In particu<strong>la</strong>r, we have the inequality<br />

1<br />

1 −|z| 2 Wn,T (z)Wn,T (z) ∗ + Dn,T (I − zT ) −n (I −¯zT ∗ ) −n Dn,T<br />

<br />

1<br />

(1 −|z| 2 ) n IDn,T in L(Dn,T ), z ∈ D.<br />

Proof. Let the space W be as in Lemma 4.1. By Theorem 4.1 and Remark 4.1 the space W<br />

is contractively embed<strong>de</strong>d into In,T . By this we mean that W ⊂ In,T and f 2 An f 2 W for<br />

f ∈ W. Recall that the space An(Dn,T ) is the orthogonal sum of the subspaces Vn(H) and In,T .<br />

The repro<strong>du</strong>cing kernel function for the space In,T is given by<br />

KIn,T (z, ζ ) = Kn(z, ζ ) − KVn(H)(z, ζ )<br />

=<br />

1<br />

(1 − ¯ζz) n IDn,T − Dn,T (I − zT ) −n (I − ¯ζT ∗ ) −n Dn,T , (z,ζ)∈ D × D,<br />

where the <strong>la</strong>st equality follows by Proposition 4.2 and (4.4). The repro<strong>du</strong>cing kernel function<br />

KW for the space W was computed in Lemma 4.1. It is known that a contractive embedding<br />

W ⊂ In,T is equiva<strong>le</strong>nt to the domination re<strong>la</strong>tion KW ≪ KIn,T of repro<strong>du</strong>cing kernel functions<br />

(see [7, Section I.7]). We conclu<strong>de</strong> that the function<br />

L(z, ζ ) = KIn,T (z, ζ ) − KW(z, ζ ), (z, ζ ) ∈ D × D,<br />

is positive <strong>de</strong>finite on D × D. This gives the positive <strong>de</strong>finiteness assertion in the theorem. The<br />

<strong>la</strong>st inequality in the theorem follows by setting z = ζ noticing that L(z, z) 0inL(Dn,T ) by<br />

positive <strong>de</strong>finiteness of the function L. This comp<strong>le</strong>tes the proof of the theorem. ✷<br />

In the case n = 2 of sca<strong>la</strong>r-valued Bergman inner functions the inequality in Theorem 4.2<br />

seems first to have appeared in Zhu [27, Theorem 4.2].<br />

Let us examine the case n = 1 somewhat closer.<br />

Corol<strong>la</strong>ry 4.2. Let T ∈ L(H) be a contraction in the c<strong>la</strong>ss C0·. Then<br />

1<br />

1<br />

IDT =<br />

1 − ¯ζz 1 − ¯ζz WT (z)WT (ζ ) ∗ + DT (I − zT ) −1 (I − ¯ζT ∗ ) −1 DT ,<br />

(z, ζ ) ∈ D × D.<br />

Proof. The space A1(DT ) is the orthogonal sum of the subspaces V1(H) and I1,T . By Corol<strong>la</strong>ry<br />

4.1 we know that the characteristic operator function WT is an isometric multiplier from<br />

A1(DT ∗) onto I1,T . The repro<strong>du</strong>cing kernel functions <strong>de</strong>compose accordingly. ✷


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 541<br />

We remark that the result of Corol<strong>la</strong>ry 4.2 is a well-known formu<strong>la</strong> in the context of unitary<br />

systems which can be proved by direct computation (see [10, Section XXVIII.2]). We have<br />

inclu<strong>de</strong>d it here merely as an illustration of our methods.<br />

5. Shift invariant subspaces in Bergman spaces<br />

The consi<strong>de</strong>rations in the previous sections have some consequences concerning general shift<br />

invariant subspaces in Bergman spaces. Let I be a shift invariant subspace of An(E). Weset<br />

H = An(E) ⊖ I and T = S ∗ n |H.<br />

The operator T in L(H) is an n-hypercontraction in the c<strong>la</strong>ss C0·. We can now mo<strong>de</strong>l this operator<br />

T as part of the adjoint shift operator S∗ n on the space An(Dn,T ) by means of the formu<strong>la</strong><br />

(Vnf )(z) = Dn,T (I − zT ) −n f = <br />

<br />

k + n − 1 Dn,T T<br />

k<br />

k f z k , z∈D. k0<br />

Furthermore, by a uniqueness property of this operator mo<strong>de</strong>l there exists an isometry ˆVn :<br />

Dn,T → E such that the functions f ∈ H all admit the representation<br />

<br />

f(z)= ˆVn (Vnf )(z) , z∈D. (5.1)<br />

The isometry ˆVn : Dn,T → E of coefficient spaces is uniquely <strong>de</strong>termined by (5.1) and acts as<br />

ˆVn : Dn,T f ↦→ f(0) for f ∈ H.<br />

Full <strong>de</strong>tails of this construction can be found in [22, Sections 6 and 7].<br />

We write Ê = ˆVn(Dn,T ) ⊂ E. ThemapˆVn naturally extends to an isometry of An(Dn,T ) into<br />

An(E) with range equal to An(Ê) by setting<br />

<br />

( ˆVnf )(z) = ˆVn f(z) , z∈D, for f ∈ An(Dn,T ).<br />

We have the following <strong>de</strong>scription of a general shift invariant subspace of An(E).<br />

Theorem 5.1. Let I be a shift invariant subspace of An(E), and <strong>le</strong>t H = An(E) ⊖ I and T =<br />

S ∗ n |H in L(H) be as above. Then the space I <strong>de</strong>composes as an orthogonal sum<br />

I = An(E ⊖ Ê) ⊕ ˆVn(In,T ),<br />

that is, a function f ∈ An(E) belongs to I if and only if it has the form of an orthogonal sum<br />

f = f1 + ˆVng, where f1 ∈ An(E) has all its Taylor coefficients in E ⊖ Ê and g belongs to the<br />

shift invariant subspace In,T of An(Dn,T ).<br />

Proof. By formu<strong>la</strong> (5.1) we have that ˆVn(Vn(H)) = H. In particu<strong>la</strong>r, the space H is contained<br />

already in An(Ê). Passing to the orthogonal comp<strong>le</strong>ment we have that


542 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

This comp<strong>le</strong>tes the proof of the theorem. ✷<br />

I = An(E) ⊖ H = An(E ⊖ Ê) ⊕ An(Ê) ⊖ H <br />

<br />

= An(E ⊖ Ê) ⊕ ˆVn An(Dn,T ) ⊖ Vn(H) <br />

= An(E ⊖ Ê) ⊕ ˆVn(In,T ).<br />

Recall that the wan<strong>de</strong>ring subspace EI for a shift invariant subspace I of An(E) is the subspace<br />

EI = I ⊖ Sn(I)<br />

of I. We have the following <strong>de</strong>scription of a general wan<strong>de</strong>ring subspace.<br />

Theorem 5.2. Let I be a shift invariant subspace of An(E), and <strong>le</strong>t H = An(E) ⊖ I and T =<br />

S ∗ n |H in L(H) be as above. Then a function f ∈ An(E) belongs to the wan<strong>de</strong>ring subspace EI<br />

for I if and only if it has the form<br />

f(z)= a0 + ˆVnWn,T (z)g, z ∈ D, (5.2)<br />

where a0 ∈ E ⊖ Ê and g belongs to the <strong>de</strong>fect space D ∗ n,T . The norm of a function f ∈ An(E) of<br />

the form (5.2) is given by<br />

f 2 An =a0 2 +g 2 n =a0 2 + <br />

where bk is the kth Taylor coefficient of g as in (4.2).<br />

μ 2 n;k<br />

μn;k+1<br />

k0<br />

bk 2 ,<br />

Proof. The form of the invariant subspace I was calcu<strong>la</strong>ted in Theorem 5.1. By this <strong>de</strong>scription<br />

we have that<br />

EI = (E ⊖ Ê) ⊕ ˆVn(En,T ),<br />

where En,T is the wan<strong>de</strong>ring subspace for the shift invariant subspace In,T of An(Dn,T ). The<br />

wan<strong>de</strong>ring subspace En,T was <strong>de</strong>scribed in Theorem 3.3 as the space of all functions of the form<br />

f(z)= Wn,T (z)g, z ∈ D,<br />

where g belongs to D∗ n,T , and norm given by f 2 An =g2n . The expression for the norm ·n<br />

using Taylor coefficients follows by Proposition 1.1. This yields the conclusion of the theorem.<br />

✷<br />

We have the following characterization of a re<strong>la</strong>ted c<strong>la</strong>ss of operator-valued analytic functions.<br />

Theorem 5.3. Let W be an L(F, E)-valued analytic function in the unit disc D such that for<br />

every x ∈ F the function Wx : z ↦→ W(z)x belongs to An(E) with the properties that


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 543<br />

• the norm equality Wx2 An =x2 holds for every x ∈ F, and<br />

• Wx ⊥ Sk nWx for all k 1 and x ∈ F.<br />

Then W(F) ={Wx: x ∈ F} is a wan<strong>de</strong>ring subspace in An(E). Denote by I the shift invariant<br />

subspace generated by W(F) in An(E), that is, I =[W(F)]. Let H = An(E) ⊖ I and T =<br />

S ∗ n |H ∈ L(H). Then there exists a unitary operator<br />

of F onto (E ⊖ Ê) ⊕ D ∗ n,T<br />

U =<br />

such that<br />

U1<br />

U2<br />

<br />

: F → (E ⊖ Ê) ⊕ D ∗ n,T<br />

W(z)x = U1x + ˆVnWn,T (z)U2x, x ∈ F, z∈ D. (5.3)<br />

Furthermore, the equality (5.3) <strong>de</strong>termines the operators U1 and U2 uniquely.<br />

Proof. C<strong>le</strong>arly W(F) is a closed subspace of An(E). By po<strong>la</strong>rization we have that Wx ⊥ Sk nWy for k 1 and x,y ∈ F. This shows that W(F) satisfies the <strong>de</strong>fining property of a wan<strong>de</strong>ring<br />

subspace, that is, W(F) ⊥ Sk nW(F) for k 1.<br />

We have that EI = W(F) (see Remark 5.1). By Theorem 5.2 the wan<strong>de</strong>ring subspace EI =<br />

W(F) for I consists of all functions f in An(E) of the form (5.2) with norm equality<br />

f 2 An =a0 2 +g 2 n .<br />

For x ∈ F we set Ux = (a0,g) when f = Wx is given by (5.2). This gives us a unitary map<br />

U from F onto (E ⊖ Ê) ⊕ D ∗ n,T . The operators U1 and U2 are uniquely <strong>de</strong>termined by (5.3) by<br />

uniqueness of the representation (5.2). This comp<strong>le</strong>tes the proof of the theorem. ✷<br />

Remark 5.1. It is a general fact often accredited to Halmos [12] that a wan<strong>de</strong>ring subspace is<br />

uniquely <strong>de</strong>termined by the invariant subspace it generates. Let T ∈ L(H) be a boun<strong>de</strong>d linear<br />

operator, and <strong>le</strong>t E be a closed subspace of H such that E ⊥ T k (E) for k 1. Set I =[E]T <br />

=<br />

k0 T k (E). Then I = E ⊕ ( <br />

k1 T k (E)), which gives that EI = I ⊖ T(I) = E.<br />

We recall that the operator-valued analytic function Wn,T is a contractive multiplier from the<br />

Hardy space A1(D ∗ n,T ) into the Bergman space An(Dn,T ) (see Theorem 4.1), and that a re<strong>la</strong>ted<br />

upper bound is avai<strong>la</strong>b<strong>le</strong> (see Theorem 4.2).<br />

Specializing to the case n = 1 we obtain a parametrization of the shift invariant subspaces of<br />

the Hardy space A1(E) for a general not necessarily separab<strong>le</strong> Hilbert space E.<br />

Corol<strong>la</strong>ry 5.1. Let I be a shift invariant subspace of the Hardy space A1(E). Then a function f<br />

in A1(E) belongs to the subspace I if and only if it has the form<br />

f(z)= f1(z) + ˆV1WT (z)g(z), z ∈ D, (5.4)<br />

for some functions f1 ∈ A1(E ⊖ Ê) and g ∈ A1(DT ∗). Furthermore, we have the norm equality<br />

f 2 A1 =f12 A1 +g2 A1 when f ∈ I, f1 ∈ A1(E ⊖ Ê) and g ∈ A1(DT ∗) are re<strong>la</strong>ted by (5.4).


544 A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545<br />

Proof. Recall the form of an invariant subspace calcu<strong>la</strong>ted in Theorem 5.1. By Corol<strong>la</strong>ry 4.1 the<br />

characteristic operator function WT is an isometric multiplier from the Hardy space A1(DT ∗) into<br />

the Hardy space A1(DT ) with range equal to I1,T . This comp<strong>le</strong>tes the proof of the corol<strong>la</strong>ry. ✷<br />

The previous results yield the following consequence concerning the in<strong>de</strong>x dim EI of a shift<br />

invariant I of An(E).<br />

Corol<strong>la</strong>ry 5.2. Let I be a shift invariant subspace of An(E) with E separab<strong>le</strong>. Set H = An(E)⊖I<br />

and T = S ∗ n |H. Then dim Dn,T dim E. If the <strong>de</strong>fect in<strong>de</strong>x dim Dn,T is finite, then<br />

dim EI = dim E − dim Dn,T + dim D ∗ n,T .<br />

Proof. The first inequality dim Dn,T dim E is evi<strong>de</strong>nt by the fact that the operator ˆVn ∈<br />

L(Dn,T , E) is an isometry (see [22, Section 7]). The second inequality is evi<strong>de</strong>nt by the <strong>de</strong>scription<br />

of the wan<strong>de</strong>ring subspace EI in Theorem 5.2. ✷<br />

We notice also that dim EI = dim D ∗ n,T if ˆVn(Dn,T ) = E.<br />

It has been known for some time that even in the sca<strong>la</strong>r case E = C the in<strong>de</strong>x dim EI of a shift<br />

invariant subspace I of An(C) for n 2 can equal any positive integer or +∞. This was first<br />

proved by Apostol et al. [5] using <strong>du</strong>al algebras, and <strong>la</strong>ter more explicit constructions have been<br />

found by He<strong>de</strong>nmalm et al. [19] and others.<br />

In the context of the Hardy space A1(E) with E separab<strong>le</strong> it is a result of Halmos [12] that the<br />

in<strong>de</strong>x dim EI of a shift invariant subspace I of A1(E) cannot exceed the in<strong>de</strong>x of the who<strong>le</strong> space<br />

A1(E) meaning that dim EI dim E. This inequality is naturally interpreted as an inequality of<br />

<strong>de</strong>fect in<strong>de</strong>xes as follows.<br />

Proposition 5.1. Let T ∈ L(H) be a contraction operator in the c<strong>la</strong>ss C0· acting on a separab<strong>le</strong><br />

Hilbert space H. Then dim DT ∗ dim DT .<br />

Proof. It is known that the characteristic operator function WT has non-tangential boundary values<br />

WT (e iθ ) in the strong operator topology for a.e. e iθ ∈ T. A well-known argument then shows<br />

that the operator WT (e iθ ) is an isometry in L(DT ∗, DT ) for a.e. e iθ ∈ T (see [26, Chapter V]).<br />

This gives the conclusion of the proposition. ✷<br />

For an n-hypercontraction T ∈ L(H) in the c<strong>la</strong>ss C0· and n 2 the corresponding inequality<br />

dim D ∗ n,T dim Dn,T of <strong>de</strong>fect in<strong>de</strong>xes is not true in general for the reasons quoted above.<br />

References<br />

[1] J. Ag<strong>le</strong>r, The Arveson extension theorem and coanalytic mo<strong>de</strong>ls, Integral Equations Operator Theory 5 (1982)<br />

608–631.<br />

[2] J. Ag<strong>le</strong>r, Hypercontractions and subnormality, J. Operator Theory 13 (1985) 203–217.<br />

[3] A. A<strong>le</strong>man, S. Richter, C. Sundberg, Beurling’s theorem for the Bergman space, Acta Math. 177 (1996) 275–310.<br />

[4] C.-G. Ambrozie, M. Engliš, V. Mül<strong>le</strong>r, Operator tup<strong>le</strong>s and analytic mo<strong>de</strong>ls over general domains in C n ,J.Operator<br />

Theory 47 (2002) 287–302.<br />

[5] C. Apostol, H. Bercovici, C. Foias, C. Pearcy, Invariant subspaces, di<strong>la</strong>tion theory and the structure of the pre<strong>du</strong>al<br />

of a <strong>du</strong>al algebra. I, J. Funct. Anal. 63 (1985) 369–404.<br />

[6] J. Arazy, M. Engliš, Analytic mo<strong>de</strong>ls for commuting operator tup<strong>le</strong>s on boun<strong>de</strong>d symmetric domains, Trans. Amer.<br />

Math. Soc. 355 (2003) 837–864.


A. Olofsson / Journal of Functional Analysis 236 (2006) 517–545 545<br />

[7] N. Aronszajn, Theory of repro<strong>du</strong>cing kernels, Trans. Amer. Math. Soc. 68 (1950) 337–404.<br />

[8] C. Ba<strong>de</strong>a, G. Cassier, Constrained von Neumann inequalities, Adv. Math. 166 (2002) 260–297.<br />

[9] J.A. Ball, N. Cohen, <strong>de</strong> Branges–Rovnyak operator mo<strong>de</strong>ls and systems theory: A <strong>sur</strong>vey, in: Topics in Matrix and<br />

Operator Theory, Rotterdam, 1989, in: Oper. Theory Adv. Appl., vol. 50, Birkhäuser, 1991, pp. 93–136.<br />

[10] I. Gohberg, S. Goldberg, M.A. Kaashoek, C<strong>la</strong>sses of Linear Operators, vol. II, Oper. Theory Adv. Appl., vol. 63,<br />

Birkhäuser, 1993.<br />

[11] P.R. Halmos, Normal di<strong>la</strong>tions and extensions of operators, Summa Brasil. Math. 2 (1950) 125–134.<br />

[12] P.R. Halmos, Shifts on Hilbert spaces, J. Reine Angew. Math. 208 (1961) 102–112.<br />

[13] H. He<strong>de</strong>nmalm, A factorization theorem for square area-integrab<strong>le</strong> analytic functions, J. Reine Angew. Math. 422<br />

(1991) 45–68.<br />

[14] H. He<strong>de</strong>nmalm, A factoring theorem for a weighted Bergman space, Algebra i Analiz 4 (1992) 167–176; St. Petersburg<br />

Math. J. 4 (1993) 163–174.<br />

[15] H. He<strong>de</strong>nmalm, A factoring theorem for the Bergman space, Bull. London Math. Soc. 26 (1994) 113–126.<br />

[16] H. He<strong>de</strong>nmalm, B. Korenblum, K. Zhu, Theory of Bergman Spaces, Springer, 2000.<br />

[17] H. He<strong>de</strong>nmalm, Y. Perdomo, Mean value <strong>sur</strong>faces with prescribed curvature form, J. Math. Pures Appl. 83 (2004)<br />

1075–1107.<br />

[18] H. He<strong>de</strong>nmalm, K. Zhu, On the failure of optimal factorization for certain weighted Bergman spaces, Comp<strong>le</strong>x<br />

Variab<strong>le</strong>s Theory Appl. 19 (1992) 165–176.<br />

[19] H. He<strong>de</strong>nmalm, S. Richter, K. Seip, Interpo<strong>la</strong>ting sequences and invariant subspaces of given in<strong>de</strong>x in the Bergman<br />

spaces, J. Reine Angew. Math. 477 (1996) 13–30.<br />

[20] A. Olofsson, A monotonicity estimate of the biharmonic Green function, Arch. Math. (Basel) 82 (2004) 240–244.<br />

[21] A. Olofsson, Wan<strong>de</strong>ring subspace theorems, Integral Equations Operator Theory 51 (2005) 395–409.<br />

[22] A. Olofsson, An operator-valued Berezin transform and the c<strong>la</strong>ss of n-hypercontractions, submitted for publication.<br />

[23] S. Richter, Invariant subspaces of the Dirich<strong>le</strong>t shift, J. Reine Angew. Math. 386 (1988) 205–220.<br />

[24] S.M. Shimorin, Wold-type <strong>de</strong>compositions and wan<strong>de</strong>ring subspaces of operators close to isometries, J. Reine<br />

Angew. Math. 531 (2001) 147–189.<br />

[25] S.M. Shimorin, On Beurling-type theorems in weighted l 2 and Bergman spaces, Proc. Amer. Math. Soc. 131 (2003)<br />

1777–1787.<br />

[26] B. Sz.-Nagy, C. Foias, Harmonic Analysis of Operators on Hilbert Space, North-Hol<strong>la</strong>nd, 1970.<br />

[27] K. Zhu, Sub-Bergman Hilbert spaces on the unit disk, Indiana Univ. Math. J. 45 (1996) 165–176.


Journal of Functional Analysis 236 (2006) 546–580<br />

www.elsevier.com/locate/jfa<br />

Hua operators and Poisson transform for boun<strong>de</strong>d<br />

symmetric domains ✩<br />

Khalid Koufany a , Genkai Zhang b,∗<br />

a Institut Élie Cartan, UMR 7502, Université Henri Poincaré (Nancy 1), BP 239,<br />

F-54506 Vandœuvre-lès-Nancy ce<strong>de</strong>x, France<br />

b Department of Mathematics, Chalmers University of Technology and Göteborg University,<br />

S-41296 Göteborg, Swe<strong>de</strong>n<br />

Received 8 December 2005; accepted 24 February 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 4 April 2006<br />

Communicated by Paul Malliavin<br />

Abstract<br />

Let Ω be a boun<strong>de</strong>d symmetric domain of non-tube type in Cn with rank r and S its Shilov boundary. We<br />

consi<strong>de</strong>r the Poisson transform Psf(z)for a hyperfunction f on S <strong>de</strong>fined by the Poisson kernel Ps(z, u) =<br />

(h(z, z) n/r /|h(z, u) n/r | 2 ) s , (z, u)×Ω ×S, s ∈ C.Forallssatisfying certain non-integral condition we find<br />

a necessary and sufficient condition for the functions in the image of the Poisson transform in terms of Hua<br />

operators. When Ω is the type I matrix domain in Mn,m(C) (n m), we prove that an eigenvalue equation<br />

for the second or<strong>de</strong>r Mn,n-valued Hua operator characterizes the image.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Boun<strong>de</strong>d symmetric domains; Shilov boundary; Invariant differential operators; Eigenfunctions; Poisson<br />

transform; Hua systems<br />

1. Intro<strong>du</strong>ction<br />

Let Ω = G/K be a Riemannian symmetric space. Any parabolic subgroup P of G <strong>de</strong>fines a<br />

boundary G/P of the symmetric space Ω. The Poisson transform is an integral operator from<br />

hyperfunctions on G/P into the space of eigenfunctions on Ω of the algebra D(Ω) G of invariant<br />

✩ Research of the authors is partly supported by European IHP network Harmonic Analysis and Re<strong>la</strong>ted Prob<strong>le</strong>ms.<br />

Research of G. Zhang is supported by Swedish Science Council (VR).<br />

* Corresponding author.<br />

E-mail addresses: khalid.koufany@iecn.u-nancy.fr (K. Koufany), genkai@math.chalmers.se (G. Zhang).<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.02.014


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 547<br />

differential operators. Any such boundary G/P can be viewed as coset space of the maximal<br />

boundary G/Pmin <strong>de</strong>fined by a minimal parabolic subgroup Pmin. In this case the most general<br />

result was obtained by Kashiwara et al. [7] where they proved that un<strong>de</strong>r certain conditions on<br />

the eigenvalues the Poisson transform is a G-isomorphism between the space of hyperfunctions<br />

on G/Pmin and the space of eigenfunctions of invariant differential operators on Ω, namely<br />

the Helgason conjecture. It thus arises the question of characterizing the image of the Poisson<br />

transform for other smal<strong>le</strong>r boundaries.<br />

Suppose Ω is a boun<strong>de</strong>d symmetric domain in a comp<strong>le</strong>x n-dimensional space V .LetSbe its Shilov boundary and r its rank. In this paper we consi<strong>de</strong>r the characterization of the image of<br />

the Poisson transform<br />

<br />

Psϕ(z) = Ps(z, u)ϕ(u) dσ (u)<br />

on the Shilov boundary S when s satisfies the following condition:<br />

<br />

−4 b + 1 + j a<br />

2<br />

S<br />

<br />

n<br />

+ (s − 1) /∈{1, 2, 3,...} for j = 0 and 1, (1)<br />

r<br />

where a and b are some structure constants of Ω. For a specific value of s (s = 1 in our parameterization)<br />

the kernel Ps(z, u), (z, u) ∈ Ω × S, is the so-cal<strong>le</strong>d Poisson kernel for harmonic<br />

functions, and the corresponding Poisson transform P := P1 maps hyperfunctions on S to harmonic<br />

functions on Ω; here harmonic functions are <strong>de</strong>fined as the smooth functions that are<br />

annihi<strong>la</strong>ted by all invariant differential operators that annihi<strong>la</strong>te the constant functions. When Ω<br />

is a tube domain Johnson and Korányi [6] proved that the image of the Poisson transform P is<br />

exactly the set of all Hua-harmonic functions. For non-tube domains the characterization of the<br />

image of the Poisson transform P was done by Berline and Vergne [1] where certain third-or<strong>de</strong>r<br />

differential Hua operator was intro<strong>du</strong>ced to characterize the image.<br />

In his paper [12] Shimeno consi<strong>de</strong>red the Poisson transform Ps on tube domains; it is proved<br />

that Poisson transform maps hyperfunctions on the Shilov boundary to certain solution space<br />

of the Hua operator. For general domains and for other boundaries, the image of the Poisson<br />

transform was characterized in [13]. However for the Shilov boundary of a non-tube domain<br />

the prob<strong>le</strong>m is still open. We will construct two Hua operators of the third or<strong>de</strong>r and use them<br />

to give a characterization. For the matrix ball Ir,r+b of r × (r + b)-matrices some eigenvalue<br />

equation for the second-or<strong>de</strong>r Hua operator (constructed by Hua [5] and reformu<strong>la</strong>ted by Berline<br />

and Vergne [1]) is proved to give the characterization. We proceed to exp<strong>la</strong>in the content of our<br />

paper.<br />

Hua operator of the second-or<strong>de</strong>r H for a general symmetric domain is <strong>de</strong>fined as a kC-valued<br />

operator, see Section 4. For tube domains it maps the Poisson kernels into the center of kC,<br />

namely the Poisson kernels are its eigenfunctions up to an e<strong>le</strong>ment in the center, but it is not<br />

true for non-tube domains, see Section 5. However for type I domains of non-tube type, see<br />

Section 6, there is a variant of the Hua operator, H (1) , by taking the first component of the<br />

operator, since in this case kC = k (1)<br />

C + k(2)<br />

C is a sum of two irre<strong>du</strong>cib<strong>le</strong> i<strong>de</strong>als. We prove that<br />

operator H (1) has the Poisson kernels as its eigenfunctions and we find the eigenvalues. We<br />

prove further that the eigenfunctions of the Hua operator H (1) are also eigenfunctions of invariant<br />

differential operators on Ω. For that purpose we compute the radial part of Hua operator H (1) ,


548 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

see Proposition 6.3. We give eventually the characterization of the image of Poisson transform<br />

in terms of Hua operator for type Ir,r+b domains:<br />

Theorem 1.1 (Theorem 6.1). Suppose s ∈ C satisfies the following condition:<br />

−4 b + 1 + j + (r + b)(s − 1) /∈{1, 2, 3,...} for j = 0 and 1.<br />

A smooth function f on Ir,r+b is the Poisson transform Ps(ϕ) of a hyperfunction ϕ on S if and<br />

only if<br />

H (1) f = (r + b) 2 s(s − 1)f Ir.<br />

Our method of proving the characterization is the same as that in [8] by proving that the boundary<br />

value of the Hua eigenfunctions satisfy certain differential equations and is thus <strong>de</strong>fined only<br />

on the Shilov boundary, neverthe<strong>le</strong>ss it requires several technically <strong>de</strong>manding computations. In<br />

Section 7 we study the characterization of the range of Poisson transform for general non-tube<br />

domains. We construct two new Hua operators of the third or<strong>de</strong>r and prove, by essentially the<br />

same method as for the previous theorem, the characterization of the image of Poisson transform<br />

using the third-or<strong>de</strong>r Hua-type operators U and W:<br />

Theorem 1.2 (Theorem 7.2). Let Ω be a boun<strong>de</strong>d symmetric non-tube domain of rank r in C n .<br />

Let s ∈ C and put σ = n r s. If a smooth function f on Ω is the Poisson transform Ps of a hyperfunction<br />

in B(S), then<br />

Conversely, suppose s satisfies the condition<br />

<br />

−4 b + 1 + j a<br />

2<br />

<br />

U − −2σ 2 + 2pσ + c<br />

σ(2σ − p − b) W<br />

<br />

f = 0. (2)<br />

<br />

n<br />

+ (s − 1) /∈{1, 2, 3,...} for j = 0 and 1.<br />

r<br />

Let f be an eigenfunction f ∈ M(λs) (see (8)) with λs given by (11). Iff satisfies (2) then it is<br />

the Poisson transform Ps(ϕ) of a hyperfunction ϕ on S.<br />

After this paper was finished we were informed by Professor T. Oshima that he and N. Shimeno<br />

have obtained some simi<strong>la</strong>r results about Poisson transforms and Hua operators.<br />

2. Preliminaries and notation<br />

2.1. General setting<br />

We recall some basic facts about the Jordan trip<strong>le</strong> characterization of boun<strong>de</strong>d symmetric<br />

domains and fix notations. Our presentation is mainly based on [9]. Let Ω be an irre<strong>du</strong>cib<strong>le</strong><br />

boun<strong>de</strong>d symmetric domain in a comp<strong>le</strong>x n-dimensional space V .LetG be the i<strong>de</strong>ntity component<br />

of the group of biholomorphic automorphisms of Ω, and K be the isotropy subgroup of G


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 549<br />

at the point 0 ∈ Ω. Then K is a maximal compact subgroup of G and as a Hermitian symmetric<br />

space, Ω = G/K. Letg be the Lie algebra of G, and<br />

g = k + p<br />

be its Cartan <strong>de</strong>composition. The Lie algebra k of K has one-dimensional center z. Then there<br />

exists an e<strong>le</strong>ment Z0 ∈ z such that ad Z0 <strong>de</strong>fines the comp<strong>le</strong>x structure of p.Let<br />

gC = p + ⊕ kC ⊕ p −<br />

be the corresponding eigenspace <strong>de</strong>composition of gC, the comp<strong>le</strong>xification of g. We will use the<br />

Jordan theoretic characterization of Ω; the corresponding Lie theoretic characterization will be<br />

then more transparent and which we will also use.<br />

There exists a quadratic map Q : V → End( ¯V,V) (here ¯V is the comp<strong>le</strong>x conjugate of V ),<br />

such that<br />

p ={ξv: v ∈ V },<br />

where ξv(z) = v − Q(z) ¯v. We will hereafter i<strong>de</strong>ntify p + with V by the natural mapping<br />

and p − with ¯V by the mapping<br />

1<br />

2 (ξv − iξiv) = v ↦→ v,<br />

− 1<br />

2 (ξv + iξiv) = Q(z) ¯v ↦→ ¯v ∈ ¯V ;<br />

we will write ¯v = Q(z) ¯v when viewed as e<strong>le</strong>ment in the Lie algebra and when no ambiguity<br />

would arise.<br />

Let {z ¯vw} be the po<strong>la</strong>rization of Q(z) ¯v, i.e.<br />

{z ¯vw}=Q(z + w)¯v − Q(z) ¯v − Q(w) ¯v.<br />

This <strong>de</strong>fines a trip<strong>le</strong> pro<strong>du</strong>ct V × ¯V × V → V , with respect to which V is a JB ⋆ -trip<strong>le</strong>, see [17].<br />

We <strong>de</strong>fine D(z, ¯v) ∈ End(V ) by<br />

D(z, ¯v)w ={z ¯vw}.<br />

The space V carries a K-invariant inner pro<strong>du</strong>ct<br />

〈z, w〉= 1<br />

tr D(z, ¯w), (4)<br />

p<br />

where “tr” is the trace functional on End(V ), and p = p(Ω) is the genus of Ω (see Eq. (6)).<br />

Besi<strong>de</strong> the Eucli<strong>de</strong>an norm, V carries also the spectral norm,<br />

<br />

<br />

z= <br />

1 <br />

D(z, ¯z) <br />

2 <br />

1/2<br />

,<br />

(3)


550 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

where the norm of an operator in End(V ) is taken with respect to the Hilbert norm 〈·,·〉 1/2 on V .<br />

The domain Ω can now be realized as the open unit ball of V with respect to the spectral norm,<br />

Ω = z ∈ V : z < 1 .<br />

An e<strong>le</strong>ment c ∈ V is a tripotent if {c ¯cc}=c. In the matrix Cartan domains (of the type I, II,<br />

and III, see below) the tripotents are exactly the partial isometries. Each tripotent c ∈ V gives<br />

rise to a Peirce <strong>de</strong>composition of V ,<br />

where<br />

V = V0(c) ⊕ V1(c) ⊕ V2(c),<br />

Vj (c) = v ∈ V : D(c, ¯c)v = jv .<br />

Two tripotents c1 and c1 are orthogonal if D(c1, ¯c2) = 0. Orthogonality is a symmetric re<strong>la</strong>tion.<br />

A tripotent c is minimal if it cannot be written as a sum of two non-zero orthogonal tripotents.<br />

A tripotent c is maximal if V0(c) ={0}. AJordan frame is a maximal family of pairwise orthogonal,<br />

minimal tripotents. It is known that the group K acts transitively on Jordan frames. In<br />

particu<strong>la</strong>r, the cardinality of all Jordan frames is the same, and is equal to the rank r of Ω.Every<br />

z ∈ V admits a (unique) spectral <strong>de</strong>composition z = r j=1 sj vj , where {vj } is a Jordan frame<br />

and s1 s2 ··· sr 0arethespectral values of z. The spectral norm of z is equal to the<br />

<strong>la</strong>rgest spectral value s1.<br />

Let us choose a Jordan frame {cj } r j=1 in V . Then, by the transitivity of K on frames, each<br />

e<strong>le</strong>ment z ∈ V admits a po<strong>la</strong>r <strong>de</strong>composition z = k r j=1 sj cj , where k ∈ K and sj are the spectral<br />

values of z. Lete = c1 + c2 +···+cr; then e is a maximal tripotent and the G-orbit, G · e,<br />

is the Shilov boundary S of Ω.Let<br />

V = <br />

0jkr<br />

be the joint Peirce <strong>de</strong>composition of V associated with the Jordan frame {cj } r j=1 , where<br />

Vj,k = v ∈ V : D(cℓ, ¯cℓ)v = (δℓ,j + δℓ,k)v: 1ℓ r <br />

for (j, k) = (0, 0) and V0,0 ={0}. By the minimality of cj , Vj,j = Ccj ,1 j r. The transitivity<br />

of K on the frames implies that the integers<br />

Vj,k<br />

a := dim Vj,k (1 j


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 551<br />

V2 becomes a Jordan algebra for the pro<strong>du</strong>ct xy ={xēy} with i<strong>de</strong>ntity e<strong>le</strong>ment e.Letn1 = dim V1<br />

and n2 = dim V2. Then we have<br />

The genus of Ω is<br />

Thus<br />

n1 = rb, n2 = r +<br />

and this is true for every minimal tripotent in V .<br />

Let<br />

r(r − 1)<br />

a and n = n1 + n2.<br />

2<br />

p = p(Ω) = 1<br />

tr D(e,ē) = (r − 1)a + b + 2. (6)<br />

r<br />

〈cj ,cj 〉= 1<br />

p tr D(cj , ¯cj ) = 1<br />

tr D(e,ē) = 1,<br />

rp<br />

a = Rξ1 +···+Rξr, ξj = ξcj ,j= 1,...,r.<br />

Then, a is the maximal abelian subspace of p.Let{βj } r j=1 ⊂ a∗ be the basis of a ∗ <strong>de</strong>termined by<br />

and <strong>de</strong>fine an or<strong>de</strong>ring on a ∗ such that<br />

βj (ξk) = 2δj,k, 1 j,k r,<br />

βr >βr−1 > ···>β1 > 0. (7)<br />

The restricted roots system Σ(g, a) of g re<strong>la</strong>tive to a is of type Cr or BCr and it consists of<br />

the roots ±βj (1 j r) with multiplicity 1, the roots ± 1 2 βj ± 1 2 βk (1 j = k r) with<br />

multiplicity a, and possibly the roots ± 1 2 βj (1 j r) with multiplicity 2b. The set positive<br />

roots Σ + (g, a) consists of 1 2 (βk ± βj ) (1 j


552 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

The Iwasawa <strong>de</strong>composition is then given by<br />

g = k ⊕ a ⊕ n − .<br />

Let, as usual, m = Zk(a) be the centralizer of a in k, then we have<br />

g = n − ⊕ m ⊕ a ⊕ n + .<br />

We <strong>le</strong>t P = Pmin = MAN be the minimal parabolic subgroup of G, with M, A and N the<br />

corresponding Lie groups with Lie algebras m, a and n− .<br />

Let t −<br />

C be the subspace<br />

t −<br />

C = CD(c1, ¯c1) +···+CD(cr, ¯cr)<br />

of kC. Then t −<br />

C is abelian and we extend it to a Cartan subalgebra tC = t −<br />

C + t+<br />

C of kC. The root<br />

system Ψ := Σ(gC, tC) of gC with respect to tC, when restricted to t −<br />

C is of the form<br />

Ψ | −<br />

tC = Σ(gC, t −<br />

<br />

C ) = ± 1<br />

2 (γk ± γj ), 1 j = k r; ±γj , ± 1<br />

2 γj<br />

<br />

, 1 j r ,<br />

where γj are the Harish-Chandra strongly orthogonal roots <strong>de</strong>fined by<br />

<br />

γj D(ck, ¯ck) = 2δjk, γj | +<br />

t = 0, 1 j,k r.<br />

C<br />

The set of compact roots Ψc := Σ(kC, tC) is such that<br />

Ψc| −<br />

tC =<br />

<br />

1<br />

2 (γk − γj ), 1 j = k r; ± 1<br />

2 γj<br />

<br />

, 1 j r ,<br />

and the set of non-compact roots Ψn satisfies<br />

Ψn| −<br />

tC =<br />

<br />

± 1<br />

2 (γk + γj ), 1 j = k r; ±γj , ± 1<br />

2 γj<br />

<br />

, 1 j r .<br />

We choose a consistent or<strong>de</strong>ring with (3) and (7)<br />

γr >γr−1 > ···>γ1.<br />

We will also need the set of positive non-compact roots Ψn| +<br />

,<br />

Ψn| +<br />

t −<br />

<br />

1<br />

=<br />

C 2 (γk + γj ), 1 j = k r; 1<br />

2 γj<br />

<br />

,γj , 1 j r .<br />

t −<br />

C


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 553<br />

2.2. Boun<strong>de</strong>d symmetric domain of type Ir,r+b<br />

Let V = Mr,r+b(C) be the vector space of comp<strong>le</strong>x r × (r + b)-matrices. V is a Jordan trip<strong>le</strong><br />

system for the following trip<strong>le</strong> pro<strong>du</strong>ct:<br />

Then the endomorphisms D(z, ¯v) are given by<br />

{x ¯yz}=xy ∗ z + zy ∗ x.<br />

D(z, ¯v)w ={z ¯vw}=zv ∗ w + wv ∗ z.<br />

There is a canonical and natural choice of frames. One consi<strong>de</strong>rs the standard matrix units {ei,j ,<br />

1 i r, 1 j r + b} and <strong>de</strong>fines cj = ej,j, 1 j r. Then the Pierce <strong>de</strong>composition<br />

V = <br />

0jkr Vj,k of V is given by<br />

Let<br />

Vj,j = Ccj , 1 j r,<br />

Vj,k = Cej,k + Cek,j, 1 j


554 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

is easily seen to be a maximal compact subgroup of G. The Lie algebra g of G <strong>de</strong>composes into<br />

g = k + p, where k, the Lie algebra of K, consists of all matrices<br />

<br />

a 0<br />

, a ∈ Mr,r(C), d ∈ Mr+b,r+b(C), a<br />

0 d<br />

∗ =−a, d ∗ =−d,<br />

and p consists of all matrices<br />

<br />

0 v<br />

v∗ <br />

, v∈Mr,r+b(C). 0<br />

The in<strong>du</strong>ced vector fields are given respectively by<br />

z ↦→ az − zd, and z ↦→ ξv(z) = v − zv ∗ z.<br />

The comp<strong>le</strong>x Lie algebra kC is given by the set of all matrices<br />

<br />

a<br />

0<br />

<br />

0<br />

,<br />

d<br />

a ∈ Mr,r(C), d ∈ Mr+b,r+b(C), tr(a) + tr(d) = 0.<br />

Hence, kC can be written as the sum<br />

where k (1)<br />

C<br />

and<br />

and k(2)<br />

C<br />

kC = k (1)<br />

C ⊕ k(2)<br />

C ,<br />

are the i<strong>de</strong>als consisting respectively of the matrices<br />

<br />

a 0<br />

0 − tr(a)<br />

r+b Ir+b<br />

<br />

, a ∈ Mr,r(C),<br />

<br />

0 0<br />

, d ∈ Mr+b,r+b(C), tr(d) = 0.<br />

0 d<br />

Then, i<strong>de</strong>ntifying kC as linear transformations of V ,wehave<br />

kC = span D(u, ¯v), u,v ∈ V <br />

and<br />

where the endomorphism D(u, ¯v) (1) is given by<br />

k (1)<br />

C = span D(u, ¯v) (1) ,u,v∈ V ,<br />

D(u, ¯v) (1) z = uv ∗ z.


3. The Poisson transform<br />

K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 555<br />

Let D(Ω) G be the algebra of all invariant differential operators on Ω. Recall the <strong>de</strong>finition<br />

of the Harish-Chandra eλ-function: eλ, forλ∈a∗ C is the unique N-invariant function on Ω such<br />

that<br />

<br />

eλ exp(t1ξ1 +···+trξr) · 0 = e 2t1(λ1+ρ1)+···+2tr (λr +ρr )<br />

.<br />

Then eλ are the eigenfunctions of T ∈ D(Ω) G and we <strong>de</strong>note χλ(T ) the corresponding eigenvalues.<br />

Denote further<br />

M(λ) = f ∈ C ∞ (Ω): Tf = χλ(T )f, T ∈ D(Ω) G . (8)<br />

Recall the parabolic subgroup P = Pmin intro<strong>du</strong>ced in Section 2.1. Corresponding to P there<br />

is the Poisson transform on the maximal boundary G/P = K/M.Forλ∈a∗ C ,thePoisson transform<br />

Pλ,K/M is <strong>de</strong>fined by<br />

<br />

Pλ,K/Mf(gK)=<br />

on the space B(K/M) of hyperfunctions on K/M.<br />

It is proved by Kashiwara et al. [7] that for λ ∈ a ∗ C ,if<br />

K<br />

−1<br />

eλ k g f(k)dk<br />

−2 〈λ,α〉<br />

/∈{1, 2, 3,...} (9)<br />

〈α, α〉<br />

for all α ∈ Σ + (g, a), then the Poisson transform is a G-isomorphism from B(K/M) onto M(λ).<br />

We now intro<strong>du</strong>ce the Poisson transform on the Shilov boundary. Let h(z) be the unique<br />

K-invariant polynomial on V whose restriction to Rc1 +···+Rcr is given by<br />

h<br />

r<br />

j=1<br />

tj cj<br />

<br />

=<br />

r<br />

j=1<br />

2<br />

1 − tj .<br />

As h is real-valued, we may po<strong>la</strong>rize it to get a polynomial on V × V , <strong>de</strong>noted by h(z, w),<br />

holomorphic in z and antiholomorphic in w such that h(z, z) = h(z). Recall that the function h<br />

is re<strong>la</strong>ted to the Bergman operator (see (12)) by that<br />

The Poisson kernel P(z,u)on Ω × S is<br />

<strong>de</strong>t b(z, ¯z) = h(z, ¯z) p .<br />

<br />

h(z, z)<br />

P(z,u)=<br />

|h(z, u)| 2<br />

n/r .


556 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

For a comp<strong>le</strong>x number s we <strong>de</strong>fine the Poisson transform Psϕ on the space B(S) of hyperfunctions<br />

ϕ on S by<br />

<br />

(Psϕ)(z) =<br />

S<br />

P(z,u) s ϕ(u) dσ (u).<br />

The kernel P(z,u) s has the following transformation property:<br />

P(gz,gu) s = Jg(u) 2ns<br />

− rp s<br />

P(z,u) , ∀g ∈ G, (10)<br />

where Jg(u) is the Jacobian of g at u.<br />

The kernel P(z,u) s ,foru = e is a special case of the eλ-function. The Poisson transform Ps<br />

on S can be viewed as a restriction of the Poisson transform Pλ,K/M. However for fixed s there<br />

are various choices of λ and we will find a specific λ so that the above condition (9) is valued<br />

when s satisfies (1). Let<br />

and consi<strong>de</strong>r the <strong>de</strong>composition<br />

ξc = ξ1 +···+ξr<br />

a = Rξc ⊕ ξ ⊥ c = Rξc<br />

r−1<br />

⊕ R(ξj − ξj+1)<br />

un<strong>de</strong>r the (negative) Killing form on g. We <strong>de</strong>note ξ ∗ c the <strong>du</strong>al vector, ξ ∗ c (ξc) = 1. We extend ξ ∗ c<br />

to a by the orthogonal projection <strong>de</strong>fined above. Observe first that<br />

We have then<br />

j=1<br />

ρ(ξc) = n = nξ ∗ c (ξc).<br />

Psf(z)= Pλs,K/Mf(z),<br />

where f on S is viewed as a function on K and thus on K/M, λs ∈ a∗ C is given by<br />

Thus<br />

λs = ρ + 2n(s − 1)ξ ∗ c . (11)<br />

PsB(S) ⊂ Pλs,K/MB(K/M) ⊂ M(λs).<br />

When s satisfies (1) we have then Pλs,K/MB(K/M) = M(λs).


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 557<br />

4. The second-or<strong>de</strong>r Hua operator<br />

We shall <strong>de</strong>fine the Hua operator both in terms of the enveloping algebra and using the covariant<br />

Cauchy–Riemann operator, the <strong>la</strong>ter having the advantage of being geometric and more<br />

explicit. To avoid some extra constants we fix and normalize the Killing form B on gC by requiring<br />

that on p + × p − it is given by<br />

B(u, ¯v) =〈u, v〉= 1<br />

p tr D(u, ¯v), u ∈ V = p+ , ¯v ∈ ¯V = p − ,<br />

where the trace is computed on the space V . (So the standard Killing form, (X, Y ) ↦→<br />

tr Ad(X) Ad(Y ), is−pB(X,Y).)<br />

Let {vj } and {v ∗ j } be <strong>du</strong>al bases of p+ and p − with respect to the normalized Killing form B.<br />

Let U(gC) be the enveloping algebra of gC. Since [p + , p − ]⊂kC, the operator<br />

H = HkC =−viv<br />

∗ j ⊗[vj ,v ∗ i ]<br />

i,j<br />

is an e<strong>le</strong>ment of U(gC) ⊗ kC, and is in<strong>de</strong>pen<strong>de</strong>nt of choice of the basis; it is cal<strong>le</strong>d the secondor<strong>de</strong>r<br />

Hua operator. If we i<strong>de</strong>ntify U(gC) with <strong>le</strong>ft-invariant differential operators on G, H <strong>de</strong>fines<br />

a homogeneous operator from C∞ (G/K) to the C∞-sections of G ×K kC. H can also be<br />

viewed as a differential operator from C∞ (G) to C∞ (G, kC).<br />

For X ∈ kC, <strong>de</strong>fine<br />

H X =− <br />

[X, vj ]v ∗ j ∈ U(gC).<br />

j<br />

Let S be a linear subspace of k, SC its comp<strong>le</strong>xification. Let {Xj } be a basis of SC and {X∗ j }<br />

be the <strong>du</strong>al basis with respect to the Killing form B. Then the projection of H onto U(gC) ⊗ SC<br />

is<br />

<br />

= H Xj ∗<br />

⊗ Xj .<br />

HSC<br />

It can also be <strong>de</strong>fined in<strong>de</strong>pen<strong>de</strong>ntly of basis, see e.g. [8, Proposition 1].<br />

Symbolically we may write<br />

where<br />

j<br />

H = D b(z, ¯z)¯∂,∂ ,<br />

b(z, ¯w) = 1 − D(z, ¯w) + Q(z)Q( ¯w) (12)<br />

is the Bergman operator, see [3]. (Operator H can also be <strong>de</strong>fined by using the covariant Cauchy–<br />

Riemannian operator b(z, ¯z)¯∂, see [2,14,20]. For brevity we will not go into the <strong>de</strong>tails.) Using


558 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

an orthonormal basis {ej } of V with respect to the Hermitian sca<strong>la</strong>r pro<strong>du</strong>ct (4), operator H can<br />

also be written as<br />

Hf(z)= <br />

D <br />

b(z, ¯z)ēi,ej ¯∂i∂j f(z).<br />

5. The Poisson kernel and the second-or<strong>de</strong>r Hua operator<br />

i,j<br />

We will compute the action of the Hua operator on the Poisson kernel. Let us first recall the<br />

notion of quasi-inverse in the Jordan trip<strong>le</strong> V ; see [9]. Let z ∈ V and ¯w ∈ ¯V . The e<strong>le</strong>ment z is<br />

cal<strong>le</strong>d quasi-invertib<strong>le</strong> with respect to ¯w, ifb(z, ¯w) is invertib<strong>le</strong>. The quasi-inverse of z with<br />

respect to ¯w is then given by<br />

z ¯w = b(z, ¯w) −1 z − Q(z) ¯w .<br />

For examp<strong>le</strong>, in the type Ir,r+b case (see Section 2.2), <strong>le</strong>t x,y ∈ V = Mr,r+b(C), then<br />

b(x, ¯y)z = (I − xy ∗ )z(I − y ∗ x).<br />

If I − xy ∗ is invertib<strong>le</strong>, then the quasi-inverse of x is<br />

x ¯y = b(x, ¯y) −1 x − Q(x) ¯y = (I − xy ∗ ) −1 (x − xy ∗ x)(I − y ∗ x) −1 = (I − xy ∗ ) −1 x.<br />

Fix a Jordan frame {cj }1jr and choose an orthonormal basis {eα} of V consisting of the<br />

frame {cj }1jr, orthonormal basis of each of the subspaces Vjk and an orthonormal basis of<br />

each of the subspaces Vj0. The following <strong>le</strong>mma can be easily proved by direct computations.<br />

Lemma 5.1.<br />

(1) For any irre<strong>du</strong>cib<strong>le</strong> boun<strong>de</strong>d symmetric domain Ω it holds:<br />

(a) r α=1 D(eα, ēα) = pZ0.<br />

(b) <br />

eα∈Vjk D(eα, ēα) = a 2 [D(cj , ¯cj ) + D(ck, ¯ck)].<br />

(2) If Ω is of type Ir,r+b, then <br />

eα∈Vj0 D(eα, ēα) (1) = bD(cj , ¯cj ) (1) .<br />

We need also the following <strong>le</strong>mma.<br />

Lemma 5.2. Let ¯w ∈ ¯V . For any comp<strong>le</strong>x number s, the holomorphic and the anti-holomorphic<br />

differential of the function z ↦→ h(z, ¯w) are given by<br />

∂h(z, ¯w) s =−sh(z, ¯w) s ¯w z ,<br />

Proof. This is a consequence of the formu<strong>la</strong><br />

see [21, Proposition 3.1]. ✷<br />

¯∂h(z, ¯w) s =−sh(z, ¯w) s w ¯z .<br />

¯w z =−∂ log <strong>de</strong>t b(z, ¯w) 1/p =−∂ log h(z, ¯w),


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 559<br />

Theorem 5.3. For u fixed in S, function<br />

satisfies the following differential equation:<br />

z ↦→ Ps,u(z) := P(z,u) s<br />

<br />

n<br />

HPs,u(z) =<br />

r s<br />

2 D b(z, ¯z) z ¯z − u ¯z , ¯z z −ū z <br />

n<br />

−<br />

r sp<br />

<br />

Z0 Ps,u(z).<br />

Proof. Choose a basis {eα} as in Lemma 5.1. Then<br />

HPs,u(z) = <br />

D <br />

b(z, ¯z)eα, ēβ ¯∂α∂βPs,u(z).<br />

According to Lemma 5.2,<br />

α,β<br />

<br />

h(z, z)<br />

∂Ps,u(z) = ∂<br />

|h(z, u)| 2<br />

n<br />

r s<br />

=− n<br />

r s<br />

<br />

h(z, z)<br />

|h(z, u)| 2<br />

n<br />

r s¯z z z<br />

−ū ,<br />

where we have i<strong>de</strong>ntified (p− ) ′ with p + by the Hermitian form (4). Performing one more time<br />

differentiation, we get<br />

<br />

¯∂∂Ps,u(z)<br />

n<br />

=<br />

r s<br />

2 Ps,u(z) z ¯z − u ¯z ⊗ ¯z z −ū z − n<br />

r sPs,u(z)¯∂ ¯z z −ū z .<br />

Moreover,<br />

¯∂ ¯z z −ū z = ¯∂ ¯z z = ¯∂∂ log h(z, ¯z) −1 = b(z, ¯z) −1 Id,<br />

where Id is the i<strong>de</strong>ntity form in (p + ) ′ ⊗ (p− ) ′ . Hence,<br />

<br />

¯∂∂Ps,u(z)<br />

n<br />

=<br />

r s<br />

2 Ps,u(z) z ¯z − u ¯z ⊗ ¯z z −ū z <br />

n<br />

−<br />

r s<br />

<br />

b(z, ¯z) −1 Id.<br />

Consequently<br />

<br />

n<br />

HPs,u(z) =<br />

r s<br />

2 <br />

<br />

n<br />

−<br />

r s<br />

α,β<br />

<br />

α<br />

z ¯z − u ¯z ,eα<br />

<br />

D(eα, ēα) Ps,u(z)<br />

<br />

z z<br />

¯z −ū , ēβ D b(z, ¯z)eα, ēβ<br />

<br />

n<br />

=<br />

r s<br />

2 D b(z, ¯z) z ¯z − u ¯z , ¯z z −ū z <br />

n<br />

−<br />

r s<br />

<br />

pZ0 Ps,u(z),<br />

since <br />

α D(eα, ēα) = pZ0. ✷<br />

If Ω is of tube type, then the genus p is given by p = 2 n r and Theorem 5.3 becomes:


560 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

Corol<strong>la</strong>ry 5.4. Let Ω be a tube type domain. For any u ∈ S, the function z ↦→ Ps,u(z) satisfies<br />

the Hua equation<br />

where I is the i<strong>de</strong>ntity operator.<br />

2 n<br />

HPs,u(z) = 2 s(s − 1)Ps,u(z)I, (13)<br />

r<br />

This corol<strong>la</strong>ry has been proved also by Faraut and Korányi [3, Theorem XIII.4.4]. Notice that<br />

the first factor 2 in (13) is because in this case our Hua operator is twice the Hua operator of<br />

Faraut and Korányi. In fact, we are using the <strong>de</strong>finition [u, ¯v]=D(u, ¯v) so that for tube domain<br />

it is twice the “square” operator ✷ of Faraut and Korányi.<br />

In [12, Theorem 4.1] Shimeno gives the following characterization of the image of Poisson<br />

transform for tube type domains.<br />

Theorem 5.5. Let Ω be a tube type domain. Suppose s ∈ C satisfies the following condition:<br />

<br />

−4 1 + j a<br />

2<br />

<br />

n<br />

+ (s − 1) /∈{1, 2, 3,...} for j = 0 and 1.<br />

r<br />

A smooth function f on Ω is the Poisson transform Ps of a hyperfunction on S if and only if f<br />

satisfies the following Hua equation<br />

2 n<br />

Hf = 2 s(s − 1)f Z0.<br />

r<br />

This is a slightly different formu<strong>la</strong>tion of Shimeno’s result. In fact, if s ′ <strong>de</strong>notes the Shimeno’s<br />

parameter, then our parameter s is<br />

s = r<br />

<br />

s<br />

2n<br />

′ + n<br />

<br />

.<br />

r<br />

6. The main result for type Ir,r+b domains<br />

In this section we restrict ourself to the case Ω = Ir,r+b. Recall that in Section 2.2 we have<br />

fixed a <strong>de</strong>composition kC = k (1)<br />

C ⊕ k(2)<br />

C .We<strong>le</strong>tH(1) be the first component of the Hua operator H.<br />

Symbolically H (1) is given by<br />

and can be i<strong>de</strong>ntified with the operator<br />

H (1) = D b(z, ¯z)¯∂,∂ (1) ,<br />

(Ir − zz ∗ )¯∂z · (Ir+b − z ∗ z) · t ∂z<br />

intro<strong>du</strong>ced by Hua [5], since in this case b(z, ¯z)v = (I − zz ∗ )v(I − z ∗ z).<br />

We state now the main theorem of this section.


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 561<br />

Theorem 6.1. Suppose s ∈ C satisfies the following condition:<br />

−4 b + 1 + j + (r + b)(s − 1) /∈{1, 2, 3,...} for j = 0 and 1. (14)<br />

A smooth function f on Ir,r+b is the Poisson transform Ps(ϕ) of a hyperfunction ϕ on S if and<br />

only if f satisfies the following Hua equation:<br />

where Ir is the i<strong>de</strong>ntity matrix of rank r.<br />

H (1) f = (r + b) 2 s(s − 1)f Ir, (15)<br />

Note here that the constant r + b = n/r for the domain Ir,r+b.<br />

6.1. The necessity of the Hua equation (15)<br />

To show the necessity of the Hua equation it is sufficient to show that the function Ps,u satisfies<br />

(15) for every u ∈ S.<br />

Proposition 6.2. If Ω is of type Ir,r+b, then<br />

H (1) Ps,u(z) = (r + b) 2 s(s − 1)Ps,u(z)Ir.<br />

Proof. It is sufficient to prove the formu<strong>la</strong> at z = 0. Specifying the result of Theorem 5.3 to the<br />

type Ir,r+b domain we get for any u ∈ S,<br />

H (1) <br />

n<br />

Ps,u(0) =<br />

r s<br />

2 D(u, u) (1) <br />

n<br />

−<br />

r sp<br />

<br />

Z (1)<br />

<br />

0 Ps,u(0).<br />

Now, obviously D(u, u) (1) = Ir and Z (1)<br />

0<br />

= r+b<br />

2r+b Ir. Therefore,<br />

H (1) Ps,u(0) = (r + b) 2 s(s − 1)Ps,u(0)Ir. ✷<br />

6.2. The Hua operator and the eigenfunctions of invariant differential operators<br />

We give first the expression for the radial part of the Hua operator H (1) , i.e. its restriction to<br />

K-invariant functions. We fix a Jordan frame {cj } r j=1 , then every e<strong>le</strong>ment of V can be written as<br />

z = k<br />

r<br />

j=1<br />

tj cj ,<br />

with k ∈ K, and tj 0. If f is a function on Ω invariant un<strong>de</strong>r K, we write<br />

f(z)= F(t1,...,tr).<br />

The function F is a symmetric function of the variab<strong>le</strong>s t1,...,tr, <strong>de</strong>fined on the unit cube<br />

0 tj < 1.


562 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

Proposition 6.3. Let Ω be the type Ir,r+b domain. Let f be C 2 and K-invariant function, then<br />

for a = r j=1 tj cj ,<br />

H (1) f(a)=<br />

r<br />

Hj F(t1,...,tr)D(cj , ¯cj ) (1) , (16)<br />

j=1<br />

where the sca<strong>la</strong>r-valued operators Hj are given by<br />

Hj = 1 − t 2 <br />

2 ∂2 j<br />

∂t2 j<br />

k=j<br />

+ 1<br />

tj<br />

∂<br />

∂tj<br />

+ <br />

2 2<br />

1 − tj 1 − tk <br />

+ 2b 1 − t 2 1<br />

j<br />

tj<br />

∂<br />

.<br />

∂tj<br />

<br />

1<br />

tj − tk<br />

∂<br />

∂tj<br />

− ∂<br />

<br />

+<br />

∂tk<br />

1<br />

<br />

∂<br />

+<br />

tj + tk ∂tj<br />

∂<br />

<br />

∂tk<br />

Proof. The proof is simi<strong>la</strong>r to the proof of [3, Theorem XIII.4.7] and we will only show how<br />

one can compute the <strong>la</strong>st term of the radial part H (1)<br />

j , namely 2b(1 − t 2 j ) 1 tj<br />

∂<br />

∂tj .Let{eα} be an<br />

orthonormal basis of V consisting of the frame {cj } r j=1 , an orthonormal basis of each of the<br />

subspaces Vjk and an orthonormal basis of each of the subspaces Vj0; and <strong>le</strong>t zα = xα + iyα be<br />

the comp<strong>le</strong>x coordinates. Let f beafunctiononΩ and fix a = r k=1 tkck. Then<br />

H (1) f(a)= <br />

α,β<br />

For any X ∈ g and any v ∈ V , it is known that<br />

D (1) ∂<br />

b(a,ā)eα, ēβ<br />

2<br />

∂zα∂ ¯zβ<br />

∂Xv∂Xvf + ∂ X 2 v f = 0.<br />

f(a).<br />

We will apply this formu<strong>la</strong> for different e<strong>le</strong>ments in g. Suppose eα = eβ ∈ Vj,0. For the e<strong>le</strong>ment<br />

X = i(D(eα, ¯cj ) + D(cj , ēα)) ∈ k,wehave<br />

and<br />

Therefore<br />

which implies<br />

Xa = i D(eα, ¯cj ) + D(cj , ēα) a = iD(eα, ¯cj )a = iD(a,ēα)eα = itj eα,<br />

X 2 <br />

a = X(itjeα) =−tj D(eα, ¯cj ) + D(cj , ēα) a =−tjcj .<br />

∂itjeα ∂itjeα f(a)+ ∂−tj cj f(a)= 0,<br />

∂2 ∂y2 α<br />

f = 1 ∂<br />

F.<br />

tj ∂tj


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 563<br />

Simi<strong>la</strong>rly, for X = D(eα, ¯cj ) − D(cj , ēα) ∈ k,wehave<br />

and<br />

Hence,<br />

From this we obtain<br />

Summarizing, we find on Vj,0,<br />

Xa = D(eα, ¯cj ) − D(cj , ēα) a = tj eα,<br />

X 2 <br />

a = X(tjeα) = tj D(eα, ¯cj ) − D(cj , ēα) eα =−tjcj .<br />

∂ 2<br />

4 f =<br />

∂zα∂ ¯zβ<br />

∂tj eα∂tj eαf(a)+ ∂−tj cj f(a)= 0.<br />

∂2 ∂x2 α<br />

f = 1 ∂<br />

F.<br />

tj ∂tj<br />

0 if α = β,<br />

∂ 2<br />

∂x 2 α<br />

+ ∂2<br />

∂y2 <br />

1 ∂<br />

f = 2 F if α = β.<br />

α<br />

tj ∂tj<br />

Furthermore,<br />

D (1) 2 (1) 2<br />

b(a,ā)eα, ēα = D 1 − tj eα, ēα = 1 − tj D(eα, ēα) (1) .<br />

Hence,<br />

r<br />

<br />

j=1 eα,eβ∈Vj,0<br />

=<br />

= 2<br />

r<br />

<br />

j=1 eα∈Vj,0<br />

= 2b<br />

D (1) ∂<br />

b(a,ā)eα, ēβ<br />

2<br />

∂zα∂ ¯zβ<br />

2<br />

1 − tj D(eα, ēα) (1) 2 1<br />

r 2 1 ∂F<br />

1 − tj tj ∂tj<br />

j=1<br />

<br />

eα∈Vj,0<br />

tj<br />

f(a)<br />

∂F<br />

∂tj<br />

D(eα, ēα) (1)<br />

r 2 1 ∂F<br />

1 − tj D(cj , ¯cj )<br />

tj ∂tj<br />

(1) ,<br />

j=1<br />

since we already proved in Lemma 5.1, that<br />

<br />

D(eα, ēα) (1) = bD(cj ,cj ) (1) .<br />

This finishes the proof. ✷<br />

eα∈Vj,0


564 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

The next proposition c<strong>la</strong>ims that the Hua equation (15) for s ∈ C is sufficient for f being an<br />

eigenfunction of D(Ω) G . A simi<strong>la</strong>r result for general tube domains is proved in [12].<br />

Proposition 6.4. Let Ω be the type Ir,r+b domain. Let s ∈ C and <strong>le</strong>t λs be given by (11). Suppose<br />

f on Ω satisfies the Hua equation (15). Then f is an eigenfunction of all T ∈ D(Ω) G with<br />

eigenvalues χλs (T ).<br />

Proof. Let f be a function on Ω solution of the Hua equation. Let g ∈ G, then the function<br />

<br />

Φ(z) = f(gk· z) dk, z ∈ Ω,<br />

is a K-biinvariant solution of differential system,<br />

K<br />

Hj Φ = (r + b)s(s − 1)Φ, j = 1,...,r.<br />

Thus by a result of Yan [18], 1 Φ is proportional to the unique spherical function<br />

<br />

ϕλs (z) =<br />

K<br />

eλs (k · z) dk<br />

in M(λs), i.e.Φ(z) = cϕλs (z). It is easy to see that c = f(g· 0), then<br />

<br />

f(gk· z) dk = ϕλs (z)f (g · 0);<br />

K<br />

and consequently, by [4, Chapter IV, Proposition 2.4], f is a joint eigenfunction of all T ∈<br />

D(Ω) G with eigenvalues χλs (T ). ✷<br />

6.3. The sufficiency of the Hua equation (15)<br />

We suppose in the rest of Section 6 that s ∈ C satisfies condition (14) and that f satisfies<br />

the sufficient condition (15) in Theorem 6.1. It follows immediately from Proposition 6.4 that<br />

f ∈ M(λs), and thus by Kashiwara et al. [7], f is the Poisson transform of a function ϕ on the<br />

Furstenberg boundary G/Pmin, f = Ps(ϕ). To prove that ϕ is a function on the Shilov boundary<br />

S, we follow a method by Berline and Vergne [1] (see also [8]), the rea<strong>de</strong>r is refereed that<br />

paper for some general arguments.<br />

We need first two e<strong>le</strong>mentary <strong>le</strong>mmas for general boun<strong>de</strong>d symmetric domain Ω; the first<br />

one gives explicit formu<strong>la</strong>s for the root spaces g α , α ∈ Σ(g, a), and can easily be <strong>de</strong><strong>du</strong>ced from<br />

the Peirce <strong>de</strong>composition (see [9,15,16]). The second is essentially stated in [1] in terms of the<br />

Cay<strong>le</strong>y transform, it has however an easier form in terms of the Jordan trip<strong>le</strong> and can easily be<br />

proved using the first. To state them we need some notational preparation. Recall the quadratic<br />

1 Roughly speaking, the differential equations used in [18] is obtained from (16) by the change of coordinates xj =<br />

−t2 j /1 − t2 j .


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 565<br />

map z → Q(z) given in Section 2. For the fixed Jordan frame {cj } and the corresponding Peirce<br />

<strong>de</strong>composition (5), the map<br />

τ : z → τ(z)= Q(e)¯z,<br />

where e = c1 +···+cr, <strong>de</strong>fines a real involution of V2 and thus a real form<br />

A(e) = z ∈ V : τ(z)= z <br />

of V2; <strong>le</strong>tV2 = A(e) ⊕ iA(e) be corresponding <strong>de</strong>composition with A(e) being a real Jordan<br />

algebra. Let<br />

then A(e) = iB(e).For1 j k r, <strong>le</strong>t<br />

B(e) = z ∈ V : τ(z)=−z ,<br />

Vjk = Ajk ⊕ Bjk<br />

be the <strong>de</strong>composition of the space Vjk into real and imaginary part re<strong>la</strong>tive to the real form A(e).<br />

Lemma 6.5. The root spaces g α , α ∈ Σ(g, a), are explicitly given as follows:<br />

for 1 j,k r.<br />

g ±βj<br />

<br />

= R ξicj ∓ 2iD(cj , ¯cj ) ,<br />

g (βk−βj<br />

)/2<br />

= ξa + D(ck − cj , ā): a ∈ Ajk ,<br />

g ±(βk+βj<br />

)/2<br />

= ξb ∓ D(ck + cj , ¯b): b ∈ Bjk ,<br />

g ±βj<br />

<br />

/2<br />

= ξv ± (D(cj , ¯v) − D(v, ¯cj ): v ∈ V0j<br />

Lemma 6.6. The corresponding root spaces for the positive compact roots γk−γj<br />

2 , 1 j


566 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

therefore, system (15) implies in particu<strong>la</strong>r<br />

However,<br />

Lemma 6.7. We have<br />

H (1)<br />

k +<br />

C<br />

= <br />

k>j<br />

H (1)<br />

k +<br />

C<br />

H (1)<br />

k (γ k −γ j )/2<br />

C<br />

f = 0. (17)<br />

+<br />

γj /2<br />

(1)<br />

kC = 0.<br />

r<br />

H<br />

j=1<br />

(1)<br />

k γj /2<br />

C<br />

Proof. In<strong>de</strong>ed, using Lemma 6.6, <strong>le</strong>t D(cj , ¯v) ∈ k γj /2<br />

C , with v = ej,j+m ∈ Vj,0 (m>0). Then<br />

Hence, from (17) it follows<br />

Lemma 6.8. We have<br />

D(cj , ¯v) (1) = ej,je ∗ j,j+m = 0. ✷<br />

k +<br />

C<br />

<br />

H<br />

k>j<br />

(1)<br />

k (γk−γj )/2<br />

C<br />

(1) = <br />

k>j<br />

.<br />

f = 0. (18)<br />

k (γk−γj )/2<br />

C<br />

and the right-hand si<strong>de</strong> is a linear direct sum, namely the spaces (k (γk−γj )/2<br />

C ) (1) are linearly<br />

in<strong>de</strong>pen<strong>de</strong>nt.<br />

Proof. This follows easily from Lemma 6.6 by observing that D(ck, ¯v) ↦→ D(ck, ¯v) (1) is a linear<br />

homomorphism, and that D(ck, ¯v) (1) =¯αek,j for v = αej,k + bek,j ∈ Vj,k. ✷<br />

We conclu<strong>de</strong> from (18) and the above <strong>le</strong>mma that<br />

for any positive compact root (γk − γj )/2.<br />

Let Ψ +,(i)<br />

c<br />

H (1)<br />

k (γk−γj )/2<br />

C<br />

(1),<br />

f = 0 (19)<br />

be the set positive compact roots in k (i)<br />

C ,fori = 1, 2. Then (19) implies<br />

H β f = 0 forβ ∈ Ψ +,(1)<br />

c<br />

with β ≡ γk − γj<br />

2<br />

(k > j),


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 567<br />

where H β is the component of H given by<br />

H β = <br />

[Eβ,vα]¯vα<br />

α∈Ψ + n<br />

and Eβ is the root vector of β.<br />

Now we fix for the rest of this section β ∈ Ψ +,(1)<br />

c such that<br />

β| t −<br />

2<br />

C = γj − γj−1<br />

The root vector Eβ has the form Eβ = D(cj , ¯w) with w = ej,j−1 or w = ej−1,j being one of the<br />

basis vectors {vα}. Observe that [Eβ,vα]=0 un<strong>le</strong>ss α is in the set Ψ1 ∪ Ψ2 ∪ Ψ3 where<br />

<br />

Ψ1 = α ∈ Ψ + n : α| t −<br />

<br />

Ψ2 = α ∈ Ψ + n : α| t −<br />

2<br />

C = γk + γj−1<br />

2<br />

C = γk + γj−1<br />

<br />

Ψ3 = α ∈ Ψ + n : α| t −<br />

<br />

γj−1<br />

= .<br />

C 2<br />

Consi<strong>de</strong>r the Poincaré–Birkhoff–Witt <strong>de</strong>composition<br />

and <strong>le</strong>t π be the projection<br />

.<br />

<br />

,k j − 1 ,<br />

<br />

,k j ,<br />

U(gC) = U(gC)kC + U(aC + n −<br />

C )<br />

π : U(gC) = U(gC)kC + U(aC + n −<br />

C ) → U(aC + n −<br />

C ).<br />

The function f is now viewed as a function on G = NAK, and the group A will be i<strong>de</strong>ntified<br />

as (R + ) r . Un<strong>de</strong>r this i<strong>de</strong>ntification, f satisfies, furthermore, the equation<br />

R π H β f = 0,<br />

where R is the mapping from U(aC + nC) to differential operators on NA <strong>de</strong>fined by<br />

∂<br />

R(ξk) = tk , R(X−α) = t<br />

∂tk<br />

α X−α,<br />

for ξk ∈ a, 1 k r, and X−α ∈ n i<strong>de</strong>ntified with the corresponding <strong>le</strong>ft-invariant differential<br />

operator.<br />

We will prove that operator t − 1 2 (βj −βj−1) R(π(H β )) has analytic coefficient near t = 0 and<br />

study the in<strong>du</strong>ced equation of t − 1 2 (βj −βj−1) R(π(H β ))f = 0.


568 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

6.3.1. The projection of the Hua operator in the PBW-<strong>de</strong>composition<br />

We will compute the Poincaré–Birkhoff–Witt components of the Hua operators as an e<strong>le</strong>ment<br />

in the universal algebra of gC. Let now Ω be a general boun<strong>de</strong>d symmetric domain.<br />

Lemma 6.9. The Iwasawa <strong>de</strong>composition of v ∈ p + and ¯v ∈ p − in gC = aC + n −<br />

C + kC is given<br />

as follows.<br />

(1) For v ∈ Vkj , r k>j 1,<br />

v = ζv + ζ ′ v − D(v, ¯ck), ¯v = η ¯v + η ′ ¯v − D(ck, ¯v),<br />

where ζv,η¯v ∈ g −(βk−βj )/2<br />

C , ζ ′ v ,η′ ¯v ∈ g−(βk+βj )/2<br />

C are given by<br />

(2) For v = cj ∈ Vjj , 1 j r,<br />

with<br />

(3) For v ∈ Vj0, 1 j r,<br />

with<br />

ζv = 1<br />

v − τ(v)+ D(v, ¯ck −¯cj )<br />

2<br />

,<br />

ζ ′ 1<br />

v = v + τ(v)+ D(v, ¯cj +¯ck)<br />

2<br />

,<br />

η ¯v = 1<br />

¯v − τ(v)+ D(cj − ck, ¯v)<br />

2<br />

,<br />

η ′ 1<br />

¯v = τ(v)+¯v + D(cj + ck, ¯v)<br />

2<br />

.<br />

cj = 1<br />

2 ξj − 1<br />

2 ζj − D(cj , ¯cj ), ¯cj =− 1<br />

2 ξj − i<br />

2 ζj − D(cj , ¯cj ),<br />

ζj = i ξicj + 2D(cj , ¯cj ) .<br />

v = ζv − D(v, ¯cj ), ¯v = η ¯v − D(cj , ¯v),<br />

ζv = v + D(v, ¯cj ) ∈ n −βj /2<br />

C , η¯v =¯v + D(cj , ¯v) ∈ n −βj /2<br />

C<br />

We <strong>de</strong>note by π n 0 C the projection onto the nilpotent subalgebra<br />

n 0 C<br />

= <br />

k>j1<br />

g −(βk−βj )/2<br />

C<br />

.


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 569<br />

in the Iwasawa <strong>de</strong>composition of gC,<br />

Then, it follows from Lemma 6.9,<br />

gC = kC + aC + <br />

kj0<br />

πn0 ( ¯v) =−π<br />

C<br />

n0 C<br />

g −(βk+βj )/2<br />

C + n 0 C .<br />

τ(v) , (20)<br />

which we will need in the next proposition.<br />

Return back to type Ir,r+b domains. We compute now the projection π(H β ) of H β . Recall<br />

that the β-root vector is Eβ = D(cj , ¯w), with w = ej,j−1 or w = ej−1,j .<br />

Proposition 6.10. The projection π(H β ) is given by<br />

where the <strong>la</strong>st term<br />

π H β j−2<br />

=<br />

<br />

k=1 vα∈Vk,j−1<br />

(ζ{cj ¯wvα} + ζ ′ {cj ¯wvα} )(η ¯vα + η′ ¯vα ) + jη¯w + jη ′ ¯w<br />

+ (ζτ(w) + ζ ′ τ(w) )<br />

<br />

− 1<br />

2 ξj−1 − i<br />

2 ζj−1<br />

<br />

+<br />

r <br />

(ζ{cj ¯wvα} + ζ ′ {cj ¯wvα} )(η ¯vα + η′ ¯vα )<br />

+<br />

k=j+1 vα∈Vk,j−1<br />

<br />

1<br />

2 ξj − 1<br />

2 ζj<br />

<br />

(η ¯w + η ′ <br />

¯w ) +<br />

vα∈Vj−1,0<br />

<br />

b(η ¯w + η<br />

Jβ =<br />

′ ¯w ) if w = ej−1,j ,<br />

0 if w = ej,j−1,<br />

ζ{cj ¯wvα}η ¯vα<br />

and where the sum <br />

vα∈Vkj is taken over the orthonormal basis {vα}of Vk,j.<br />

Proof. We compute the projection <br />

α∈Ψ + n π([Eβ,vα]v∗ α ). For α ∈ Ψ + n<br />

[Eβ,vα]=0 un<strong>le</strong>ss α ∈ Ψ1 ∪ Ψ2 ∪ Ψ3.<br />

Case I. α ∈ Ψ1, with α| −<br />

tC = (γk + γj−1)/2. Then vα ∈ Vk,j−1 and<br />

vβ+α := [Eβ,vα]= <br />

D(ci, ¯w),vα = D(ci, ¯w)vα ∈ Vj,k.<br />

By the previous <strong>le</strong>mma, for k


570 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

[Eβ,vα]v ∗ α = vβ+α ¯vα ≡ (ζvβ+α + ζ ′ vβ+α )(η ¯vα + η′ ¯vα ) − <br />

D(vβ+α,cj ), ¯vα<br />

= (ζvβ+α + ζ ′ vβ+α )(η ¯vα + η′ ¯vα ) + D(cj , ¯vβ+α)vα.<br />

To find the <strong>la</strong>st term we note first that for any Jordan trip<strong>le</strong> system [9],<br />

<br />

D(vα, ¯vα), D(w, ¯cj ) = D <br />

vα, D(cj , ¯w)vα − D D(w, ¯cj )vα, ¯vα ;<br />

we <strong>le</strong>t it act on cj and then sum over vα<br />

<br />

vα∈Vk,j−1<br />

D cj ,D(cj , ¯w)vα<br />

<br />

vα = a <br />

D(cj−1,cj−1) + D(ck,ck)<br />

2<br />

w<br />

by using Proposition 5.1. It is further w, since a = 2 for type I domains. Thus<br />

<br />

≡<br />

<br />

(ζvβ+α + ζ ′ vβ+α )(η ¯vα + η′ ¯vα ) + (j − 2)η ¯w + (j − 2)η ′ ¯w .<br />

vα∈Vk,j−1<br />

k


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 571<br />

Now <strong>le</strong>t k = j, then [Eβ,vα]=D(cj , ¯w)vα = D(vα, ¯w)cj =〈vα,w〉cj , which vanishes except<br />

when vα = w and in that case,<br />

and<br />

<br />

1<br />

[Eβ,vα]¯vα = cj ¯vα =<br />

2 ξj − 1<br />

2 ζj<br />

<br />

− D(cj , ¯cj ) ¯w,<br />

<br />

1<br />

[Eβ,vα]¯vα ≡<br />

2 ξj − 1<br />

2 ζj<br />

<br />

(η ¯w + η ′ ¯w ) + η ¯w + η ′ ¯w .<br />

1<br />

Case III. α ∈ Ψ3, with α| −<br />

t =<br />

C 2γj−1, and the root vector vα ∈ Vj−1,0. In this case, we have<br />

[Eβ,vα]¯vα = D(cj , ¯w)vα ¯vα ≡ ζD(cj , ¯w)vα − D <br />

D(cj , ¯w)vα,cj η ¯vα<br />

≡ ζD(cj , ¯w)vαη ¯vα + D <br />

cj , D(cj ,w)vα vα.<br />

However, by the commutator re<strong>la</strong>tion (JP15) in [9] we have<br />

<br />

D(w, ¯vα), D(cj , ¯cj ) = D <br />

D(w, ¯vα)cj , ¯cj − D cj , D(vα, ¯w)cj =−D cj , D(vα, ¯w)cj<br />

since D(w, ¯vα)cj = 0 by the Peirce ru<strong>le</strong> that D(w, ¯vα)cj ∈{Vj,j−1 ¯Vj−1,0Vjj }={0}, thus<br />

D <br />

cj , D(cj , ¯w)vα vα = D(cj ,cj ), D(w, ¯vα) vα = D(cj ,cj )D(w, ¯vα)vα = D(vα, ¯vα)w<br />

since D(cj ,cj )vα = 0.<br />

It is easy to see, by direct matrix computation, that,<br />

Hence, mo<strong>du</strong>lo U(gC)kC<br />

Consequently,<br />

<br />

<br />

vα∈Vj−1,0<br />

<br />

vα∈Vj−1,0<br />

[Eβ,vα]v<br />

vα∈Vj−1,0<br />

∗ α<br />

and this finishes the proof. ✷<br />

<br />

b ¯w if w = ej−1,j ,<br />

D(vα, ¯vα)w =<br />

0 ifw = ej,j−1.<br />

<br />

b(η ¯w + η<br />

D(vα, ¯vα)w ≡<br />

′ ¯w ) if w = ej−1,j ,<br />

0 if w = ej,j−1.<br />

≡ <br />

vα<br />

ζD(cj , ¯w)vαη ¯vα +<br />

<br />

b(η ¯w + η ′ ¯w ) if w = ej−1,j ,<br />

0 if w = ej,j−1


572 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

6.3.2. The in<strong>du</strong>ced equation<br />

We apply now the theory of boundary values of eigenfunctions of D(Ω) G on symmetric<br />

spaces, see [7,10,11,13].<br />

We i<strong>de</strong>ntify the space G/K with NA and A with (R + ) r . It follows from Proposition 6.10 that<br />

operator t − 1 2 (βj −βj−1) R[π(H β )] has analytic coefficients near t = 0. Then the in<strong>du</strong>ced equation<br />

for the differential equation t − 1 2 (βj −βj−1) R[π(H β )]f = 0is<br />

where t = (t1,t2,...,tr) ∈ A = (R + ) r , and<br />

lim<br />

t→0 tλ−ρ t − 1 2 (βj<br />

−βj−1) β<br />

R π H t ρ−λ (Bλf)= 0,<br />

t μ = t μ(ξ1)<br />

1 ···t μ(ξr )<br />

r<br />

for μ ∈ a ∗ C .HereBλf is the boundary value of f .<br />

Proposition 6.11. The boundary value Bλf of f satisfies the following in<strong>du</strong>ced equation:<br />

R[ζτ(w)](Bλf)= 0. (21)<br />

Observe, using (20), that the in<strong>du</strong>ced Eq. (21) is equiva<strong>le</strong>nt to the following one:<br />

R[η ¯w](Bλf)= 0.<br />

Proof. Let us compute the limit of the differential operator<br />

t λ−ρ t − 1 2 (βj −βj−1) R π H β t ρ−λ<br />

when t → 0. We will consi<strong>de</strong>r each term in the projection π(H β ).<br />

• The differential operator corresponding to j(η¯w + η ′ ¯w ) is<br />

t λ−ρ t − 1 2 (βj −βj−1) j t 1 2 (βj −βj−1) R(η ¯w) + t 1 2 (βj +βj−1) R(η ′ ¯w ) t ρ−λ ,<br />

and its limit when t ↦→ 0is<br />

jR(η ¯w). (22)<br />

• Consi<strong>de</strong>r the quadratic term (ζτw + ζ ′ 1<br />

τw )(− 2ξj−1 − i<br />

2ζj−1). The corresponding differential<br />

operator is<br />

t λ−ρ t − 1 2 (βj<br />

<br />

t 1<br />

−βj−1)<br />

2 (βj −βj−1)<br />

R(ζτ(w) + t 1 2 (βj +βj−1) ′<br />

R(η τ(w) ) <br />

<br />

× − 1<br />

2 tj−1<br />

∂<br />

∂tj−1<br />

− i<br />

2 tβj−1 <br />

R(ζj−1) t ρ−λ<br />

× R(ζτ(w)) + t βj−1 R(ζ ′ τ(w) ) <br />

− 1<br />

2 (ρ − λ)(ξj−1) − i<br />

2 tβj−1 R(ζj−1)<br />

<br />

.


Its limit when t → 0is<br />

K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 573<br />

− 1<br />

(ρ − λ)(ξj−1)R(ζτ(w)).<br />

2<br />

• For the quadratic term ( 1 2ξj − 1 2ζj )(η ¯w − η ′ ¯w ), the corresponding differential operator is<br />

t λ−ρ t − βj −β <br />

j−1 1<br />

2<br />

2 tj<br />

Its limit is<br />

which is<br />

lim<br />

t→0 tλ−ρ t − βj −βj−1 2<br />

= lim<br />

∂<br />

∂tj<br />

t<br />

t→0 λ−ρ t − βj −βj−1 2<br />

− 1<br />

2 tβj <br />

t βj −βj−1 R(ζj ) 2 R(η ¯w) + t βj +βj−1 2 R(η ′ ¯w )t<br />

ρ−λ .<br />

<br />

1 ∂ βj −βj−1 t 2 t<br />

2 ∂tj<br />

ρ−λ R(η ¯w) + t βj +βj−1 2 t ρ−λ R(η ′ ¯w )<br />

<br />

1 βj − βj−1<br />

(ξj ) + (ρ − λ)(ξj ) t<br />

2 2<br />

βj −βj−1 2 t ρ−λ R(η ¯w)<br />

+ 1<br />

<br />

βj + βj−1<br />

(ξj ) + (ρ − λ)(ξj ) t<br />

2 2<br />

βj +βj−1 2 t ρ−λ R(η ′ ¯w )<br />

<br />

,<br />

1<br />

1 + (ρ − λ)(ξj )<br />

2<br />

R(η ¯w).<br />

• The in<strong>du</strong>ced equation corresponding to the <strong>la</strong>st term of the projection π(Hβ ) is<br />

<br />

bR(η ¯w),<br />

(23)<br />

0.<br />

• Now, it is easy to see, using the same computations, that the in<strong>du</strong>ced equation of the remaining<br />

terms of π(Hβ ) is zero.<br />

It follows now from (22), (23) and (20), that the boundary value Bλf of f satisfies<br />

where C1 is given by<br />

C1R(ζτ(w))(Bλf)= 0,<br />

C1 = 1<br />

2 (ρ − λ)(ξj<br />

<br />

12<br />

+ j + b,<br />

− ξj−1) +<br />

1<br />

2 + j.<br />

Now if C1 = 0, then the in<strong>du</strong>ced equation is R(ζτ(w))(Bλf)= 0. On the other hand, if C1 = 0,<br />

we may rep<strong>la</strong>ce f by tκ(β j −βj−1 2 ) f for sufficiently <strong>la</strong>rge κ>0, consi<strong>de</strong>r the differential operator<br />

t − 1 2 (βj −βj−1) κ(<br />

t βj −βj−1 2<br />

) β<br />

R π H t −κ(β j −βj−1 2 )<br />

,<br />

and we still prove that R(ζτ(w))(Bλf)= 0, see also [12]. ✷


574 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

We continue the proof of the necessity condition of the Hua equation. We now get<br />

R(ζτ(w))(Bλs f) = 0 for any root β such that β ≡ 1 2 (γj − γj−1). Since { 1 2 (γj − γj−1),<br />

2 j r} is the set of simp<strong>le</strong> roots of the system { 1 2 (γk − γj ), 1 j


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 575<br />

Let f be in M(λs) and suppose f satisfies (25). Then f = Ps(ϕ) is the Poisson transform of a<br />

hyperfunction ϕ on S.<br />

7.2. The necessity of the Hua equation (25)<br />

Consi<strong>de</strong>r the operator<br />

Proposition 7.3. Let s ∈ C. We have<br />

and<br />

In particu<strong>la</strong>r, for σ = 0,(p+ b)/2,<br />

Y = U − −2σ 2 + 2pσ + c<br />

σ(2σ − p − b) W.<br />

WPs,u(z) = Ps,u(z)σ 2 (2σ − p − b)u<br />

UPs,u(z) = Ps,u(z)σ −2σ 2 + 2pσ + c u.<br />

YPs,u(z) = 0,<br />

and the image f = Ps(ϕ) of the Poisson transform of a hyperfunction ϕ on S satisfies<br />

Yf = 0.<br />

Proof. We compute first W on Ps(z, u). By the covariant property of W and transformation<br />

property (10) of the kernel Ps(z, u) we need only to prove that the formu<strong>la</strong> is valid at z = 0. We<br />

use the differentiation h(z, z)¯∂ in p<strong>la</strong>ce of ¯∂ as it will pro<strong>du</strong>ce some more compact formu<strong>la</strong>s; the<br />

eventual result will be the same at z = 0. (The operator h(z, z)¯∂ can be geometrically <strong>de</strong>fined,<br />

see Section 4.) Proceeding as in the proof of Theorem 5.3 by using Lemma 5.2, we have<br />

<br />

h(z, z)¯∂ ∂Ps(z, u) = σ 2 Ps,u(z) b(z, z) z ¯z − u ¯z ⊗ ¯z z −ū z − σPs,u(z)Id.<br />

Its image un<strong>de</strong>r −Ad V ⊗ ¯V →kC is,<br />

−Ad V ⊗ ¯V →kC<br />

since<br />

and<br />

h(z, z)¯∂ ∂Ps(z, u) = σ 2 Ps,u(z)D b(z, z) z ¯z − u ¯z , ¯z z −ū z − σPs,u(z)pZ0,<br />

−AdV ⊗ ¯V →kC<br />

(u ⊗¯v) =−[u, ¯v]=D(u, ¯v)<br />

−AdV ⊗ ¯V →kC<br />

Id = pZ0.<br />

To compute ¯∂Ad V ⊗ ¯V →kC (h(z, z)¯∂)∂Ps(z, u) at z = 0 we observe that for any function f ,<br />

¯∂f(0) is the coefficient of ¯z in the expansion of f near z = 0. By direct computation we find:


576 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

−¯∂AdV ⊗ ¯V<br />

¯∂∂Ps(z, →kC<br />

u)|z=0 = σ 3 u ⊗ D(u,ū) − pσ 2 u ⊗ Z0<br />

+ σ<br />

2 <br />

j<br />

vj ⊗ D Q(u) ¯vj , ū − D(u, ¯vj ) ,<br />

where {vj } is an orthonormal basis of V .Nowforv ∈ V,X∈ kC, AdV ⊗kC→(v ⊗ X) =[v,X]=<br />

−Xv, where Xv is the <strong>de</strong>fining action of kC on V . We get<br />

AdV ⊗kC→V ¯∂AdV ⊗ ¯V<br />

¯∂∂Ps(z, →kC<br />

u)|z=0<br />

= 2σ 3 u − pσ 2 <br />

2<br />

u + σ D Q(u) ¯vj , ū <br />

vj − D(u, ¯vj )vj .<br />

j<br />

The sum can further be evaluated by using the Peirce <strong>de</strong>composition with respect to u, and we<br />

obtain eventually<br />

AdV ⊗kC→V ¯∂AdV ⊗ ¯V<br />

¯∂∂Ps(z, →kC<br />

u)|z=0 = σ 2 (2σ − p − b)u.<br />

This proves the first formu<strong>la</strong>.<br />

For the second formu<strong>la</strong> we have first<br />

and<br />

¯∂Ps(z, u) = σPs(z, u)b(z, z) z ¯z − u ¯z<br />

∂ ¯∂ ¯∂Ps(z, u)(0)|z=0 = σ <br />

j,k<br />

Performing the differentiation we find then<br />

So that<br />

∂vk ¯∂vj<br />

∂ ¯∂ ¯∂Ps(z, u)|z=0 = σ 3 ū ⊗ u ⊗ u − σ<br />

Ps(z, u)b(z, z) z ¯z − u ¯z |z=0.<br />

2 <br />

− σ <br />

¯vk ⊗ vj ⊗ D(vk, ¯vj )u.<br />

j,k<br />

UPs(z, u)|z=0 =−σ 3 D(u,ū)u + σ<br />

k<br />

2 <br />

+ σ <br />

D(vj , ¯vk)D(vk, ¯vj )u.<br />

Again <br />

k D(vk, ¯vk)v = pv for v ∈ V and<br />

<br />

D(vj , ¯vk)D(vk, ¯vj )u = cp<br />

by [19, Lemma 2.5] with c given as in (24). ✷<br />

j,k<br />

j,k<br />

k<br />

¯vk ⊗ (vk ⊗ u + u ⊗ vk) <br />

<br />

D(vk, ¯vk)u + D(u, ¯vk)vk


K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 577<br />

In particu<strong>la</strong>r, if s = 1, WPs,u(z) = 0; the simi<strong>la</strong>r result, VPs,u(z) = 0 for the Hua operator V<br />

was proved by Berline and Vergne [1, Proposition 3.3].<br />

7.3. The sufficiency of the Hua equation (25)<br />

The i<strong>de</strong>a of the proof is simi<strong>la</strong>r to that in Section 6, and many technical computations on the<br />

various <strong>de</strong>composition involving the third-or<strong>de</strong>r Hua operators W and U are paral<strong>le</strong>l to those<br />

in [1] for the Berline–Vergne’s Hua operator V, so we will not present all <strong>de</strong>tails.<br />

Suppose hereafter that f ∈ M(λs) satisfies (25). We first observe that the operator U can also<br />

be written as<br />

U = <br />

vγ v ∗ αv∗ β ⊗ vα, [vβ,v ∗ γ ]<br />

α,β,γ<br />

since [[v ∗ γ ,vα],vβ]=[vα, [vβ,v ∗ γ ]] by the Jacobi i<strong>de</strong>ntity and by [vα,vβ]=0. Writing<br />

with<br />

U δ ⊗ vδ = <br />

α+β−γ =δ<br />

we have, mo<strong>du</strong>lo U(gC)kC,<br />

U = U δ ⊗ vδ, W = W δ ⊗ vδ,<br />

vγ v ∗ α v∗ β ⊗ vδ, W δ ⊗ vδ = <br />

U δ ⊗ vδ − W δ <br />

<br />

⊗ vδ =<br />

α+β−γ =δ<br />

α+β−γ =δ<br />

|Cα,β,γ | 2<br />

<br />

v ∗ δ ⊗ vδ,<br />

v ∗ α v∗ β vγ ⊗ vδ, (26)<br />

where Cα,β,γ are given by [vα, [vβ,v ∗ γ ]] = Cα,β,γ vδ.<br />

Writing Y = <br />

δ Yδ ⊗ vδ as above with Y δ ∈ U(g) we have then mo<strong>du</strong>lo U(gC)kC<br />

with<br />

Thus<br />

C1 := <br />

α+β−γ =δ<br />

Y δ = C1v ∗ δ<br />

δ + C2W<br />

|Cα,β,γ | 2 , C2 := 1 − −2σ 2 + 2pσ + c<br />

σ(2σ − p − b) .<br />

Y δ f = 0. (27)<br />

for any non-compact root δ ≡ (γj + γj−1)/2 mo<strong>du</strong>lo t −<br />

C , by our assumption on f . We will henceforth<br />

fix one such δ, and study the in<strong>du</strong>ced equation of (27).


578 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

Recall projection π from U(gC) onto U(aC + n −<br />

C ). Here we will not be ab<strong>le</strong> to find explicit<br />

formu<strong>la</strong> for π(Yδ ) as in Proposition 6.10. Neverthe<strong>le</strong>ss we can compute the in<strong>du</strong>ced equation.<br />

Consi<strong>de</strong>r the <strong>de</strong>composition of U(aC + n −<br />

C ) un<strong>de</strong>r aC,<br />

where<br />

is root <strong>la</strong>ttice of Σ − (g, a).<br />

U(aC + n −<br />

<br />

C ) =<br />

p∈Π −<br />

U(aC + n −<br />

C )p, (28)<br />

Π − <br />

= p = <br />

β∈Σ − (g,a)<br />

<br />

cββ, 0 cβ, cβ∈ Z<br />

Lemma 7.4. Let α + β − γ = δ ≡ (γj + γj−1)/2 be as in (26). Decomposing π(¯vα ¯vβvγ ) ∈<br />

U(aC + n −<br />

C ) according to (28),<br />

π(¯vα ¯vβvγ ) = <br />

π(¯vα ¯vβvγ )p, (29)<br />

p∈Π −<br />

we have that p −(βj − βj−1)/2, for any p appearing in (29) so that π(¯vα ¯vβvγ )p = 0.<br />

Proof. The <strong>le</strong>mma can be proved by a case by case computation of the projection by using<br />

Lemma 6.9, and is essentially contained in [1]. We sketch another somewhat more systematic<br />

method. We <strong>de</strong>note the Iwasawa <strong>de</strong>composition as ¯vα = π(¯vα) + y with y ∈ kC. Thus<br />

The Iwasawa projection of the first term is<br />

¯vα ¯vβvγ = π(¯vα) ¯vβvγ + y ¯vβvγ .<br />

π(¯vα)π( ¯vβvγ ).<br />

The projection of the second term is<br />

π <br />

[y, ¯vβ]vγ + π ¯vβ[y,vγ ] .<br />

Observe by Lemma 6.9 that the e<strong>le</strong>ment y is a positive compact root vector, so that all these projections<br />

involved are of the form π(¯vδvɛ) with vδ and vɛ being non-compact positive root vectors.<br />

Our <strong>le</strong>mma re<strong>du</strong>ces to the following c<strong>la</strong>im, which can be proved easily by using Lemma 6.9. The<br />

weights p of π(¯vδvɛ) satisfy the inequalities<br />

p − βj − βj−1<br />

2<br />

+ βk − βk ′<br />

2<br />

p − βj − βj−1<br />

2<br />

if δ − ɛ = γj + γj−1<br />

2<br />

if δ − ɛ = γj + γj−1<br />

2<br />

− γk + γk ′<br />

2<br />

− γk,<br />

with k>k ′ ,


and<br />

K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580 579<br />

p − βj − βj−1<br />

2<br />

+ βk<br />

2<br />

if δ − ɛ = γj + γj−1<br />

2<br />

− γk<br />

. ✷<br />

2<br />

From this it follows that the operator t − 1 2 (βj −βj−1) R[π(Y δ )] has analytic coefficients in t near<br />

t = 0 and thus the in<strong>du</strong>ced equation<br />

lim<br />

t→0 tλ−ρ t 1 2 (βj −βj−1) R π Y δ t ρ−λ (Bλf)= 0 (30)<br />

of the equation Yδf = 0 is well <strong>de</strong>fined.<br />

Consi<strong>de</strong>r next the eigenspace <strong>de</strong>composition of the space n −<br />

rj=1 ξj :<br />

1<br />

2<br />

with<br />

n −1<br />

C<br />

= <br />

kj1<br />

and correspondingly<br />

Let<br />

n −<br />

C<br />

= n−1<br />

C<br />

g −(βk+βj )/2<br />

C , n −1/2<br />

C<br />

U(aC + n −<br />

C ) = U(aC + n −<br />

+ n−1/2<br />

C + n0 C<br />

<br />

=<br />

k1<br />

C )n −1<br />

C<br />

π Y δ = Y1 + Y0<br />

g −βk/2<br />

C , n 0 C<br />

C un<strong>de</strong>r the e<strong>le</strong>ment 1 2 ξc =<br />

= <br />

kj<br />

g −(βk−βj )/2<br />

C ;<br />

<br />

+ n−1/2<br />

C + U aC + n 0 <br />

C . (31)<br />

be the <strong>de</strong>composition of π(Yδ ) according to (31). As the <strong>de</strong>compositions (28) and (31) are consistent,<br />

we see that the weights p that appear in the <strong>de</strong>composition of Y1 according to (28) satisfy<br />

p −δ, and p μ for μ such that n μ<br />

C ⊂ n−1<br />

C + n−1/2<br />

C . The first implies that the in<strong>du</strong>ced equation<br />

is well <strong>de</strong>fined, and the second that the in<strong>du</strong>ced Eq. (30) now re<strong>du</strong>ces to<br />

lim<br />

t→0 tλ−ρ t − 1 2 (βj −βj−1) R[Y1]t ρ−λ (Bλf)= 0. (32)<br />

The e<strong>le</strong>ment Y0 can be found along the same lines as in [1], where the constant term C1ζvδ<br />

was found.<br />

Lemma 7.5. The e<strong>le</strong>ment Y0 is given by<br />

<br />

1<br />

2<br />

C1 + C2 −ξj − ξ<br />

2<br />

2 j−1 + ξj<br />

<br />

′<br />

ξj−1 + C 1ξ ζvδ ,<br />

where ξ ∈ aC, C ′ 1 is some constants in<strong>de</strong>pen<strong>de</strong>nt of λ.


580 K. Koufany, G. Zhang / Journal of Functional Analysis 236 (2006) 546–580<br />

In particu<strong>la</strong>r the in<strong>du</strong>ced Eq. (32) is of the form<br />

<br />

C1 + C2D(λ) R(ζvδ )(Bλf)= 0,<br />

where<br />

D(λ) = 1<br />

−(ρ − λ)(ξj )<br />

2<br />

2 − (ρ − λ)(ξj−1) 2 + (ρ − λ)(ξj )(ρ − λ)(ξj−1) <br />

+ C ′ 1 (ρ − λ)(ξ).<br />

Observe first that C1 > 0. If C2 = 0 it follows immediately that R(ζvδ )(Bλf)= 0, so we need<br />

only to consi<strong>de</strong>r the case C2 = 0. If C1 + C2D(λ) = 0 we get again R(ζvδ )(BλF) = 0. Finally,<br />

if C1 + C2D(λ) = 0 we may rep<strong>la</strong>ce f by t κγj for sufficiently <strong>la</strong>rge κ and still prove<br />

that R(ζvδ )(Bλf)= 0; see [12]. This comp<strong>le</strong>tes the proof.<br />

References<br />

[1] N. Berline, M. Vergne, Équations <strong>de</strong> Hua et noyau <strong>de</strong> Poisson, in: Noncommutative Harmonic Analysis and Lie<br />

Groups, Marseil<strong>le</strong>, 1980, in: Lecture Notes in Math., vol. 880, Springer, Berlin, 1981, pp. 1–51.<br />

[2] M. Englis, J. Peetre, Covariant Lap<strong>la</strong>cian operators on Käh<strong>le</strong>r manifolds, J. Reine Angew. Math. 478 (1996) 17–56.<br />

[3] J. Faraut, A. Korányi, Analysis on Symmetric Cones, Oxford Math. Monogr., C<strong>la</strong>rendon Press, Oxford, 1994.<br />

[4] S. Helgason, Groups and Geometric Analysis. Integral Geometry, Invariant Differential Operators, and Spherical<br />

Functions, Pure Appl. Math., vol. 113, Aca<strong>de</strong>mic Press, Or<strong>la</strong>ndo, FL, 1984.<br />

[5] L.K. Hua, Harmonic Analysis of Functions of Several Comp<strong>le</strong>x Variab<strong>le</strong>s in the C<strong>la</strong>ssical Domains, Amer. Math.<br />

Soc., Provi<strong>de</strong>nce, RI, 1963, iv+164 pp.<br />

[6] K. Johnson, A. Korányi, The Hua operators on boun<strong>de</strong>d symmetric domains of tube type, Ann. of Math. (2) 111<br />

(1980) 589–608.<br />

[7] M. Kashiwara, A. Kowata, K. Minemura, K. Okamoto, T. Oshima, M. Tanaka, Eigenfunctions of invariant differential<br />

operators on a symmetric space, Ann. of Math. (2) 107 (1978) 1–39.<br />

[8] M. Lassal<strong>le</strong>, Les équations <strong>de</strong> Hua d’un domaine borné symétrique <strong>du</strong> type tube, Invent. Math. 77 (1984) 129–161.<br />

[9] O. Loos, Boun<strong>de</strong>d symmetric domains and Jordan pairs, University of California, Irvine, 1977.<br />

[10] T. Oshima, A <strong>de</strong>finition of boundary values of solutions of partial differential equations with regu<strong>la</strong>r singu<strong>la</strong>rities,<br />

Publ. Res. Inst. Math. Sci. 19 (1983) 1203–1230.<br />

[11] T. Oshima, Boundary value prob<strong>le</strong>ms for systems of linear partial differential equations with regu<strong>la</strong>r singu<strong>la</strong>rities,<br />

in: Adv. Stud. Pure Math., vol. 4, 1984, pp. 391–432.<br />

[12] N. Shimeno, Boundary value prob<strong>le</strong>ms for the Shilov boundary of a boun<strong>de</strong>d symmetric domain of tube type,<br />

J. Funct. Anal. 140 (1996) 124–141.<br />

[13] N. Shimeno, Boundary value prob<strong>le</strong>ms for various boundaries of Hermitian symmetric spaces, J. Funct. Anal. 170<br />

(2000) 265–285.<br />

[14] G. Shimura, Differential operators, holomorphic projection, and singu<strong>la</strong>r forms, Duke Math. J. 76 (1994) 141–173.<br />

[15] A. Unterberger, H. Upmeier, The Berezin transform and invariant differential operators, Comm. Math. Phys. 164<br />

(1994) 563–597.<br />

[16] H. Upmeier, Jordan algebras and harmonic analysis on symmetric spaces, Amer. J. Math. 108 (1986) 1–25.<br />

[17] H. Upmeier, Jordan algebras in analysis, operator theory, and quantum mechanics, in: CBMS Reg. Conf. Ser. Math.,<br />

vol. 67, Amer. Math. Soc., Provi<strong>de</strong>nce, RI, 1987.<br />

[18] Z. Yan, A c<strong>la</strong>ss of generalized hypergeometric functions in several variab<strong>le</strong>s, Canad. J. Math. 44 (1992) 1317–1338.<br />

[19] G. Zhang, Invariant differential operators on Hermitian symmetric spaces and their eigenvalues, Israel J. Math. 119<br />

(2000) 157–185.<br />

[20] G. Zhang, Shimura invariant differential operators and their eigenvalues, Math. Ann. 319 (2001) 235–265.<br />

[21] G. Zhang, Nearly holomorphic functions and re<strong>la</strong>tive discrete series of weighted L 2 -spaces on boun<strong>de</strong>d symmetric<br />

domains, J. Math. Kyoto Univ. 42 (2002) 207–221.


Journal of Functional Analysis 236 (2006) 581–591<br />

www.elsevier.com/locate/jfa<br />

Symmetry of <strong>la</strong>rge solutions of nonlinear elliptic<br />

equations in a ball<br />

A<strong>le</strong>ssio Porretta a,1 , Laurent Véron b,∗<br />

a Dipartimento di Matematica, Università di Roma Tor Vergata, Via <strong>de</strong>l<strong>la</strong> Ricerca Scientifica 1, 00133 Roma, Italy<br />

b Laboratoire <strong>de</strong> Mathématiques et Physique Théorique, CNRS UMR 6083, Université François Rabe<strong>la</strong>is,<br />

Tours 37200, France<br />

Received 13 December 2005; accepted 2 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 2 May 2006<br />

Communicated by H. Brezis<br />

Abstract<br />

Let g be a locally Lipschitz continuous real-valued function which satisfies the Kel<strong>le</strong>r–Osserman condition<br />

and is convex at infinity, then any <strong>la</strong>rge solution of −u + g(u) = 0 in a ball is radially symmetric.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Elliptic equations; Boundary blow-up; Kel<strong>le</strong>r–Osserman condition; Radial symmetry; Spherical Lap<strong>la</strong>cian<br />

1. Intro<strong>du</strong>ction<br />

Let BR <strong>de</strong>note the open ball of center 0 and radius R>0inRN , N 2. A c<strong>la</strong>ssical result<br />

<strong>du</strong>e to Gidas, Ni and Nirenberg [9] asserts that, if g is a locally Lipschitz continuous real-valued<br />

function, any u ∈ C2 (Ω) which is a positive solution of<br />

<br />

−u + g(u) = 0 inBR,<br />

(1.1)<br />

u = 0 on∂BR<br />

is radially symmetric. The proof of this result is based on the ce<strong>le</strong>brated A<strong>le</strong>xandrov–Serrin<br />

moving p<strong>la</strong>ne method [17]. Later on, this method was used in many occasions, with a lot of<br />

* Corresponding author.<br />

E-mail address: veronl@lmpt.univ-tours.fr (L. Véron).<br />

1 The author acknow<strong>le</strong>dges the support of RTN European project FRONTS-SINGULARITIES, RTN contract HPRN-<br />

CT-2002-00274.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.010


582 A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591<br />

refinements for obtaining se<strong>le</strong>cted symmetry results and a priori estimates for solutions of <strong>semi</strong>linear<br />

elliptic equations. If the boundary condition is rep<strong>la</strong>ced by u = k ∈ R, c<strong>le</strong>arly the radial<br />

symmetry still holds if u − k does not change sign in BR. Starting from this observation, it was<br />

conjectured by Brezis [5] that any solution u of<br />

<br />

−u + g(u) = 0 inBR,<br />

(1.2)<br />

lim|x|→R u(x) =∞<br />

is in<strong>de</strong>ed radially symmetric. Notice that this prob<strong>le</strong>m admits a solution (usually cal<strong>le</strong>d a “<strong>la</strong>rge<br />

solution”) if and only if g satisfies the Kel<strong>le</strong>r–Osserman condition [10,14] g h on [a,∞), for<br />

some a>0 where h is non <strong>de</strong>creasing and satisfies<br />

∞<br />

a<br />

s<br />

ds<br />

√ < ∞, where H(s)= h(t) dt. (1.3)<br />

H(s)<br />

Up to now, at <strong>le</strong>ast to our know<strong>le</strong>dge, only partial results were known concerning the radial<br />

symmetry of solutions of (1.2): in [13], the authors prove this result assuming (besi<strong>de</strong>s the Kel<strong>le</strong>r–<br />

Osserman condition) that g ′ (s)/ √ G(s) →∞as s →∞, or for the special case when g(s) = s q ,<br />

using the estimates for the second term of the asymptotic expansion of the solution near the<br />

boundary. Of course, the symmetry can also be obtained via uniqueness, however uniqueness is<br />

known un<strong>de</strong>r an assumption of global monotonicity and convexity [11,12]. Otherwise, it is easy<br />

to prove, by a one-dimensional topological argument, that uniqueness for prob<strong>le</strong>m (1.2) holds<br />

for almost all R>0 un<strong>de</strong>r the mere monotonicity assumption. However, if g is not monotone,<br />

uniqueness may not hold (see e.g. [1,13,15]), and it turns out to be very important to know<br />

whether all the solutions constructed in a ball are radially symmetric, a fact that would <strong>le</strong>ad to<br />

a full c<strong>la</strong>ssification of all possib<strong>le</strong> solutions. Let us point out that the interest in such qualitative<br />

properties of <strong>la</strong>rge solutions has being raised in the <strong>la</strong>st few years from different prob<strong>le</strong>ms (see<br />

e.g. [1,6–8] and the references therein).<br />

In this artic<strong>le</strong> we prove that Brezis’ conjecture is verified un<strong>de</strong>r an assumption of asymptotic<br />

convexity upon g, namely we prove<br />

Theorem 1.1. Let g be a locally Lipschitz continuous function. Assume that g is positive and<br />

convex on [a,∞) for some a>0, and satisfies the Kel<strong>le</strong>r–Osserman condition. Then any C 2<br />

solution of (1.2) is radially symmetric and increasing.<br />

Notice that the Kel<strong>le</strong>r–Osserman condition implies that the function g is superlinear at infinity.<br />

The convexity assumption on g is then very natural in such context.<br />

In or<strong>de</strong>r to prove Theorem 1.1, we prove first a suitab<strong>le</strong> adaptation of Gidas–Ni–Nirenberg<br />

moving p<strong>la</strong>ne method to the framework of <strong>la</strong>rge solutions, without requiring any monotonicity<br />

assumption on g. This first result, which can have an interest in its own, reads as follows.<br />

Theorem 1.2. Assume that g is locally Lipschitz continuous and <strong>le</strong>t u be a solution of (1.2) which<br />

satisfies<br />

<br />

lim|x|→R ∂ru =∞,<br />

(1.4)<br />

|∇τ u|=o(∂ru) as |x|→R, ∀τ ⊥ x such that |τ|=1,<br />

a


A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591 583<br />

where ∂ru and ∇τ u are respectively the radial <strong>de</strong>rivative and the tangential <strong>gradient</strong> of u. Then<br />

u is radially symmetric and ∂ru>0 on BR \{0}.<br />

Thus, in view of the previous statement, our main point in or<strong>de</strong>r to <strong>de</strong><strong>du</strong>ce the general result<br />

of Theorem 1.1 is to prove that condition (1.4) always holds (even in a stronger form) if we<br />

assume that g is asymptotically convex, and this is achieved by providing sharp informations on<br />

the radial and the tangential behaviour of u near the boundary.<br />

2. Proof of the results<br />

Let B ={e1,...,eN } be the canonical basis of RN .IfP ∈ RN and ρ>0, we <strong>de</strong>note by<br />

Bρ(P ) the open ball with center P and radius ρ, and for simplicity Bρ(0) = Bρ. We consi<strong>de</strong>r the<br />

prob<strong>le</strong>m<br />

<br />

−u + g(u) = 0 inBR,<br />

(2.1)<br />

u(x) =∞ on ∂BR,<br />

where R>0. By a solution of (2.1), we mean that u ∈ C 2 (BR) is a c<strong>la</strong>ssical solution in the<br />

interior of the ball and that u(x) tends to infinity uniformly as |x| tends to R.<br />

We shall consi<strong>de</strong>r the following assumptions on g:<br />

and satisfies<br />

g : R → R is locally Lipschitz continuous, (2.2)<br />

∃a >0 such that g is positive and convex on [a,∞), (2.3)<br />

<br />

+∞<br />

a<br />

t<br />

1<br />

√ dt < +∞, where G(t) = g(s)ds. (2.4)<br />

G(t)<br />

Note that convexity and (2.4) imply that g is increasing on [b,∞) for some b>0.<br />

If u ∈ C 1 (BR) we <strong>de</strong>note by ∂u/∂r(x) =〈Du(x), x/|x|〉 the radial <strong>de</strong>rivative of u, and by<br />

∇τ u(x) = (Du(x) −|x| −1 ∂u/∂r(x))x the tangential <strong>gradient</strong> of u. Our first technical result,<br />

which is a reformu<strong>la</strong>tion in the framework of <strong>la</strong>rge solutions of the famous original proof of<br />

Gidas, Ni and Nirenberg [9], is the following.<br />

Theorem 2.1. Assume that g satisfies (2.2), and <strong>le</strong>t u be a solution of (2.1). If there holds<br />

(ii)<br />

then u is radially symmetric and ∂u/∂r > 0 in BR \{0}.<br />

a<br />

∂u<br />

(i) lim (x) =∞,<br />

|x|→R ∂r<br />

<br />

∇τ (x) <br />

<br />

∂u<br />

= o<br />

∂r (x)<br />

<br />

as |x|→R, (2.5)


584 A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591<br />

Proof. Since the equation is invariant by rotation, it is sufficient to prove that (2.5) implies that<br />

u is symmetric in the x1 direction.<br />

We c<strong>la</strong>im first that for any P ∈ ∂B + := ∂BR ∩{x ∈ R N : x1 > 0}, there exists δ ∈ (0,R)such<br />

that<br />

∂u<br />

(x) > 0 ∀x ∈ BR ∩ Bδ(P ). (2.6)<br />

∂x1<br />

In<strong>de</strong>ed, thanks to (2.5) we have,<br />

∂u<br />

=<br />

∂x1<br />

∂u x1<br />

∂r |x| +<br />

<br />

Du − ∂u<br />

<br />

x<br />

· e1<br />

∂r |x|<br />

= ∂u<br />

<br />

x1<br />

∂r |x| +<br />

<br />

∂u<br />

−1<br />

Du −<br />

∂r<br />

∂u<br />

<br />

x<br />

· e1<br />

∂r |x|<br />

= ∂u<br />

<br />

x1<br />

+ o(1) as |x|→R.<br />

∂r |x|<br />

Since P ∈ ∂B + , the c<strong>la</strong>im follows straightforwardly.<br />

Next we follow the construction in [9]. For any λ 0<br />

∂x1<br />

inUε ∩ BR. (2.8)<br />

By <strong>de</strong>finition of μ there holds u uμ in Σμ; thus, if we <strong>de</strong>note Dε = BR−ɛ/2 ∩ Σμ, wehave<br />

<br />

(u − uμ) = a(x)(u − uμ)<br />

u − uμ 0<br />

in Dε,<br />

inDε,<br />

(2.9)


A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591 585<br />

where a(x) = g(u) − g(uμ)/(u − uμ). Thanks to (2.2) and since Dε is in the interior of BR,<br />

a(x) is a boun<strong>de</strong>d function in Dε, and the strong maximum princip<strong>le</strong> applies to (2.9). Since u<br />

tends to infinity at the boundary and is finite in the interior, for ε small we c<strong>le</strong>arly have u ≡ uμ<br />

in Dε: therefore we conclu<strong>de</strong> that u>uμ in Dε, and, since u = uμ on Tμ ∩ ∂Dε and ∂uμ/∂x1 =<br />

−∂u/∂x1 on Tμ, it follows from Hopf boundary <strong>le</strong>mma that<br />

∂u<br />

> 0 onTμ∩∂Dε. ∂x1<br />

Since u ∈ C 1 (BR), the <strong>la</strong>st assertion, together with (2.8), implies that there exists σ>0 such that<br />

∂u<br />

> 0<br />

∂x1<br />

inBR∩{x: μ − σuμ in Σμ. On the other<br />

hand, we can also exclu<strong>de</strong> that ¯x ∈ Tμ; in<strong>de</strong>ed, we have<br />

u(xn) − u <br />

(xn)λn = 2(xn − λn) ∂u<br />

(ξn)<br />

∂x1<br />

for a point ξn ∈ ((xn)λn ,xn). If (a subsequence of) xn converges to a point in Tμ, then for n <strong>la</strong>rge<br />

we have dist(ξn,Tμ)0 contradicting (2.11). We<br />

are <strong>le</strong>ft with the possibility that ¯x ∈ ∂Σμ \ Tμ: but this is also a contradiction since u blows up at<br />

the boundary and it is locally boun<strong>de</strong>d in the interior, so that u(xn) − u((xn)λn ) would converge<br />

to infinity.<br />

Thus μ = 0 and (2.7) holds in the who<strong>le</strong> {x ∈ BR: x1 > 0}. We <strong>de</strong><strong>du</strong>ce that u is symmetric<br />

in the x1 direction and ∂u/∂x1 > 0. Applying to any other direction we conclu<strong>de</strong> that u is radial<br />

and ∂u/∂r > 0. ✷<br />

Remark 2.1. Let us recall that in some special examp<strong>le</strong>s (for instance when g(s) has an exponential<br />

or a power-like growth) the asymptotic behaviour at the boundary of the <strong>gradient</strong> of the<br />

<strong>la</strong>rge solutions has already been studied (see e.g. [2,4,16]) so that the previous result could be<br />

directly applied to prove symmetry. In general, through a blow-up argument, we are ab<strong>le</strong> to prove<br />

(2.5) if<br />

s ↦→ g(s)<br />

√ G(s)<br />

∞<br />

s<br />

1<br />

√ 2G(ξ) dξ (2.12)


586 A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591<br />

is boun<strong>de</strong>d at infinity; however, this assumption does not inclu<strong>de</strong> the case when g has a slow<br />

growth at infinity (such as g(r) ≡ r(ln r) α with α>2) and is not so general as (2.3).<br />

Theorem 2.2. Assume that (2.2)–(2.4) hold. Then any solution u of (2.1) is radial and ∂u/∂r > 0<br />

in BR \{0}.<br />

The following preliminary result is a consequence of more general results in [3,11,12,18].<br />

However, we provi<strong>de</strong> here a simp<strong>le</strong> self-contained proof for the radial case.<br />

Lemma 2.1. Let h be a convex increasing function satisfying the Kel<strong>le</strong>r–Osserman condition<br />

<br />

+∞<br />

a<br />

s<br />

ds<br />

√ < ∞, H(s)= h(t) dt, (2.13)<br />

H(s)<br />

for some a>0. Then the prob<strong>le</strong>m<br />

<br />

−v + h(v) = 0 in BR,<br />

lim|x|→R v(x) =∞<br />

has a unique solution.<br />

a<br />

(2.14)<br />

Proof. Since h is increasing, there exist a maximal and a minimal solution v and v, which are<br />

both radial, so that it is enough to prove that v = v. To this purpose, observe that if v is radial<br />

we have (v ′ rN−1 ) ′ = rN−1 h(v) so that, since v ′ (0) = 0, and rep<strong>la</strong>cing H by H ˜ = H − H(min v)<br />

which is nonnegative on the range of values of v,wehave<br />

which yields<br />

Define now<br />

(v ′ r N−1 ) 2<br />

2<br />

<br />

=<br />

0<br />

r<br />

s 2(N−1) h(v)v ′ ds r 2(N−1) ˜<br />

H(v)<br />

0 v ′ <br />

< 2H(v). ˜<br />

(2.15)<br />

<br />

w = F(v)=<br />

v<br />

∞<br />

ds<br />

.<br />

2H(s) ˜<br />

A straightforward computation, and condition (2.13), show that w solves the prob<strong>le</strong>m<br />

<br />

w = b(w)(|Dw| 2 − 1) in BR,<br />

w = 0 on∂BR,<br />

(2.16)


A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591 587<br />

<br />

where b(w) = h(v)/ 2H(v). ˜ One can easily check that the convexity of h implies that<br />

<br />

h(v)/ 2H(v)is ˜ non<strong>de</strong>creasing, hence b(w) is nonincreasing with respect to w. Moreover, since<br />

|Dw|=|w ′ |=v ′ <br />

/ 2H(v), ˜ from (2.15) one gets |w ′ | < 1. Note that the transformation v ↦→ w<br />

establishes a one-to-one monotone correspon<strong>de</strong>nce between the <strong>la</strong>rge solutions of (2.14) and<br />

the solutions of (2.16), so that w = F(v) and w = F(v) are respectively the minimal and the<br />

maximal solutions of (2.16). Thus we have<br />

′ N−1<br />

(w − w ) r ′ N−1<br />

= r b(w) |w ′ | 2 − 1 − b(w ) |w ′ | 2 − 1 r N−1 b(w) |w ′ | 2 −|w ′ | 2 ,<br />

so that the function z = (w − w ) ′ r N−1 satisfies<br />

z ′ a(r)z, z(0) = 0, where a(r) = b(w)(w ′ + w ′ ).<br />

Because a is locally boun<strong>de</strong>d on [0,R), we <strong>de</strong><strong>du</strong>ce that z 0, hence w − w is non<strong>de</strong>creasing.<br />

Since w − w is nonnegative and w(R) = w(R) = 0 we <strong>de</strong><strong>du</strong>ce that w = w, hence v = v. ✷<br />

Lemma 2.2. Assume that g satisfies (2.3) and (2.4), and that u is a solution of (2.1). Then<br />

and the two limits hold uniformly with respect to {x: |x|=r}.<br />

(i) lim<br />

|x|→R ∇τ u(x) = 0,<br />

(ii)<br />

∂u<br />

lim (x) =∞,<br />

|x|→R ∂r<br />

(2.17)<br />

Proof. In spherical coordinates (r, σ ) ∈ (0, ∞) × S N−1 the Lap<strong>la</strong>ce operator takes the form<br />

u = ∂2u N − 1 ∂u<br />

+<br />

∂r2 r ∂r<br />

1<br />

+ su,<br />

r2 where s is the Lap<strong>la</strong>ce–Beltrami operator on SN−1 .If{γj } N−1<br />

j=1<br />

on SN−1 crossing orthogonally at ˜σ = γj (0), there holds<br />

su(r, ˜σ)= <br />

j1<br />

d 2 u(r, γj (t))<br />

dt 2<br />

<br />

<br />

<br />

t=0<br />

is a system of N − 1 geo<strong>de</strong>sics<br />

. (2.18)<br />

On the sphere the geo<strong>de</strong>sics are <strong>la</strong>rge circ<strong>le</strong>s. The system of geo<strong>de</strong>sics can be obtained by consi<strong>de</strong>ring<br />

a set of skew symmetric matrices {Aj } N−1<br />

j=1 such that 〈Aj ˜σ,Ak ˜σ 〉=δk j , and by putting<br />

γj (t) = etAj ˜σ .<br />

Step 1. Two-si<strong>de</strong> estimate on the tangential first <strong>de</strong>rivatives.<br />

By assumption (2.3) g can be written as<br />

g(s) = g∞(s) +˜g(s),


588 A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591<br />

where g∞(s) is a convex increasing function satisfying (2.4) and ˜g(s) is a locally Lipschitz<br />

function such that ˜g ≡ 0in[M,∞) for some M>0. In particu<strong>la</strong>r, u satisfies<br />

u − g∞(u) =˜g(u).<br />

Since u blows up uniformly, there holds u(x) M if |x|∈[r0,R)for a certain r0 2,<br />

ln r−ln R<br />

ln r0−ln R if N = 2,<br />

and v h (x) = u h (x) +|h|LP (|x|); then v h = u h , and since g∞ is increasing,<br />

(2.22)<br />

v h − u h<br />

g∞ v − g∞(u) in BR \ Br0 . (2.23)


A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591 589<br />

Observe that u h ,asu, also satisfies (2.21), so that in particu<strong>la</strong>r<br />

u h (x) − u(x) → 0 as|x|→R. (2.24)<br />

Therefore vh (x)−u(x) → 0as|x|→R too, whi<strong>le</strong> by construction vh u on ∂Br0 . We conclu<strong>de</strong><br />

from (2.23) (e.g. using the test function (vh − u + ε)−, which is compactly supported, and then<br />

<strong>le</strong>tting ε go to zero) that<br />

v h = u h +|h|LP (r) u.<br />

We recall that the Lie <strong>de</strong>rivative LAj u of u(r, ·) following the vector field tangent to SN−1 η ↦→<br />

Aj η is <strong>de</strong>fined by<br />

so we get, by <strong>le</strong>tting h → 0,<br />

LAj u(r, σ ) = <strong>du</strong>(r,etAj σ)<br />

dt<br />

<br />

<br />

<br />

t=0<br />

<br />

LAj u(r, ˜σ) LP (r) < C(R − r). (2.25)<br />

Step 2. One-si<strong>de</strong> estimate on the tangential second <strong>de</strong>rivatives.<br />

Next we <strong>de</strong>fine the function w h by<br />

w h = uh + u−h − 2u<br />

h2 .<br />

As before, <strong>le</strong>t r0 0 such that<br />

w h ˜L on ∂Br0 .<br />

Moreover, from (2.25) we get that wh = 0on∂BR. We conclu<strong>de</strong> that<br />

h<br />

w <br />

+ (x) ˜LP |x| ,<br />

,


590 A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591<br />

where P(r)is <strong>de</strong>fined in (2.22). Letting h tend to zero we obtain<br />

d 2 u(r, e tAj ˜σ)<br />

dt 2<br />

<br />

<br />

<br />

t=0<br />

Using (2.18), and the fact that ˜σ is arbitrary, we <strong>de</strong>rive<br />

˜LP (r) for r ∈[r0,R). (2.26)<br />

su(r, σ ) (N − 1) ˜LP(r) ∀(r, σ ) ∈[r0,R)× S N−1 . (2.27)<br />

Step 3. Estimate on the radial <strong>de</strong>rivative.<br />

Using (2.22) and (2.27) we <strong>de</strong><strong>du</strong>ce that<br />

Therefore<br />

<br />

∂<br />

r<br />

∂r<br />

(su) + (x) = o(1) uniformly as |x|→R.<br />

<br />

N−1 ∂u<br />

= r<br />

∂r<br />

N−1<br />

<br />

g(u) − 1<br />

su<br />

r2 <br />

r N−1 g∞(u) − o(1)<br />

uniformly as r → R. (2.28)<br />

Now one can easily conclu<strong>de</strong>: <strong>le</strong>t z(r) <strong>de</strong>note the minimal (hence radial) solution of<br />

⎧<br />

⎨ z = g∞(z) in BR \ Br0<br />

⎩<br />

,<br />

z = min∂Br u<br />

0<br />

on ∂Br0 ,<br />

limr→R z =∞.<br />

We have u(x) z(x) if |x| ∈[r0,R), hence g∞(u) g∞(z). Because this <strong>la</strong>st function is not<br />

integrab<strong>le</strong> near ∂BR, one obtains<br />

lim<br />

r<br />

r→R<br />

r0<br />

C<strong>le</strong>arly (2.28) implies<br />

s N−1 <br />

g∞ u(s, σ ) ds →∞ uniformly for σ ∈ SN−1 .<br />

∂u r→R<br />

(r, σ ) −−−→∞<br />

∂r<br />

uniformly for σ ∈ SN−1 .<br />

This comp<strong>le</strong>tes the proof of (2.17). ✷<br />

Proof of Theorem 2.2. By assumptions (2.3) and (2.4), and Lemma 2.2, we <strong>de</strong><strong>du</strong>ce that u satisfies<br />

(2.5), hence we apply Lemma 2.1 to conclu<strong>de</strong>. ✷<br />

Finally, <strong>le</strong>t us point out that thanks to Lemma 2.2 and using the moving p<strong>la</strong>ne method as in<br />

Theorem 2.1, we can <strong>de</strong>rive a result <strong>de</strong>scribing the boundary behaviour of any solution of<br />

<br />

−u + g(u) = 0 inΓR,r ={x ∈ RN : r


A. Porretta, L. Véron / Journal of Functional Analysis 236 (2006) 581–591 591<br />

which extends a simi<strong>la</strong>r result in [9].<br />

Corol<strong>la</strong>ry 1. Assume that g satisfies (2.2)–(2.4). Then any solution of (2.29) satisfies (2.17) and<br />

verifies ∂ru>0 on ΓR,(R+r)/2.<br />

References<br />

[1] A. Aftalion, M. <strong>de</strong>l Pino, R. Letelier, Multip<strong>le</strong> boundary blow-up solutions for nonlinear elliptic equations, Proc.<br />

Roy. Soc. Edinburgh Sect. A 133 (2) (2003) 225–235.<br />

[2] C. Band<strong>le</strong>, M. Essen, On the solutions of quasilinear elliptic prob<strong>le</strong>ms with boundary blow-up, in: Sympos. Math.,<br />

vol. 35, Cambridge Univ. Press, 1994, pp. 93–111.<br />

[3] C. Band<strong>le</strong>, M. Marcus, Large solutions of <strong>semi</strong>linear elliptic equations: Existence, uniqueness and asymptotic behaviour,<br />

J. Anal. Math. 58 (1992) 9–24.<br />

[4] C. Band<strong>le</strong>, M. Marcus, Asymptotic behaviour of solutions and their <strong>de</strong>rivatives, for <strong>semi</strong>linear elliptic prob<strong>le</strong>ms<br />

with blowup on the boundary, Ann. Inst. H. Poincaré Anal. Non Linéaire 12 (2) (1995) 155–171.<br />

[5] H. Brezis, personal communication, January 2005.<br />

[6] Y. Du, Z. Guo, Boundary blow-up solutions and their applications in quasilinear elliptic equations, J. Anal. Math. 89<br />

(2003) 277–302.<br />

[7] Y. Du, Z. Guo, Uniqueness and <strong>la</strong>yer analysis for boundary blow-up solutions, J. Math. Pures Appl. 83 (6) (2004)<br />

739–763.<br />

[8] Y. Du, S. Yan, Boundary blow-up solutions with a spike <strong>la</strong>yer, J. Differential Equations 205 (1) (2004) 156–184.<br />

[9] B. Gidas, W.M. Ni, L. Nirenberg, Symmetry and re<strong>la</strong>ted properties via the maximum princip<strong>le</strong>, Comm. Math.<br />

Phys. 68 (1979) 209–243.<br />

[10] J.B. Kel<strong>le</strong>r, On solutions of u = f(u), Comm. Pure Appl. Math. 10 (1957) 503–510.<br />

[11] M. Marcus, L. Véron, Uniqueness and asymptotic behaviour of solutions with boundary blow-up for a c<strong>la</strong>ss of<br />

nonlinear elliptic equations, Ann. Inst. H. Poincaré 14 (1997) 237–274.<br />

[12] M. Marcus, L. Véron, Existence and uniqueness results for <strong>la</strong>rge solutions of general nonlinear elliptic equations,<br />

J. Evol. Equ. 3 (2003) 637–652.<br />

[13] P.J. McKenna, W. Reichel, W. Walter, Symmetry and multiplicity for nonlinear elliptic differential equations with<br />

boundary blow-up, Nonlinear Anal. 28 (7) (1997) 1213–1225.<br />

[14] R. Osserman, On the inequality u f(u), Pacific J. Math. 7 (1957) 1641–1647.<br />

[15] S.I. Pohožaev, The boundary value prob<strong>le</strong>m for equation U = U 2 , Dokl. Akad. Nauk SSSR 138 (1961) 305–308<br />

(in Russian).<br />

[16] A. Porretta, L. Véron, Asymptotic behaviour for the <strong>gradient</strong> of <strong>la</strong>rge solutions to some nonlinear elliptic equations,<br />

Adv. Nonlinear Stud., in press.<br />

[17] J. Serrin, A symmetry prob<strong>le</strong>m in potential theory, Arch. Rat. Mech. Anal. 43 (1971) 304–318.<br />

[18] L. Véron, Semilinear elliptic equations with uniform blow-up on the boundary, J. Anal. Math. 59 (1992) 231–250.


Journal of Functional Analysis 236 (2006) 592–608<br />

www.elsevier.com/locate/jfa<br />

Non-structural control<strong>la</strong>bility of linear e<strong>la</strong>stic systems<br />

with structural damping<br />

Luc Mil<strong>le</strong>r a,b,∗<br />

a Équipe Modal’X, EA 3454, Université Paris X, Bât. G, 200 Av. <strong>de</strong> <strong>la</strong> République, 92001 Nanterre, France<br />

b Centre <strong>de</strong> Mathématiques Laurent Schwartz, UMR CNRS 7640, Éco<strong>le</strong> Polytechnique, 91128 Pa<strong>la</strong>iseau, France<br />

Received 16 December 2005; accepted 1 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 4 April 2006<br />

Communicated by Paul Malliavin<br />

Abstract<br />

This paper proves that any initial condition in the energy space for the p<strong>la</strong>te equation with square root<br />

damping ¨ζ − ρ˙ζ + 2ζ = u on a smooth boun<strong>de</strong>d domain, with hinged boundary conditions ζ = ζ = 0,<br />

can be steered to zero by a square integrab<strong>le</strong> input function u supported in arbitrarily small time interval<br />

[0,T] and subdomain. As T tends to zero, for initial states with unit energy norm, the norm of this u<br />

grows at most like exp(Cp/T p ) for any real p>1andsomeCp > 0. In<strong>de</strong>ed, this fast control<strong>la</strong>bility<br />

cost estimate is proved for more general linear e<strong>la</strong>stic systems with structural damping and non-structural<br />

controls satisfying a spectral observability condition. Moreover, un<strong>de</strong>r some geometric optics condition on<br />

the subdomain allowing to apply the control transmutation method, this estimate is improved into p = 1<br />

and the <strong>de</strong>pen<strong>de</strong>nce of Cp on the subdomain is ma<strong>de</strong> explicit. These results are analogous to the optimal<br />

ones known for the heat flow.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Control<strong>la</strong>bility; Observability; Control cost; P<strong>la</strong>tes; Structural damping; Transmutation<br />

A wi<strong>de</strong> variety of dissipative linear e<strong>la</strong>stic control systems may be represented by a secondor<strong>de</strong>r<br />

differential equation in a Hilbert space:<br />

* Fax:+330169333019.<br />

E-mail address: mil<strong>le</strong>r@math.polytechnique.fr.<br />

¨ζ(t)+ D ˙ζ(t)+ Sζ(t) = Bu(t), t ∈ R+, (1)<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.001


L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 593<br />

where each dot <strong>de</strong>notes a <strong>de</strong>rivative with respect to the time variab<strong>le</strong> t and the function ζ represents<br />

the evolution of the system un<strong>de</strong>r the action of the input function u.Thestructural vibration<br />

mo<strong>de</strong>s of the conservative system represented by (1) with B = D = 0 are prescribed by the positive<br />

self-adjoint operator S. This i<strong>de</strong>al system is perturbed by a dissipative mechanism prescribed<br />

by the positive self-adjoint operator D. The system is actuated through a control mechanism prescribed<br />

by the operator B (possibly unboun<strong>de</strong>d to take into account trace operators prescribing<br />

the boundary value of distributed states). Throughout this paper, control<strong>la</strong>bility will always mean<br />

the ability of steering any initial state (z(0), ˙z(0)) to zero over a finite time by some appropriate<br />

input function u (i.e. exact control<strong>la</strong>bility to zero or null control<strong>la</strong>bility).<br />

This paper concerns the specific dissipative mechanism D = S α with α ∈ (0, 1) cal<strong>le</strong>d structural<br />

damping, which generalizes the square root damping mo<strong>de</strong>l α = 1/2 intro<strong>du</strong>ced in [6]:<br />

“The basic property of structural damping, which is said to be consistent with empirical studies,<br />

is that the amplitu<strong>de</strong>s of the normal mo<strong>de</strong>s of vibration are attenuated at rates which are proportional<br />

to the oscil<strong>la</strong>tion frequencies.” This mo<strong>de</strong>l was also studied un<strong>de</strong>r the name “proportional<br />

damping” (cf. [3]). The quite different case α = 1 is known as “Kelvin–Voigt” damping. When<br />

B is the i<strong>de</strong>ntity and α ∈ (0, 1], this is the first c<strong>la</strong>ss of parabolic-like control mo<strong>de</strong>ls consi<strong>de</strong>red<br />

in [13,24] with the extra assumption that S has compact resolvent, dispensed with in [2].<br />

This paper focuses on the cost of fast controls as in [2,24]. The control<strong>la</strong>bility results known<br />

for these systems hold for a control time which can be chosen as small as wished. This asymptotic<br />

is referred to as fast control. Thecost over a given time is the supremum over every initial state<br />

with unit energy norm of the smal<strong>le</strong>st norm of an input function which steers it to zero over the<br />

given time. The study of the cost of fast controls was initiated by Seidman (cf. references in [17])<br />

and recently revived by Da Prato who connected it to some properties of stochastic differential<br />

equations (cf. references in [2]).<br />

The earlier results restricted to e<strong>le</strong>mentary forms of control operators (mainly B is the i<strong>de</strong>ntity<br />

or has rank one, cf. [2,11,13,14,21,23,24]). A key point in their proofs is (loosely speaking) the<br />

existence of a common eigenbasis for the three operators S, D and B, mo<strong>de</strong>ling respectively the<br />

structure, the damping and the control. On the contrary, the control<strong>la</strong>bility results of this paper<br />

apply to non-structural controls, e.g. locally distributed control.<br />

The main application is to the p<strong>la</strong>te equation with square root damping on a smooth boun<strong>de</strong>d<br />

domain M of R d with hinged boundary conditions:<br />

¨ζ − ρ˙ζ + 2 ζ = u on R+ × M, ζ = ζ = 0 onR+ × ∂M, (2)<br />

where ρ>0 and the input function u is supported on a non-empty subdomain Ω (cf. Theorem 2).<br />

Fast control<strong>la</strong>bility is proved to hold for any control region Ω. As the control time T tends to<br />

zero, the cost is proved to grow at most like exp(Cβ/T β ) for any β>1 and some Cβ > 0. If<br />

the <strong>le</strong>ngth LΩ of the longest generalized ray of geometrical optics in M which does not intersect<br />

Ω is finite (this is the condition of [5]) and ρLΩ and some positive C which does not <strong>de</strong>pend on Ω. These results<br />

are analogous to the optimal fast control<strong>la</strong>bility cost known for the heat flow (cf. [10,17]). They<br />

confirm the formal analogy: ∂ 2 t −2∂t + 2 = (∂t −) 2 . On the contrary, when Ω = M (i.e. B is<br />

the i<strong>de</strong>ntity), [2,24] prove that the fast control<strong>la</strong>bility cost grows like 1/T β for some β 1/2, as<br />

in finite-dimensional systems (cf. [22]).<br />

Earlier methods to estimate the cost of fast controls were global parabolic Car<strong>le</strong>man estimates<br />

(cf. [2,10]), the Fourier transform method for constructing functions bi-orthogonal to exponential<br />

series (cf. [17,23] and references therein) and the transmutation control method (cf. [17,19]). The


594 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

<strong>la</strong>st two are combined in Section 2 to take the geometry of the control region into account and<br />

improve the cost estimate for (2) as stated above and more precisely in Theorem 2.<br />

The proof of the abstract result, Theorem 1 of Section 1, applies a new method using the control<br />

strategy of Lebeau and Robbiano in [15] as imp<strong>le</strong>mented in [16] (the companion paper [20]<br />

applies this method to a simp<strong>le</strong>r mo<strong>de</strong>l: the holomorphic <strong>semi</strong>group generated by exp(−tS β ),<br />

β>0). The key assumption is an observability condition on the spectral subspaces of S with<br />

respect to B stated in Definition 1. It is an abstract version of a result on sums of eigenfunctions<br />

of the Dirich<strong>le</strong>t Lap<strong>la</strong>cian proved in [12,16] by local elliptic Car<strong>le</strong>man estimates and exten<strong>de</strong>d to<br />

non-compact manifolds in [18]. Therefore the abstract result applies to (2) (such concrete mo<strong>de</strong>ls<br />

with other forms of controls are consi<strong>de</strong>red e.g. in [14, Chapter 3]) even if, e.g., M is unboun<strong>de</strong>d,<br />

the Dirich<strong>le</strong>t Lap<strong>la</strong>cian is positive with non-compact resolvent and Ω is the exterior of a compact<br />

subdomain.<br />

It is c<strong>le</strong>arly <strong>de</strong>sirab<strong>le</strong> to study p<strong>la</strong>te equations with other boundary conditions or with locally<br />

distributed controls on the boundary instead of the interior. Other open prob<strong>le</strong>ms are mentioned<br />

in the remarks of Section 2.1.<br />

1. A structurally damped linear e<strong>la</strong>stic control system<br />

Before stating the abstract mo<strong>de</strong>l and the theorem precisely, we need to intro<strong>du</strong>ce a few notations.<br />

Let H0 and U be Hilbert spaces with respective norms ·0 and ·.LetA be a self-adjoint,<br />

positive and boun<strong>de</strong>dly invertib<strong>le</strong> unboun<strong>de</strong>d operator on H0 with domain D(A). We intro<strong>du</strong>ce<br />

the Sobo<strong>le</strong>v sca<strong>le</strong> of spaces based on A. For any positive integer p, <strong>le</strong>tHp <strong>de</strong>note the Hilbert<br />

space D(A p/2 ) with the norm xp =A p/2 x0 (which is equiva<strong>le</strong>nt to the graph norm x0 +<br />

A p/2 x0). We i<strong>de</strong>ntify H0 and U with their <strong>du</strong>als. Let H−p <strong>de</strong>note the <strong>du</strong>al of Hp. Since Hp is<br />

<strong>de</strong>nsely continuously embed<strong>de</strong>d in H0, the pivot space H0 is <strong>de</strong>nsely continuously embed<strong>de</strong>d in<br />

H−p, and H−p is the comp<strong>le</strong>tion of H0 with respect to the norm x−p =A −p/2 x0.<br />

Let the observation operator C be in L(H2,U), which <strong>de</strong>notes boun<strong>de</strong>d operators from H2<br />

to U, and <strong>le</strong>t B ∈ L(U, H−2) <strong>de</strong>note the <strong>du</strong>al of C.<br />

Let α ∈ (0, 1) <strong>de</strong>note the structural dissipation power, and <strong>le</strong>t ρ>0 <strong>de</strong>note the dissipativity<br />

constant. With the structural operator A and the control operator B, they <strong>de</strong>fine the second-or<strong>de</strong>r<br />

Cauchy prob<strong>le</strong>m with input function u:<br />

¨ζ(t)+ ρA 2α ˙ζ(t)+ A 2 ζ(t)= Bu(t),<br />

ζ(0) = ζ0 ∈ H2, ˙ζ(0) = ζ1 ∈ H0, u∈L 2 loc (R; U). (3)<br />

In or<strong>de</strong>r to <strong>de</strong>fine the (mild) solution of this prob<strong>le</strong>m, we assume that B ∈ L(U, H0) (which<br />

is enough for the application in Section 2) or, more generally, that B is admissib<strong>le</strong> in a sense<br />

specified <strong>la</strong>ter in (9).<br />

To state the “observability condition” on the spectral subspaces of A with respect to C of<br />

the main theorem, we first intro<strong>du</strong>ce our spectral notations. Given γ>0 and μ>1, applying<br />

the functional calculus for self-adjoint operators to the positive operator A γ and the boun<strong>de</strong>d<br />

function on R + <strong>de</strong>fined by 1λμ = 1ifλ μ and 1λμ = 0, otherwise, yields the spectral<br />

projector 1A γ μ. The image of H0 un<strong>de</strong>r this projection operator is just the spectral subspace<br />

1A γ μH0 of A γ .


L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 595<br />

Definition 1. Let γ>0. The observability of low mo<strong>de</strong>s of A γ through C at exponential cost<br />

holds if there are positive constants D0 and D1 such that<br />

∀μ>1, ∀v ∈ 1A γ μH0, v0 D0e D1μ Cv. (4)<br />

This abstract condition is satisfies in some concrete applications given in the next section. As<br />

illustrated in the proof of the following main theorem (cf. Section 1.3), it allows to compare the<br />

free dissipation of high mo<strong>de</strong>s to the cost of controlling low mo<strong>de</strong>s.<br />

Theorem 1. Assume that observability of low mo<strong>de</strong>s of A γ through C at exponential cost<br />

holds for some γ ∈ (0, 1) (cf. Definition 1). For all ρ>0 and α ∈ (γ /2, 1 − γ/2), for all<br />

β>(2min{α, 1 − α}/γ − 1) −1 , there are positive constants C1 and C2 such that for all<br />

T ∈ (0, 1], for all ζ0 and ζ1, there is an input function u such that the solution ζ of (3) satisfies<br />

ζ(T)= ˙ζ(T)= 0 with the cost estimate:<br />

T<br />

0<br />

<br />

u(t) 2 dt C2 exp<br />

<br />

C1<br />

T β<br />

1.1. The <strong>du</strong>ality between observation and control<br />

<br />

ζ0 2 2 +ζ1 2 0 .<br />

The proof of Theorem 1 uses the well-known equiva<strong>le</strong>nce between control<strong>la</strong>bility and observability<br />

(cf. [9]). In this section, we c<strong>la</strong>rify in what sense the <strong>du</strong>al of the control prob<strong>le</strong>m (3) is the<br />

observation of the following Cauchy prob<strong>le</strong>m (without input):<br />

¨z(t) + ρA 2α ˙z(t) + A 2 z(t) = 0, z(0) = z0 ∈ H2, ˙z(0) = z1 ∈ H0. (5)<br />

The second-or<strong>de</strong>r differential equations (3) and (5) may be restated as first-or<strong>de</strong>r systems by<br />

setting ξ(t) = (ζ(t), ˙ζ(t)) and x(t) = (z(t), ˙z(t)):<br />

˙ξ(t)− Aξ(t) = Bu(t), ξ(0) = ξ0 ∈ X, u ∈ L 2 loc (R; U), (6)<br />

˙x(t) = Ax(t), x(0) = x0 ∈ X. (7)<br />

The state space is X = H2 × H0. The <strong>semi</strong>group generator A of (7) is <strong>de</strong>fined by<br />

<br />

0 I<br />

A =<br />

−A2 −ρA2α <br />

, D(A) = (z0,z1) ∈ H2 × H2 | A 2 z0 + ρA 2α <br />

z1 ∈ H0 .<br />

It inherits from −A the necessary and sufficient properties of Lumer–Phillips for generating a<br />

contraction <strong>semi</strong>group.<br />

The control operator is B = ΠB, where Π : X → H0 is <strong>de</strong>fined by Π(z0,z1) = z1. IfB ∈<br />

L(U, H0) (as in the application in Section 2) then B ∈ L(U, X). In<strong>de</strong>ed Theorem 1 is valid in<br />

the following more general (canonical) setting intro<strong>du</strong>ced by Weiss in [25]. Let X1 be D(A)<br />

with the norm x1 =Ax and <strong>le</strong>t X−1 be the comp<strong>le</strong>tion of X with respect to the norm<br />

x−1 =A −1 x. At first, we only assume B ∈ L(U, X−1). In or<strong>de</strong>r to <strong>de</strong>fine the unique (mild)<br />

solution ξ ∈ C(R+; X) of (6) by the integral formu<strong>la</strong>


596 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

ξ(t) = e tA t<br />

ξ(0) + e (t−s)A Bu(s) ds, (8)<br />

0<br />

we also make the admissibility assumption: forsomeT>0 (hence for all T>0) there is a<br />

positive constant KT such that<br />

∀u ∈ L 2 loc (R; U),<br />

<br />

T<br />

<br />

e<br />

<br />

tA <br />

<br />

<br />

Bu(t) dt <br />

<br />

0<br />

2<br />

KT<br />

We <strong>de</strong>fine the <strong>du</strong>ality pairing on X by<br />

<br />

(ζ0,ζ1), (z0,z1) =〈Aζ0,Az0〉0 −〈ζ1,z1〉0.<br />

T<br />

0<br />

<br />

u(t) 2 dt. (9)<br />

With respect to this pairing, X and A are their own <strong>du</strong>al, X−1 is the <strong>du</strong>al of X1 and B is the <strong>du</strong>al<br />

of the observation operator C = CΠ. The assumptions on B are equiva<strong>le</strong>nt to C ∈ L(X1,U)and,<br />

for all x0 ∈ D(A),<br />

T<br />

0<br />

<br />

tA<br />

Ce x0<br />

2 dt KT x0 2 .<br />

Therefore the output map x0 ↦→ Ce tA x0 from D(A) to L 2 ([0,T]; U) has a continuous extension<br />

to X.N.b.ifα 1/2, then X1 = H4 × H2 and X−1 = H0 × H−2.<br />

We recall the <strong>du</strong>ality between control<strong>la</strong>bility and observability (cf. [9]).<br />

Lemma 1. Let T>0 and CT > 0. The following properties are equiva<strong>le</strong>nt:<br />

(i) For all initial state ξ0 ∈ X, there is an input function u ∈ L2 loc (R; U) such that the solution<br />

ξ ∈ C(R+; X) of (6) satisfies ξ(T) = 0 and uL2 (0,T ;U) CT ξ0.<br />

(ii) For all initial state x0 ∈ X, the solution x(t) = etAx0 of (7) satisfies the observation inequality<br />

x(T ) CT Cx(t)L2 (0,T ;U) .<br />

N.b. the smal<strong>le</strong>st constant CT such that these properties hold is the control<strong>la</strong>bility cost mentioned<br />

in the intro<strong>du</strong>ction. The estimate in Theorem 1 writes C 2 T C2 exp(C1/T β ). The contractivity<br />

of e tA , (8) and (9) imply the estimates:<br />

T<br />

0<br />

<br />

ξ(T) 2 2 1 + KT C 2 <br />

ξ(0) T<br />

2 , (10)<br />

<br />

ξ(t) 2 dt 2T 1 + KT C 2 <br />

ξ(0) T<br />

2 , (11)<br />

since Kt and t<br />

0 u(s)2 ds are non<strong>de</strong>creasing, although Ct is nonincreasing.


1.2. Spectral and growth bounds<br />

L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 597<br />

The proof of Theorem 1 relies on a spectral <strong>de</strong>composition of the prob<strong>le</strong>m. We extend the<br />

action of the spectral projector 1A γ μ to X according to 1A γ μ(z0,z1) = (1A γ μz0, 1A γ μz1).<br />

It commutes with A and the generated <strong>semi</strong>group. Therefore X is the orthogonal sum of the<br />

invariant subspaces 1A γ μX (low mo<strong>de</strong>s) and 1A γ >μX (high mo<strong>de</strong>s).<br />

The restriction of e tA to 1A γ >μX satisfies the following exponential <strong>de</strong>cay bound.<br />

Proposition 1. Let γ ′ = γ/(2min{α, 1 − α}). There is an r>0 such that<br />

∀μ 1, ∀x ∈ 1A γ >μX, ∀t 0,<br />

<br />

e tA x exp 1/γ ′<br />

−rμ t x.<br />

This re<strong>du</strong>ces to a spectral bound thanks to the results of Chen and Triggiani on the differentiability<br />

of e tA (cf. [7,8]). We first prove two spectral <strong>le</strong>mmas.<br />

Lemma 2. The spectrum of A re<strong>la</strong>tes to the spectrum of A according to<br />

σ(−A) ⊂ λ ∈ C |∃μ ∈ σ(A), Pμ(λ) = 0 with Pμ(λ) = λ 2 − ρμ 2α λ + μ 2 .<br />

Proof. Let λ/∈{λ ∈ C |∃μ ∈ σ(A), Pμ(λ) = 0}. The function μ ↦→ Pμ(λ)/μ2 is continuous<br />

on (0, +∞) ⊃ σ(A), it tends to 1 as μ tends to infinity and it does not vanish on the<br />

closed set σ(A). Hence, there is an ε>0 such that, for all μ ∈ σ(A), |Pμ(λ)/μ2 | >ε. Since<br />

μ ↦→|μ2Pμ(λ) −1 | is boun<strong>de</strong>d on σ(A) (by ε−1 ), we have A2PA(λ) −1 ∈ L(H0) ⊂ L(H2,H0),<br />

PA(λ) −1 ∈ L(H0,H4) ⊂ L(H0,H2) and (λI − ρA2α )PA(λ) −1 ∈ L(H2). Therefore the operator<br />

M(λ) <strong>de</strong>fined by<br />

<br />

(λI − ρA2α )PA(λ)<br />

M(λ) =<br />

−1 −PA(λ) −1 <br />

A 2 PA(λ) −1 λPA(λ) −1<br />

is boun<strong>de</strong>d on X.ButM(λ)(λI + A) = (λI + A)M(λ) = I , so that M(λ) is the boun<strong>de</strong>d inverse<br />

of λI + A. Hence λ/∈ σ(−A). ✷<br />

Lemma 3. The roots λ± = (1 ± 1 − (2μ 1−2α /ρ) 2 )ρμ 2α /2 of Pμ(λ) satisfy:<br />

∀μ 1, min{Re λ+, Re λ−} rμ 2min{α,1−α} , with r = min{ρ/2, 1/ρ}.<br />

Proof. Let x = 2μ 1−2α /ρ. Ifx 1, then Re λ+ = Re λ− = μ 2α ρ/2. Otherwise λ± ∈ R, λ+ =<br />

(1 + √ ...)μ 2α ρ/2 μ 2α ρ/2, and λ− = (1 − √ 1 − x 2 )μ 2α ρ/2 x 2 μ 2α ρ/4 = μ 2(1−α) /ρ.<br />

Since min{μ 2α ,μ 2(1−α )}=μ 2min{α,1−α} for μ 1, gathering these lower bounds yields the<br />

<strong>le</strong>mma. ✷<br />

Proof of Proposition 1. Since etA is a differentiab<strong>le</strong> <strong>semi</strong>group for α ∈ (0, 1] ([8] proves that<br />

this <strong>semi</strong>group is of Gevrey c<strong>la</strong>ss and that it is analytic if and only if α ∈[1/2, 1]), it is eventually<br />

continuous for the operator norm topology. The <strong>semi</strong>group generated by the restriction Aμ of A<br />

to 1Aγ >μX inherits this property. But Lemma 3 and the proof of Lemma 2 imply σ(−Aμ) ⊂<br />

{λ ∈ C | Re λ rμ1/γ ′<br />

} with r = min{ρ/2, 1/ρ}. Therefore the growth bound in Proposition 1<br />

holds. ✷


598 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

1.3. Proof of Theorem 1<br />

The <strong>la</strong>st ingredient of this proof is the following cost estimate corresponding to the control<br />

operator B = I proved in [2].<br />

Proposition 2. (Avalos–Lasiecka, 2003) For all ρ>0, for all α ∈ (0, 1), there are positive constants<br />

c1 and c2 such that for all T ∈ (0, 1] the solutions of (5) satisfy:<br />

∀z0 ∈ H2, ∀z1 ∈ H0,<br />

<br />

z(T ) 2 2 + ˙z(T ) 2 c2<br />

0 T c1<br />

T<br />

0<br />

<br />

˙z(t) 2 dt. (12)<br />

(In<strong>de</strong>ed, [2] specifies how the power c1 <strong>de</strong>pends on α.)<br />

In a first step, from the stationary condition in Definition 1, Proposition 2 and the <strong>du</strong>ality in<br />

Lemma 1, we <strong>de</strong><strong>du</strong>ce the “control<strong>la</strong>bility of low mo<strong>de</strong>s at exponential cost” in the corresponding<br />

dynamics. In a second step, combining it with the <strong>de</strong>cay bound in Proposition 1 according to the<br />

iterative control strategy intro<strong>du</strong>ced by Lebeau and Robbiano in [15], we prove the control<strong>la</strong>bility<br />

of all mo<strong>de</strong>s. We estimate the control<strong>la</strong>bility cost as the control time tends to zero, like in [20],<br />

in the <strong>la</strong>st step.<br />

First step. With the notations intro<strong>du</strong>ced in Section 1.1, the observation inequality (12) in Proposition<br />

2 writes:<br />

∀x0 ∈ X,<br />

<br />

e T A x0<br />

<br />

2 c2<br />

T<br />

T c1<br />

0<br />

<br />

tA<br />

Πe x0<br />

2 dt. (13)<br />

Let τ ∈ (0, 1], μ 1 and x0 ∈ 1A γ μX. For all t ∈[0,τ], we may apply (4) to Πe tA x0 since it<br />

is in 1A γ μH0:<br />

<br />

Πe tA <br />

x02<br />

0 D2 0e2D1μ CΠe tA <br />

x02<br />

.<br />

First integrating on [0,τ], then using (13) yields:<br />

<br />

e T A x0<br />

<br />

2 D 2 0<br />

c2<br />

e2D1μ<br />

τ c1<br />

τ<br />

0<br />

<br />

Ce tA <br />

x02<br />

dt.<br />

This “low mo<strong>de</strong>s fast observability for e tA at exponential cost” is equiva<strong>le</strong>nt, by the same <strong>du</strong>ality<br />

as in Lemma 1, to the control<strong>la</strong>bility property: for all τ ∈ (0, 1] and μ>1, there is a boun<strong>de</strong>d<br />

operator S τ μ : X → L2 (0,τ; U) such that, for all ξ0 ∈ 1A γ μX, the solution ξ ∈ C(R+,X) of<br />

(6) with input function u = S τ μ ξ0 satisfies 1A γ μξ(τ) = 0, and ∃d3 > 0, S τ μ (d3/τ c1/2 )e D1μ<br />

(cost estimate).<br />

Second step. The hypothesis on α implies that the γ ′ of Proposition 1 is lower than 1. We<br />

intro<strong>du</strong>ce a dyadic sca<strong>le</strong> of mo<strong>de</strong>s μk = 2 k (k ∈ N) and a sequence of time intervals τk = σδT/μ δ k


L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 599<br />

where δ ∈ (0,γ ′−1 − 1) and σδ = (2 <br />

k∈N 2−kδ ) −1 > 0, so that the sequence of times <strong>de</strong>fined<br />

recursively by T0 = 0 and Tk+1 = Tk + 2τk converges to T . The strategy consists in steering the<br />

initial state ξ0 to 0, through the sequence of states ξk = ξ(Tk) ∈ 1Aγ >μk−1X composed of ever<br />

higher mo<strong>de</strong>s, by applying recursively the input function uk = S τk<br />

μkξk to ξk <strong>du</strong>ring a time τk and<br />

no input <strong>du</strong>ring a time τk. Intro<strong>du</strong>cing the notations<br />

εk =ξk, Ck = D2e D1μk /τ c1/2<br />

k , and ρk =<br />

the cost estimate of the previous step writes S τk<br />

μk Ck and implies:<br />

u 2<br />

L 2 (0,T ;U)<br />

= <br />

k∈N<br />

uk 2<br />

L 2 (0,τk;U)<br />

<br />

k∈N<br />

Ck+1εk+1<br />

Ckεk<br />

Since τk T 1, the estimate (10) between the times Tk and Tk + τk implies<br />

<br />

ξ(Tk + τk) 2 2 1 + K1C 2 2<br />

k εk .<br />

2<br />

, (14)<br />

C 2 k ε2 k . (15)<br />

Since 1Aγ μkξ(Tk ′<br />

−rμ1/γ + τk) = 0 and Proposition 1 imply εk+1 e k<br />

τkξ(Tk + τk), we <strong>de</strong><strong>du</strong>ce<br />

ε 2 k+1<br />

1/γ ′ <br />

2e−2rτkμk 1 + K1C 2 2<br />

k εk .<br />

Since Ck+1/Ck = 2 δc1/2 e D1μk , we <strong>de</strong><strong>du</strong>ce that, for any D3 > 4D1, there is a D4 > 0 such that<br />

ρk 2 1+δc1<br />

Since γ ′−1 − δ>1, this implies:<br />

<br />

e −2D1μk<br />

K1D<br />

+ 2 2<br />

τ c1<br />

<br />

1/γ ′<br />

4D1μk−2rτkμ<br />

e k <br />

k<br />

D4<br />

T c1 eD3μk−2rσδTμ γ ′−1−δ k . (16)<br />

∀ρ ∈ (0, 1), ∃N ∈ N, k N ⇒ ρk ρ.<br />

Therefore limk εk = 0 and the <strong>la</strong>st series in (15) converges. This comp<strong>le</strong>tes the proof of the<br />

control<strong>la</strong>bility in Theorem 1.<br />

Third step. The control<strong>la</strong>bility cost CT , formally <strong>de</strong>fined after Lemma 1, satisfies:<br />

Since<br />

l μl,<br />

<br />

0kl−1<br />

C 2 T C2 0<br />

<br />

1 + <br />

<br />

l1 0kl−1<br />

μk μl and <br />

0kl−1<br />

ρk<br />

<br />

. (17)<br />

μ γ ′−1 −δ<br />

k<br />

μ γ ′−1−δ l−1 /2,


600 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

(16) implies<br />

<br />

0kl−1<br />

ρk exp D3 + ln D4/T c1 μl − rσδTμ γ ′−1−δ l−1 .<br />

Hence, setting q = 2 γ ′−1 −δ and T ′ = rσδT/q we have<br />

∀l 1,<br />

<br />

0kl−1<br />

As in [20], plugging this in (17) yields the cost estimate:<br />

ρk exp DT ′2l − T ′ q l with DT ′ ∼<br />

T ′ →0 c1 ln(1/T ′ ).<br />

∀β >βq, ∃D6 > 0, ∃D7 > 0, C 2 T D6<br />

<br />

D7<br />

exp<br />

T ′β<br />

<br />

−1 ln q<br />

with βq = − 1 .<br />

ln 2<br />

Since T ′ is proportional to T and βq <strong>de</strong>creases to (γ ′−1 − 1) −1 = (2min{α, 1 − α}/γ − 1) −1 as<br />

δ <strong>de</strong>creases to 0, this proves the estimate in Theorem 1 restated after Lemma 1.<br />

2. Interior control<strong>la</strong>bility of structurally damped p<strong>la</strong>tes<br />

This section concerns concrete applications of the abstract mo<strong>de</strong>l studied in the previous section.<br />

The main application is to the p<strong>la</strong>te equation with square root damping and interior control<br />

in Ω with hinged boundary conditions:<br />

¨ζ − ρ˙ζ + 2 ζ = χΩu on R+ × M, ζ = ζ = 0 onR+ × ∂M,<br />

ζ(0) = ζ0 ∈ H 2 (M) ∩ H 1 0 (M), ˙ζ(0) = ζ1 ∈ L 2 (M), u ∈ L 2 loc (R+ × M). (18)<br />

In this section, M is a smooth connected comp<strong>le</strong>te d-dimensional Riemannian manifold with<br />

metric g and non-empty boundary ∂M, M <strong>de</strong>notes the interior and M = M ∪ ∂M.Let <strong>de</strong>note<br />

the Dirich<strong>le</strong>t Lap<strong>la</strong>cian on L 2 (M) with domain D() = H 1 0 (M) ∩ H 2 (M) (thus <strong>de</strong>notes a<br />

negative differential operator with variab<strong>le</strong> coefficients <strong>de</strong>pending on the metric g). N.b. the<br />

results are already interesting when (M, g) is a smooth domain of the Eucli<strong>de</strong>an space R d ,so<br />

that = ∂ 2 /∂x 2 1 +···+∂2 /∂x 2 d .LetχΩ <strong>de</strong>note the multiplication by the characteristic function<br />

of an open subset Ω =∅of M.<br />

For simplicity, in the following theorem proved in Section 2.4, we assume that M is compact.<br />

The second part of this theorem makes a geometric assumption on Ω <strong>du</strong>e to Bardos–Lebeau–<br />

Rauch based on generalized geo<strong>de</strong>sics. In this context, the generalized geo<strong>de</strong>sics are continuous<br />

trajectories t ↦→ x(t) in M which follow geo<strong>de</strong>sic curves at unit speed in M (so that on these<br />

intervals t ↦→ ˙x(t) is continuous); if they hit ∂M transversely at time t0, then they ref<strong>le</strong>ct as light<br />

rays or billiard balls (and t ↦→ ˙x(t) is discontinuous at t0); if they hit ∂M tangentially then either<br />

there exists a geo<strong>de</strong>sic in M which continues t ↦→ (x(t), ˙x(t)) continuously and they branch onto<br />

it, or there is no such geo<strong>de</strong>sic curve in M and then they gli<strong>de</strong> at unit speed along the geo<strong>de</strong>sic<br />

of ∂M which continues t ↦→ (x(t), ˙x(t)) continuously until they may branch onto a geo<strong>de</strong>sic<br />

in M. For this result and whenever generalized geo<strong>de</strong>sics are mentioned, we make the additional<br />

assumption that they can be uniquely continued at the boundary ∂M (as in [5], to en<strong>sur</strong>e this, we


L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 601<br />

may assume either that ∂M has no contacts of infinite or<strong>de</strong>r with its tangents, or that g and ∂M<br />

are real analytic).<br />

Let LΩ <strong>de</strong>note the <strong>le</strong>ngth of the longest generalized geo<strong>de</strong>sic in M which does not intersect<br />

Ω. For instance, we recall that LΩ < ∞ if Ω is a neigbourhood of the boundary of a<br />

smooth domain M of R d (in that case LΩ is the <strong>le</strong>ngth of the longest segment in M \ Ω) and<br />

that LΩ < 2D if Ω is a neighborhood of a hemisphere of the boundary of a Eucli<strong>de</strong>an ball M of<br />

diameter D.<br />

Theorem 2. For all ρ>0 and Ω =∅, for all β>1, there are C1 > 0 and C2 > 0 such that,<br />

for all T ∈ (0, 1], for all ζ0 and ζ1, there is an input function u such that the solution ζ of (18)<br />

satisfies ζ(T)= ˙ζ(T)= 0 and the cost estimate:<br />

T<br />

0<br />

<br />

M<br />

|u| 2 dxdt C2 exp <br />

C1/T<br />

β<br />

M<br />

|ζ0| 2 +|ζ1| 2 dx.<br />

For all ρ ∈ (0, 2) and L>LΩ, this result holds with this estimate improved by rep<strong>la</strong>cing<br />

exp(C1/T β ) with exp(CρL 2 /T) where Cρ does not <strong>de</strong>pend on Ω.<br />

2.1. Application of Theorem 1<br />

In<strong>de</strong>ed, Theorem 1 applies to structurally damped p<strong>la</strong>tes with interior control in Ω more<br />

general than (18):<br />

¨ζ + ρ(−) 2α ˙ζ + 2 ζ = χΩu on R+ × M,<br />

ζ(0) = ζ0 ∈ H 2 (M) ∩ H 1 0 (M), ˙ζ(0) = ζ1 ∈ L 2 (M), u ∈ L 2 [0,T]×M . (19)<br />

This is the abstract system (6) where the generator is A =−, the state and input space is<br />

H0 = U = L 2 (M), and the control and observation operator is B = C = χΩ. N.b. for square root<br />

damping (α = 1/2), X1 = H4 × H2 ={(z0,z1) ∈ H 4 (M) × H 2 (M) | z1 = z0 = z0 = 0on∂M}<br />

and the solution of (6) satisfy the boundary conditions ζ = ζ = 0onR+ × ∂M in a generalized<br />

sense.<br />

If M is not compact, assume that Ω is the exterior of a compact set K such that K ∩ Ω ∩<br />

∂M =∅and that 0 /∈ σ(). In this setting, the observability of low mo<strong>de</strong>s of (−) 1/2 through<br />

C at exponential cost holds (cf. Definition 1). When M is compact this is an inequality on sums<br />

of eigenfunctions proved as Theorem 3 in [16] and Theorem 14.6 in [12]. This was generalized<br />

to non-compact M in [18]. Applying Theorem 1 with γ = 1/2 yields:<br />

Corol<strong>la</strong>ry 1. For all ρ>0, α ∈ (1/4, 3/4), and β>min{4α − 1, 3 − 4α} −1 , there are C1 > 0<br />

and C2 > 0 such that, for all T ∈ (0, 1], for all ζ0 and ζ1, there is an input function u such that<br />

the solution ζ of (19) satisfies ζ(T)= ˙ζ(T)= 0 and the cost estimate:<br />

u 2<br />

L2 C2 exp C1/T β ζ0 2<br />

H 2 +ζ1 2<br />

L2 <br />

.<br />

Remark 1. Note that Proposition 2 proves control<strong>la</strong>bility from Ω = M when α ∈[0, 1). It results<br />

from [2] that control<strong>la</strong>bility does not hold in Corol<strong>la</strong>ry 1 for α = 1. For α = 0, it can be proved


602 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

by the transmutation control method that the control<strong>la</strong>bility for ρ = 0 (which holds if Ω satisfies<br />

the condition LΩ < ∞ of Theorem 2) implies the control<strong>la</strong>bility over the same time for ρ>0.<br />

The case α ∈ (0, 1/4]∪[3/4, 1) with Ω = M is still open.<br />

Remark 2. More generally, Theorem 1 applies to A = (−) 1/(2γ) with γ ∈ (0, 1). It does not<br />

apply to the wave equation which would correspond to γ = 1. The wave equation with square<br />

root damping (α = 1/2) is just out of reach and seems to us an interesting open prob<strong>le</strong>m (the<br />

appendix in [20] proves that it is not control<strong>la</strong>b<strong>le</strong> by a one-dimensional input). It results from [1]<br />

that the wave equation (γ = 1) with Kelvin–Voigt damping (α = 1) is not control<strong>la</strong>b<strong>le</strong> from<br />

any Ω = M. It results from [5] that the damped wave equation (γ = 1, α = 0), is control<strong>la</strong>b<strong>le</strong><br />

from Ω or not <strong>de</strong>pending on whether the control time is greater or lower than LΩ (<strong>de</strong>fined before<br />

Theorem 2).<br />

2.2. Smoothing<br />

The control transmutation method of Section 2.4 applies to initial data smoother than in Theorem<br />

2. This drawback is easily overcome by a general abstract remark, ma<strong>de</strong> here, concerning<br />

the null-control<strong>la</strong>bility of analytic <strong>semi</strong>groups: in the smoothness sca<strong>le</strong> of Sobo<strong>le</strong>v spaces <strong>de</strong>fined<br />

by the generator, if fast control<strong>la</strong>bility holds for initial data in some space, then it holds for<br />

initial data in any <strong>le</strong>ss smooth space; moreover, the same statement holds for fast control<strong>la</strong>bility<br />

at exponential cost.<br />

We recall the setting of Section 1.1. A is the boun<strong>de</strong>dly invertib<strong>le</strong> generator of a boun<strong>de</strong>d<br />

analytic <strong>semi</strong>group on the Hilbert space X. For any p>0, <strong>le</strong>t Xp <strong>de</strong>note the Hilbert space<br />

D((−A) p ) with the norm xp =(−A) p x (which is equiva<strong>le</strong>nt to the graph norm) and <strong>le</strong>t<br />

X−p be the comp<strong>le</strong>tion of X with respect to the norm x−p =(−A) −p x. There is a <strong>du</strong>ality<br />

pairing on X such that X and A are their own <strong>du</strong>al. For this <strong>du</strong>ality pairing, X−p is the <strong>du</strong>al<br />

of Xp.<br />

For any p ∈ R, the control operator B is said admissib<strong>le</strong> in Xp and fast control<strong>la</strong>bility is<br />

said to hold in Xp if B satisfies (9) and the property (i) of Lemma 1 holds for all positive T ,<br />

respectively, with X and its norm rep<strong>la</strong>ced by Xp and its norm. In this case, the admissibility<br />

constant KT and the cost CT are <strong>de</strong>noted by Kp,T and Cp,T .<br />

Proposition 3. For all real numbers p and p ′ such that p ′ p:<br />

• Admissibility in Xp implies admissibility in Xp ′. Conversely, if B ∈ L(U, Xp ′) and p′ ><br />

p − 1/2 then B is admissib<strong>le</strong> in Xp.<br />

• Fast control<strong>la</strong>bility in Xp implies fast control<strong>la</strong>bility in Xp ′. Moreover, if there are positive<br />

constants β, C1 and C2 such that Cp,T C2 exp(C1/T β ), then there are positive constants<br />

C ′ 1 and C′ 2 such that Cp ′ ,T C ′ 2 exp(C′ 1 /T β ).<br />

Proof. Since A −1 ∈ L(X), Xp ⊂ Xp ′ continuously which proves the first statement.<br />

Since e tA is an analytic <strong>semi</strong>group, it satisfies the smoothing property: ∀n >0, ∀m ∈ R,<br />

Sn := sup t>0 t n A n e tA L(Xm) < ∞.


By <strong>du</strong>ality, Kp,T is also <strong>de</strong>fined by<br />

Therefore<br />

L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 603<br />

∀x ∈ X−p ′,<br />

T<br />

0<br />

Kp,T C 2 L(X −p ′ ,U) S2 p−p ′<br />

<br />

Ce tA x 2 dt Kp,T x 2 −p .<br />

T<br />

0<br />

t 2(p′ −p) dt<br />

is finite for 2(p ′ − p) > −1.<br />

By <strong>du</strong>ality, Cp,T is also <strong>de</strong>fined by the observability inequality:<br />

∀x ∈ X−p,<br />

Therefore Cp ′ ,2T Sp−p ′Cp,T /T p−p′ . ✷<br />

2.3. Boundary control<strong>la</strong>bility<br />

<br />

e T A x −p Cp,T<br />

<br />

Ce tA L2 . (0,T ;U)<br />

This section concerns the following boundary control version of the p<strong>la</strong>te equation with square<br />

root damping (18):<br />

¨ζ − ρ˙ζ + 2 ζ = 0 onR+ × M, ζ = 0, ζ = χΓ u on R+ × ∂M,<br />

ζ(0) = ζ0 ∈ H 1 0 (M), ˙ζ(0) = ζ1 ∈ H −1 (M), u ∈ L 2 loc (R+ × M), (20)<br />

where χΓ <strong>de</strong>notes the restriction to the boundary followed by the multiplication by the characteristic<br />

function of an open subset Γ =∅of ∂M. A key ingredient of the control transmutation<br />

method of Section 2.4 is the so-cal<strong>le</strong>d “fundamental control<strong>le</strong>d solution” for (18). It is constructed<br />

in Corol<strong>la</strong>ry 2 of Theorem 3 in this section which applies [23] to estimate the cost of<br />

fast boundary controls for (20) when M is a (Eucli<strong>de</strong>an) segment.<br />

We first adapt the abstract <strong>du</strong>ality framework of Section 1.1 to this boundary control system.<br />

Here A =−, H0 = L2 (M), U = L2 (Γ ) and α = 1/2. It is convenient to use the state space<br />

X = H1 × H−1 with the <strong>du</strong>ality pairing<br />

<br />

(ζ0,ζ1), (z0,z1) =〈ζ1,z0〉0 +〈ζ0,z1〉0 + ρ A α ζ0,A α <br />

z0 0 .<br />

With respect to this pairing, X and A are their own <strong>du</strong>al, X−1 = H−1 × H−3 is the <strong>du</strong>al of<br />

X1 = H3 × H1. Multiplying by ζ(T − t) the <strong>du</strong>al homogeneous equation<br />

¨z − ρ˙z + 2 z = 0 onR+ × M, z = z = 0 onR+ × ∂M, (21)<br />

and integrating by parts on (0,T)× M, yields that the control operator B (arising when rewriting<br />

the second-or<strong>de</strong>r system (20) as the first-or<strong>de</strong>r system (6) on ξ(t) = (ζ(t), ˙ζ(t))) is the <strong>du</strong>al<br />

(with respect to the new pairing) of the Neumann observation operator C ∈ L(X1,U)<strong>de</strong>fined by


604 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

C(z0,z1) = χΓ ∂νz0, where ∂ν is a vector field normal to ∂M. As in the proof of Proposition 3,<br />

the admissibility of B results from the analyticity of e tA and C ∈ L(Xp; U) with p ∈ (1/4, 1/2).<br />

(N.b. C ∈ L(Xp,U) for p>1/4 since Xp = H2p+1 × H2p−1 and χ∂M∂ν ∈ L(Hs; U) for<br />

s>3/2.)<br />

Theorem 3. For all ρ ∈ (0, 2), there are C1 > 0 and C2 > 0 such that for all L>0 and T ∈<br />

]0, inf(π/2,L) 2 ], for all ζ0 and ζ1, there is an input function u such that the solution ζ of (20)<br />

with M = (−L,L) and Γ ={L} satisfies ζ(T)= ˙ζ(T)= 0 and the cost estimate<br />

T<br />

0<br />

u 2<br />

L2 dt C2 exp C1L 2 /T ζ0 2<br />

H 1 +ζ1 2<br />

H −1<br />

<br />

.<br />

Proof. By Lemma 1, it is enough to prove that the solution of (21) for any initial data z(0) ∈<br />

H 1 0 (−L,L) and ˙z(0) ∈ H −1 (−L,L) satisfies the observation inequality<br />

<br />

z(T ) 2<br />

H 1 + ˙z(T ) 2<br />

H −1 C2 exp(C1/T)<br />

T<br />

0<br />

<br />

∂sz(t, L) 2 ds.<br />

The scaling (t, x) ↦→ (σ 2 t,σx) re<strong>du</strong>ces the prob<strong>le</strong>m to the case L = π. Using the explicit eigenvalues<br />

and eigenfunctions of on M = (−π,π), this inequality becomes a “window prob<strong>le</strong>m”<br />

for series of comp<strong>le</strong>x exponentials which is almost the one-dimensional setting of “vibrational<br />

control with structural damping” consi<strong>de</strong>red in [23, Section 6], in<strong>de</strong>ed simp<strong>le</strong>r because more<br />

explicit. Therefore [23, Theorem 1] applies and comp<strong>le</strong>tes the proof of Theorem 3. ✷<br />

Corol<strong>la</strong>ry 2. For all ρ ∈ (0, 2) there are positive constants Cρ and C ′ ρ such that, ∀L >0,<br />

∀T ∈ (0, 1], there is a “fundamental control<strong>le</strong>d solution” k in C0 ([0,T]; H −1 (]−L,L[)) ∩<br />

C1 ([0,T]; H −3 (]−L,L[)) satisfying:<br />

T<br />

0<br />

∂ 2 t k − ρ∂2 s ∂tk + ∂ 4 s k = 0 in D′ ]0,T[×]−L,L[ ,<br />

k⌉t=0 = δ, ∂tk⌉t=0 = δ ′<br />

and k⌉t=T = ∂tk⌉t=T = 0,<br />

<br />

k(t,·) 2 H −1 (]−L,L[) + ∂tk(t,·) 2 H −3 ′<br />

dt C (]−L,L[) ρe CρL2 /T<br />

.<br />

Proof. The fast control<strong>la</strong>bility in X = H 1 0 (−L,L) × H −1 (−L,L) stated in Theorem 3 implies,<br />

by Proposition 3, the fast control<strong>la</strong>bility in X−1 = H −1 (−L,L) × H −3 (−L,L) with the same<br />

form of cost estimate. Since the Dirac mass at the origin δ is in H s (R) for all s


L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 605<br />

the control<strong>la</strong>bility cost of a system is not changed by taking its tensor pro<strong>du</strong>ct with a unitary<br />

group). It applies in particu<strong>la</strong>r when M is a rectang<strong>le</strong> or an infinite strip in the p<strong>la</strong>ne control<strong>le</strong>d<br />

from one si<strong>de</strong> (this control<strong>la</strong>bility prob<strong>le</strong>m in a rectang<strong>le</strong> with other boundary conditions was<br />

solved in [11] without the cost estimate, which was ad<strong>de</strong>d <strong>la</strong>ter at the end of [23]). N.b. in this<br />

examp<strong>le</strong>, the condition LΓ < ∞ of [5] required in Theorem 2 is not satisfied.<br />

Corol<strong>la</strong>ry 3. Let ˜M <strong>de</strong>note another smooth comp<strong>le</strong>te Riemannian manifold. For all ρ ∈ (0, 2),<br />

there are C1 > 0 and C2 > 0 such that, for all L>0 and T ∈ (0, 1] for all ζ0 and ζ1, there is<br />

an input function u such that the solution ζ of (20) with M = (−L,L) × ˜M and Γ ={L}×∂ ˜M<br />

satisfies ζ(T)= ˙ζ(T)= 0 and the cost estimate:<br />

T<br />

0<br />

u 2<br />

L2 dt C2 exp C1L 2 /T ζ0 2<br />

H 1 +ζ1 2<br />

H −1<br />

<br />

.<br />

Proof. Let (s, y) <strong>de</strong>note the variab<strong>le</strong> on M = (−L,L) × ˜M. Denoting respectively by s<br />

and y the Dirich<strong>le</strong>t Lap<strong>la</strong>cians on the segment (−L,L) and on ˜M, wehave = s + y.<br />

Since is boun<strong>de</strong>dly invertib<strong>le</strong>, (20) may also be restated as a first-or<strong>de</strong>r system on X =<br />

H −1 (M) × H −1 (M) by setting ξ(t) = (ζ(t), ˙ζ(t)). Then the <strong>semi</strong>group generator A of the<br />

<strong>du</strong>al homogeneous system (7) becomes:<br />

<br />

0 1<br />

A = R with R =<br />

, and e<br />

−1 −ρ<br />

tA = e tsR tyR tyR tsR<br />

e = e e .<br />

The observation operator Cs <strong>de</strong>fined by Csx = χΓ ∂νz = ∂sz⌉s=L commutes with e tyR . We shall<br />

estimate the cost by the <strong>du</strong>ality in Lemma 1. Fix the initial state x0 ∈ X and T>0. Applying to<br />

s ↦→ (e TyR x0)(s, y) for fixed y the observability inequality corresponding to Theorem 3 yields,<br />

with C ′ T := C2 exp(C1L 2 /T):<br />

<br />

0<br />

L<br />

<br />

e TsR<br />

<br />

TyR<br />

e x02<br />

ds C ′ T<br />

T<br />

0<br />

<br />

0<br />

L<br />

<br />

Cse tsR<br />

<br />

TyR<br />

e x02<br />

dsdt.<br />

Integrating this inequality over ˜M yields (the first and <strong>la</strong>st step use Fubini’s theorem and the<br />

commutation of operators acting separately on s and y, the second step uses that e tyR is a<br />

contraction):<br />

<br />

M<br />

<br />

e T A x0<br />

<br />

2 dsdy C ′ T<br />

C ′ T<br />

T<br />

0<br />

T<br />

0<br />

L<br />

0<br />

L<br />

0<br />

<br />

˜M<br />

<br />

˜M<br />

<br />

e TyR<br />

Cse tsR<br />

<br />

x02<br />

dydsdt<br />

<br />

e tyR<br />

Cse tsR<br />

<br />

x02<br />

dydsdt = C ′ T<br />

This is the observability inequality corresponding to Corol<strong>la</strong>ry 3. ✷<br />

T<br />

0<br />

<br />

M<br />

<br />

Cse tA <br />

x02<br />

dsdydt.


606 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

2.4. Proof of Theorem 2<br />

The first part of Theorem 2 is Corol<strong>la</strong>ry 1 for α = 1/2. We shall now prove the second part of<br />

Theorem 2 by the transmutation control method (cf. [17,19]). According to Proposition 3, it is<br />

sufficient to consi<strong>de</strong>r initial data in the space X2 = H6 × H4 which is smoother than the energy<br />

space c<strong>la</strong>imed in Theorem 2.<br />

It results from the work of Bardos–Lebeau–Rauch that (n.b. the control time and the time<br />

variab<strong>le</strong> are <strong>de</strong>noted by L and s here):<br />

Theorem 4. [4,5] Let L>LΩ. For all (w0,w1) and (w2,w3) in H 4 (M) ∩ H 1 0 (M) × H 3 (M) ∩<br />

H 1 0 (M) there is an input function v ∈ H 3 (]0,L[×M) supported in ]0,L[×Ω such that the solution<br />

w ∈ <br />

n∈N Cn ([0,L]; H 4−n (M)) of<br />

∂ 2 s w − w = v in ]0,L[×M, w = 0 on ]0,L[×∂M,<br />

with Cauchy data (w, ∂sw) = (w0,w1) at s = 0, satisfies (w, ∂sw) = (w2,w3) at s = L. Moreover,<br />

the operator SW <strong>de</strong>fined by SW ((w0,w1), (w2,w3)) = v is continuous in the corresponding<br />

norms.<br />

Let T ∈ (0, 1] and L>LΩ be fixed from now on.<br />

Let (ζ0,ζ1) ∈ X2 = H6 × H4 be an initial data for the p<strong>la</strong>te equation (18). Let v± and w± be<br />

the input function and solution for the wave equation obtained from Theorem 4 with w0 = ζ0,<br />

w1 =±ζ1 and w2 = w3 = 0. Let w(±s,·) = w±(s, ·) and v(±s,·) = v±(s, ·) for s ∈[0,L]. We<br />

<strong>de</strong>fine w ∈ <br />

n∈N Cn (R; H 4−n (M)) and v ∈ H 3 (R × M) as the extensions of w and v by zero<br />

outsi<strong>de</strong> [−L,L]×M. They inherit from w± and v± the properties:<br />

∂ 2 s w − w = v in D′ (R × M), w = 0 onR × ∂M,<br />

(w,∂sw )⌉s=0 = (ζ0,ζ1) and (w,∂sw )⌉s=±L = (0, 0). (22)<br />

Let k, Cρ and C ′ ρ be the fundamental control<strong>le</strong>d solution and corresponding constants given<br />

by Corol<strong>la</strong>ry 2. We <strong>de</strong>fine k as the extension of k by zero outsi<strong>de</strong> ]0,T[×]−L,L[. It inherits<br />

from k the following properties:<br />

∂ 2 t k − ρ∂2 s ∂tk + ∂ 4 s k = 0 inD′ ]0,T[×]−L,L[ , (23)<br />

k⌉t=0 = δ, ∂tk⌉t=0 = δ ′<br />

and k⌉t=T = ∂tk ⌉t=T = 0,<br />

T<br />

0<br />

<br />

k(t, ·) 2 H −1 (R) + ∂tk(t, ·) 2 H −3 ′<br />

dt C (R) ρe CρL2 /T<br />

. (24)<br />

The princip<strong>le</strong> of the control transmutation method is to use k as a kernel to transmute<br />

w and v into a solution ζ and an input function u for (18). Since k ∈ C 0 (R+; H −1 (R)) ∩<br />

C 1 (R+; H −3 (R)), w ∈ H 1 (R; H 3 (M)) ∩ H 3 (R; H 1 (M)) and v ∈ H 1 (R; H 2 (M)) ∩<br />

H 3 (R; L 2 (M)), the transmutation formu<strong>la</strong>s


L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608 607<br />

<br />

u(t, x) =<br />

R<br />

<br />

ζ(t,x)=<br />

R<br />

k(t, s)w(s, x) ds,<br />

ρ∂tk(t, s)v(s, x) + k(t, s) ∂ 2 s + v(s, x) ds<br />

<strong>de</strong>fine functions ζ ∈ C 0 (R+; H 3 (M)) ∩ C 1 (R+; H 1 (M)) and u ∈ L 2 (R+ × M).Thisζ satisfies<br />

the required initial conditions: k ⌉t=0 = δ and w ⌉s=0 = ζ0 imply ζ⌉t=0 = ζ0; ∂tk ⌉t=0 = ∂sδ and<br />

∂sw ⌉s=0 = ζ1 imply ∂tζ⌉t=0 = ζ1 by integrating by parts. This ζ satisfies the required final conditions:<br />

k ⌉t=T = ∂tk ⌉t=T = 0 implies ζ⌉t=T = ∂tζ⌉t=T = 0. This ζ satisfies the required boundary<br />

conditions: w ⌉∂M = 0 implies ζ⌉∂M = 0 and w ⌉∂M = ∂ 2 s w ⌉∂M = 0 implies ζ⌉∂M = 0. The<br />

input u is supported in [0,T]×Ω since k is supported in [0,T]×(−L,L) and v is supported<br />

in (−L,L) × Ω. These ζ and u satisfy the p<strong>la</strong>te equation (18): using (22) in the second step,<br />

integration by parts in the third, and (23) in the fourth,<br />

¨ζ − ρ˙ζ + 2 ζ<br />

<br />

= ∂ 2 t k w − ρ∂tkw + k 2 w<br />

<br />

=<br />

∂ 2 t k w − ρ∂tk ∂ 2 s w − v + k ∂ 2 2<br />

s ∂s w − v − v <br />

<br />

∂2 = t k − ρ∂ 2 s ∂tk + ∂ 4 s k w + ρ∂tk + k ∂ 2 s + v = u = χΩu.<br />

Finally, the cost estimate<br />

u 2<br />

L2 (R×M) C2 exp CρL 2 /T ζ0 2<br />

H 6 +ζ1 2<br />

H 4<br />

<br />

results from (24),<br />

v 2<br />

H 3 (R×M) 2SW 2 ζ0 2<br />

H 4 2<br />

+ζ1 (M) H 3 <br />

and<br />

(M)<br />

uL2 (R×M) ρ∂tk L2 (R;H −3 (R)) vH 3 (R;L2 (M)) +kL2 (R;H −1 (R)) vH 1 (R;H 2 (M)) .<br />

Note ad<strong>de</strong>d in proof<br />

After our artic<strong>le</strong> was accepted, we became aware of a paper to appear in Asymptotic Analysis:<br />

“Internal null-control<strong>la</strong>bility for a structurally damped beam equation” by Julian Edward and<br />

Louis Tebou. This paper concerns (18) when M is a segment and focuses on the limit ρ → 0.<br />

We c<strong>la</strong>im that our Theorem 2 generalizes the main result of this paper (n.b. LΩ < ∞ always<br />

hold when the dimension of M is one). Since Theorem 1 of this paper says that the cost does<br />

not <strong>de</strong>pend on ρ


608 L. Mil<strong>le</strong>r / Journal of Functional Analysis 236 (2006) 592–608<br />

with α = 1/2) implies |λ±|=μ so that |c| does not <strong>de</strong>pend on ρ here, which comp<strong>le</strong>tes the proof<br />

of our c<strong>la</strong>im.<br />

References<br />

[1] A. Atal<strong>la</strong>h-Baraket, C. Fermanian Kammerer, High frequency analysis of families of solutions to the equation of<br />

viscoe<strong>la</strong>sticity of Kelvin–Voigt, J. Hyperbolic Differ. Equ. 1 (4) (2004) 789–812.<br />

[2] G. Avalos, I. Lasiecka, Optimal blowup rates for the minimal energy null control of the strongly damped abstract<br />

wave equation, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 2 (3) (2003) 601–616.<br />

[3] A.V. Ba<strong>la</strong>krishnan, Damping operators in continuum mo<strong>de</strong>ls of f<strong>le</strong>xib<strong>le</strong> structures: Explicit mo<strong>de</strong>ls for proportional<br />

damping in beam bending with end-bodies, Appl. Math. Optim. 21 (3) (1990) 315–334.<br />

[4] C. Bardos, G. Lebeau, J. Rauch, Un exemp<strong>le</strong> d’utilisation <strong>de</strong>s notions <strong>de</strong> propagation pour <strong>le</strong> contrô<strong>le</strong> et <strong>la</strong> stabilisation<br />

<strong>de</strong> problèmes hyperboliques, in: Nonlinear Hyperbolic Equations in Applied Sciences, 1987, Rend. Sem. Mat.<br />

Univ. Politec. Torino (1988) 11–31 (special issue).<br />

[5] C. Bardos, G. Lebeau, J. Rauch, Sharp sufficient conditions for the observation, control, and stabilization of waves<br />

from the boundary, SIAM J. Control Optim. 30 (5) (1992) 1024–1065.<br />

[6] G. Chen, D.L. Russell, A mathematical mo<strong>de</strong>l for linear e<strong>la</strong>stic systems with structural damping, Quart. Appl.<br />

Math. 39 (4) (1982) 433–454.<br />

[7] S.P. Chen, R. Triggiani, Proof of extensions of two conjectures on structural damping for e<strong>la</strong>stic systems, Pacific J.<br />

Math. 136 (1) (1989) 15–55.<br />

[8] S.P. Chen, R. Triggiani, Gevrey c<strong>la</strong>ss <strong>semi</strong>groups arising from e<strong>la</strong>stic systems with gent<strong>le</strong> dissipation: The case<br />

0


Journal of Functional Analysis 236 (2006) 609–629<br />

www.elsevier.com/locate/jfa<br />

On action of diffeomorphisms of C*-algebras on<br />

<strong>de</strong>rivations<br />

Edward Kissin<br />

Department of Computing, Communications Technology and Mathematics, London Metropolitan University,<br />

166-220 Holloway Road, London N7 8DB, UK<br />

Received 9 January 2006; accepted 13 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 18 April 2006<br />

Communicated by J. Cuntz<br />

Abstract<br />

In this paper we consi<strong>de</strong>r automorphisms of the domains of closed *-<strong>de</strong>rivations of C*-algebras and show<br />

that they extend to automorphisms of C*-algebras, so we call them diffeomorphisms. The diffeomorphisms<br />

generate transformations of the sets of closed *-<strong>de</strong>rivations of C*-algebras. In this paper we study the subgroups<br />

of diffeomorphisms that <strong>de</strong>fine “boun<strong>de</strong>d” shifts of <strong>de</strong>rivations and the subgroups of the stabilizers<br />

of <strong>de</strong>rivations.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Derivations; C*-algebras; Automorphisms; Diffeomorphisms<br />

1. Intro<strong>du</strong>ction<br />

Extensive <strong>de</strong>velopment of non-commutative geometry requires e<strong>la</strong>borating of the theory of the<br />

domains of closed *-<strong>de</strong>rivations of C*-algebras whose properties in many respects are analogous<br />

to the properties of algebras of differentiab<strong>le</strong> functions. In this paper we consi<strong>de</strong>r automorphisms<br />

of the domains of <strong>de</strong>rivations and show that they extend to automorphisms of C*-algebras, so we<br />

call them diffeomorphisms. The diffeomorphisms generate transformations of the sets of closed<br />

*-<strong>de</strong>rivations of C*-algebras. In this paper we study the subgroups of diffeomorphisms that <strong>de</strong>fine<br />

“boun<strong>de</strong>d” shifts of <strong>de</strong>rivations and the subgroups of the stabilizers of <strong>de</strong>rivations.<br />

E-mail address: e.kissin@londonmet.ac.uk.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.009


610 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

Throughout the paper we <strong>de</strong>note by (A, ·) a C*-algebra. A closed linear map δ from a<br />

<strong>de</strong>nse *-subalgebra D(δ) of A into A is cal<strong>le</strong>d a closed *-<strong>de</strong>rivation if<br />

δ(AB) = Aδ(B) + δ(A)B and δ A ∗ = δ(A) ∗<br />

for A,B ∈ D(δ).<br />

The subalgebra D(δ) is cal<strong>le</strong>d the domain of δ; δ is boun<strong>de</strong>d if and only if D(δ) = A.<br />

Let A be a <strong>de</strong>nse *-subalgebra of A. Denote by Der(A) the set of all closed *-<strong>de</strong>rivations<br />

δ on A with A = D(δ). We call A a domain if Der(A) = ∅. In Section 2 we show that all<br />

*-automorphisms of a domain A of A extend to *-automorphisms of A. We call these extensions<br />

diffeomorphisms of A; they form a group that we <strong>de</strong>note by Dif(A). Each diffeomorphism φ<br />

<strong>de</strong>fines a transformation Tφ of Der(A): for every δ ∈ Der(A), the <strong>de</strong>rivation Tφ(δ) = φ−1δφ also<br />

belongs to Der(A).ThemapT : φ → Tφ is an antirepresentation of the group Dif(A) into the set<br />

of all transformations of Der(A): Tφθ = Tθ Tφ. We <strong>de</strong>note by Z(δ) the stabilizer of δ:<br />

Z(δ) = φ ∈ Dif(A): δ = Tφ(δ) <br />

and by B(δ) the subgroup of Dif(A) of diffeomorphisms that <strong>de</strong>fine boun<strong>de</strong>d shifts of δ:<br />

B(δ) = φ ∈ Dif(A): the <strong>de</strong>rivation Tφ(δ) − δ is boun<strong>de</strong>d on A in · .<br />

Denote by B(H) the algebra of all boun<strong>de</strong>d operators on a Hilbert space H and by C(H) the<br />

i<strong>de</strong>al of all compact operators. In this paper we study the structure of the groups Z(δ) and B(δ),<br />

for δ ∈ Der(A), when A are domains of C*-subalgebras A of B(H) that contain C(H).<br />

An operator F on H with the <strong>de</strong>nse domain D(F) imp<strong>le</strong>ments δ ∈ Der(A) if<br />

AD(F ) ⊆ D(F) and δ(A)|D(F) = i[F,A]|D(F) = i(FA − AF )|D(F) for all A ∈ A. (1.1)<br />

Bratteli and Robinson proved in [2] that, if C(H) ⊆ A ⊆ B(H) and A is a domain of A, then each<br />

δ ∈ Der(A) has a symmetric imp<strong>le</strong>mentation: a closed symmetric operator S on H that imp<strong>le</strong>ments<br />

δ. The operator S can be chosen (see [5, Theorem 27.21]) to be a minimal imp<strong>le</strong>mentation,<br />

that is, for each closed operator F that imp<strong>le</strong>ments δ,<br />

S + t1|D(S) ⊆ F for some t ∈ C.<br />

With each closed symmetric operator S on H , we associate a *-subalgebra<br />

AS = A ∈ B(H): AD(S) ⊆ D(S), A ∗ D(S) ⊆ D(S) and [S,A]|D(S) is boun<strong>de</strong>d (1.2)<br />

of B(H). It is the domain (see [5]) of a closed *-<strong>de</strong>rivation δS into B(H) <strong>de</strong>fined by<br />

δS(A) = i[S,A] for A ∈ AS,<br />

where [S,A] is the clo<strong>sur</strong>e of [S,A]|D(S) = (SA − AS)|D(S). Furthermore, AS = B(H) if and<br />

only if S is boun<strong>de</strong>d. If S is unboun<strong>de</strong>d, δS is unboun<strong>de</strong>d and AS is a Hermitian <strong>semi</strong>simp<strong>le</strong><br />

Banach *-algebra with respect to the norm<br />

AδS =A+ δS(A) for A ∈ AS.


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 611<br />

Denote by FS the clo<strong>sur</strong>e in ·δS of the set of all finite rank operators in AS and set<br />

JS = A ∈ AS ∩ C(H): δS(A) ∈ C(H) .<br />

Then (see [5]) FS and JS are domains of C(H) and closed *-i<strong>de</strong>als of AS. The *-<strong>de</strong>rivations<br />

δ min<br />

S = δS|FS and δ max<br />

S = δS|JS<br />

of C(H) are closed; they are the minimal and the maximal closed *-<strong>de</strong>rivations of C(H) with<br />

minimal imp<strong>le</strong>mentation S. It was proved in [6] that the clo<strong>sur</strong>e of (JS) 2 in ·δS coinci<strong>de</strong>s with<br />

FS and that JS = FS if S is selfadjoint.<br />

In Section 3 we establish a link between minimal symmetric imp<strong>le</strong>mentations of two <strong>de</strong>riva-<br />

tions from Der(A). We prove that if S and T are such imp<strong>le</strong>mentations, then the algebras FS<br />

and FT coinci<strong>de</strong> and the norms ·δS and ·δT on them are equiva<strong>le</strong>nt. It was shown in [7]<br />

that if these norms are equal then S − t1 =±UTU∗ for some unitary operator U and t ∈ R.<br />

In Section 3 we consi<strong>de</strong>r the general case and obtain some necessary conditions that S and T<br />

satisfy.<br />

Denote by US the group of all unitary operators in the algebra AS and set<br />

ZS = U ∈ US: δS(U) = λU for some λ ∈ C .<br />

We show in Section 4 that if C(H) ⊆ A ⊆ B(H) and A is a domain of A, then each φ ∈ Dif(A) is<br />

imp<strong>le</strong>mented by a unitary operator Uφ: φ(A) = UφAU ∗ φ for all A ∈ A. Moreover, if δ ∈ Der(A)<br />

then φ ∈ B(δ) if and only if Uφ ∈ US, and φ ∈ Z(δ) if and only if Uφ ∈ ZS, where S is a minimal<br />

symmetric imp<strong>le</strong>mentation of δ. I<strong>de</strong>ntifying Dif(A), B(δ) and Z(δ) with the corresponding<br />

subgroups of unitary operators, we have<br />

B(δ) = Dif(A) ∩ US and Z(δ) = Dif(A) ∩ ZS.<br />

Section 5 is <strong>de</strong>voted to the investigation of the structure of the groups ZS. In Section 6 we<br />

study the prob<strong>le</strong>m of constructing domains of C*-algebras that extend the domains JS.LetAbe a domain of a C*-subalgebra A of B(H) and <strong>le</strong>t C(H) A. Assume that there is a <strong>de</strong>rivation<br />

in Der(A) imp<strong>le</strong>mented by a symmetric operator S. Then A + JS is a <strong>de</strong>nse *-subalgebra of the<br />

C*-algebra A + C(H) and δ = δS|(A + JS) is a *-<strong>de</strong>rivation of A + C(H). We provi<strong>de</strong> some<br />

sufficient conditions for δ to be a closed <strong>de</strong>rivation which implies that A + JS is a domain of A +<br />

C(H). Numerous examp<strong>le</strong>s of such domains can be obtained by consi<strong>de</strong>ring the *-commutant<br />

CS = Ker δS = A ∈ AS: δS(A) = 0 <br />

of S. It is a W*-algebra and we prove that, for each C*-subalgebra A of CS satisfying some<br />

simp<strong>le</strong> conditions, the algebra A + JS is a domain of the C*-algebra A + C(H). In particu<strong>la</strong>r,<br />

CS + JS is a domain of the C*-algebra CS + C(H). Finally, we show that, for each symmetric<br />

operator S,<br />

B δ min<br />

S<br />

= B δ max<br />

S<br />

<br />

= US and Z δ min<br />

S<br />

All symmetric operators in this paper are assumed to be closed.<br />

max<br />

= Z δ = Z(δ) = ZS where δ = δS|(CS + JS).<br />

S


612 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

2. Extension of automorphisms from subalgebras of C*-algebras<br />

Let A be a <strong>de</strong>nse *-subalgebra of a unital C*-algebra A. It is cal<strong>le</strong>d a Q-subalgebra of A if<br />

1 ∈ A and Sp A (A) = Sp A (A) for all A ∈ A. (2.1)<br />

If A is a <strong>de</strong>nse *-subalgebra of a non-unital C*-algebra A, consi<strong>de</strong>r the unitizations A = A + C1<br />

of A and A = A + C1 of A. The algebra A is a Q-subalgebra of A if<br />

Sp A (A) = SpA (A) for all A ∈ A.<br />

The domains of closed *-<strong>de</strong>rivations of A are Q-subalgebras of A (see [2,5]).<br />

Proposition 2.1. Let A be a Q-subalgebra of a C ∗ -algebra A and <strong>le</strong>t φ be a ∗ -automorphism<br />

of A. Then φ=1,soφ extends to a ∗ -automorphism of A.<br />

Proof. Let A be unital. Since SpA (A) = SpA (φ(A)), forA∈A,wehave <br />

SpA (A) = SpA (A) = SpA φ(A) = SpA φ(A) . (2.2)<br />

If A = A ∗ ∈ A then φ(A) ∗ = φ(A ∗ ) = φ(A) and, by (2.2),<br />

Hence, for B ∈ A,<br />

A= sup |λ|= sup |λ|=<br />

λ∈SpA (A) λ∈SpA (φ(A))<br />

φ(A) .<br />

B 2 = B ∗ B = φ B ∗ B = φ(B) ∗ φ(B) = φ(B) 2 .<br />

For non-unital A, we have the proof by rep<strong>la</strong>cing in the above argument A by A and A by A. ✷<br />

For a Q-subalgebra A of a C ∗ -algebra A, <strong>de</strong>note by Der(A) the set of all closed unboun<strong>de</strong>d<br />

*-<strong>de</strong>rivations δ on A with A = D(δ). We call A a domain if Der(A) = ∅. We call a<br />

∗ -automorphism φ of A a diffeomorphism, if it preserves a domain A in A and <strong>de</strong>note by Dif(A)<br />

the group of all diffeomorphisms of A that preserve A. Proposition 2.1 yields<br />

Corol<strong>la</strong>ry 2.2. φ → φ|A is an isomorphism from Dif(A) onto the set of all ∗ -automorphisms<br />

of A.<br />

Any domain A is a Hermitian <strong>semi</strong>simp<strong>le</strong> Banach *-algebra (see [5]) with respect to each<br />

norm<br />

Aδ =A+ δ(A) for A ∈ A, where δ ∈ Der(A).<br />

For each boun<strong>de</strong>d <strong>de</strong>rivation δb on A, δ + δb ∈ Der(A). Johnson’s uniqueness of norm theorem<br />

yields<br />

Proposition 2.3. All norms ·δ, δ ∈ Der(A), on a domain A are equiva<strong>le</strong>nt.


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 613<br />

Each φ ∈ Dif(A) <strong>de</strong>fines a transformation Tφ of Der(A) by the formu<strong>la</strong><br />

Tφ(δ) = δφ = φ −1 δφ|A, for δ ∈ Der(A).<br />

Then Tφψ = TψTφ, soT : φ → Tφ is an antirepresentation of the group Dif(A) into the set of all<br />

transformations of Der(A). Denote by Z(δ) the stabilizer of δ in Dif(A):<br />

Z(δ) = <br />

φ ∈ Dif(A): δ = δφ<br />

and by B(δ) the subgroup of Dif(A) of diffeomorphisms which <strong>de</strong>fine boun<strong>de</strong>d shifts of δ:<br />

B(δ) = φ ∈ Dif(A): the <strong>de</strong>rivation δφ − δ is boun<strong>de</strong>d on A in · .<br />

If ψ ∈ B(δ) then B(δ) = B(δψ). Denote by A ∗ the <strong>du</strong>al space of A.<br />

Proposition 2.4. Let δ ∈ Der(A) and φ ∈ Dif(A). If there exists Δ ∈ Der(A) such that, for each<br />

A ∈ A and F ∈ A∗ , F(Δφn(A)) → F(δ(A)),asn→∞, then φ ∈ Z(δ).<br />

Proof. Define Fφ−1 by Fφ−1(A) = F(φ−1 (A)), forA∈A. Then Fφ−1 ∈ A∗ ,so,forA∈A, F Δφn+1(A) <br />

= Fφ−1 Δφn <br />

φ(A) → Fφ−1 δ φ(A)<br />

= F φ −1 δ φ(A) = F δφ(A) .<br />

Since F(Δ φ n+1(A)) → F(δ(A)), we have F(δφ(A)) = F(δ(A)). Thus δφ(A) = δ(A), so<br />

φ ∈ Z(δ). ✷<br />

3. Domains of C*-algebras containing C(H)<br />

For x,y ∈ H , the rank one operator x ⊗ y on H acts by the formu<strong>la</strong><br />

(x ⊗ y)z = (z, x)y for z ∈ H, and x ⊗ y=xy.<br />

Let F be an operator on H .Foru, v ∈ H ,<br />

(x ⊗ y)(u ⊗ v) = (v, x)(u ⊗ y), (x ⊗ y) ∗ = y ⊗ x,<br />

x ⊗ λy = λ(x ⊗ y) = λx ⊗ y,<br />

F(x⊗ y) = x ⊗ Fy, (x ⊗ y)F = F ∗ x ⊗ y, if y ∈ D(F), x ∈ D F ∗ . (3.1)<br />

For an algebra of operators A, <strong>de</strong>note by F(A) the subalgebra of all finite rank operators in A.<br />

Lemma 3.1. Let A be a domain of A and C(H) ⊆ A ⊆ B(H). Forδ ∈ Der(A), <strong>le</strong>t a symmetric<br />

operator S be its minimal imp<strong>le</strong>mentation. Then<br />

(i) the set of all rank one operators in A consists of all y ⊗ x with x,y ∈ D(S);<br />

(ii) F(A) ={ n i=1 xi ⊗ yi: xi,yi ∈ D(S)}=F(AS).


614 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

Proof. First <strong>le</strong>t us show that the set<br />

Eδ ={x ∈ H : x ⊗ x ∈ A}<br />

is a <strong>de</strong>nse linear subspace of H and each rank one operator in A has form y ⊗ x, forx,y ∈ Eδ.<br />

For each x ∈ H , x=1, the rank one projection x ⊗x belongs to A. It follows from [9, Proposition<br />

3.4.9] that, for every ε>0, there is a projection pε ∈ D(δ) = A such that x ⊗ x − pε


Using it, we obtain as in (3.2) that<br />

E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 615<br />

δ(yn ⊗ yn) = i(yn ⊗ Syn − Syn ⊗ yn) → i(y ⊗ Sy − Sy ⊗ y).<br />

Since δ is a closed <strong>de</strong>rivation, y ⊗y ∈ D(δ) = A. Hence y ∈ Eδ,soEδ = D(S). Part (i) is proved.<br />

C<strong>le</strong>arly, all operators n i=1 xi ⊗ yi, with xi, yi ∈ D(S), belong to F(A). Conversely, each<br />

A ∈ F(A) has form A = n i=1 xi ⊗ yi, where all xi are linearly in<strong>de</strong>pen<strong>de</strong>nt and all yi are<br />

linearly in<strong>de</strong>pen<strong>de</strong>nt. Since S imp<strong>le</strong>ments δ and D(S) is <strong>de</strong>nse in H ,<br />

Az =<br />

n<br />

(z, xi)yi ∈ D(S) for all z ∈ D(S).<br />

i=1<br />

Hence all yi ∈ D(S). As A ∗ = n i=1 yi ⊗ xi ∈ F(A), all xi ∈ D(S). Thus F(A) =<br />

{ n i=1 xi ⊗ yi: xi,yi ∈ D(S)}. From this and from [6, Lemma 3.1] it follows that F(A) =<br />

F(AS). ✷<br />

For δ ∈ Der(A), <strong>de</strong>note by F(A,δ)the clo<strong>sur</strong>e of F(A) in ·δ. Recall that FS is the clo<strong>sur</strong>e<br />

of F(AS) in ·δS .<br />

Corol<strong>la</strong>ry 3.2. Let A be a domain in A, C(H) ⊆ A ⊆ B(H). Let δ,σ ∈ Der(A) and <strong>le</strong>t S,T be,<br />

respectively, their minimal symmetric imp<strong>le</strong>mentations. Then<br />

(i) F(A,δ)is an i<strong>de</strong>al of A isometrically isomorphic to the algebra (FS, ·δS ).<br />

(ii) The algebras F(A,δ)and F(A,σ)coinci<strong>de</strong> and D(S) = D(T ).<br />

Proof. Since S imp<strong>le</strong>ments δ, it follows from (1.1) that A = D(δ) ⊆ AS and δ = δS|D(δ). Hence<br />

the norms ·δS and ·δ coinci<strong>de</strong> on A, so it follows from Lemma 3.1(ii) that F(A,δ)and FS<br />

are isometrically isomorphic. As FS is an i<strong>de</strong>al of AS (see [6]), F(A,δ)is an i<strong>de</strong>al of A.<br />

By Proposition 2.3, the norms ·δ and ·σ on A are equiva<strong>le</strong>nt, so the algebras<br />

F(A,δ) and F(A,σ) coinci<strong>de</strong>. Since A = D(δ) = D(σ), we have from Lemma 3.1(i) that<br />

D(S) = D(T ). ✷<br />

Let S,T be minimal symmetric imp<strong>le</strong>mentations of δ,σ ∈ Der(A). It follows from Proposition<br />

2.3 and Corol<strong>la</strong>ry 3.2 that the algebras FS and FT coinci<strong>de</strong> and the norms ·δS and ·δT<br />

on them are equiva<strong>le</strong>nt. It was shown in [7, Theorem 4.4] that these norms are equal if and only<br />

if S − t1 =±UTU∗ for some t ∈ R and a unitary operator U. Below we consi<strong>de</strong>r the general<br />

case and obtain some necessary conditions that S and T satisfy.<br />

Theorem 3.3. Let A be a domain in A, C(H) ⊆ A ⊆ B(H). Let symmetric operators S and T<br />

be minimal imp<strong>le</strong>mentations of δ,σ ∈ Der(A), respectively. Then<br />

(i) S and T are either both selfadjoint or both non-selfadjoint;<br />

(ii) there exist boun<strong>de</strong>d invertib<strong>le</strong> operators M from (S − i1)D(S) onto (T − i1)D(T ) and N<br />

from (S + i1)D(S) onto (T + i1)D(T ) such that<br />

T − i1 =M(S − i1) and T + i1 = N(S + i1);<br />

(iii) the <strong>de</strong>rivation σ −δ is boun<strong>de</strong>d if and only if T = S +R where R is selfadjoint and boun<strong>de</strong>d.


616 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

Proof. It was shown in [6] that the algebra (FS, ·δS ) has a boun<strong>de</strong>d approximate i<strong>de</strong>ntity if<br />

and only if S is selfadjoint. By Corol<strong>la</strong>ry 3.2, FS = FT and the norms ·δS and ·δT on them<br />

are equiva<strong>le</strong>nt. This yields (i).<br />

By Corol<strong>la</strong>ry 3.2(ii), D(S) = D(T ). Fixx∈D(S) with x=1. By Proposition 2.3, there is<br />

C>0 such that, for all y ∈ D(S),<br />

x ⊗ yσ =x ⊗ y+ σ(x⊗ y) Cx ⊗ yδ = Cx ⊗ y+C δ(x ⊗ y) .<br />

The operators T − i1 and S − i1 imp<strong>le</strong>ment σ and δ.Asx ⊗ y=xy, we have from (3.3)<br />

Therefore<br />

xy+ x ⊗ (T − i1)y − (T + i1)x ⊗ y <br />

C xy+ x ⊗ (S − i1)y − (S + i1)x ⊗ y .<br />

<br />

(T − i1)y = x ⊗ (T − i1)y x ⊗ (T − i1)y − (T + i1)x ⊗ y + (T + i1)x ⊗ y <br />

C xy+ x ⊗ (S − i1)y − (S + i1)x ⊗ y + (T + i1)x y<br />

Cy+C <br />

(S − i1)y + (S + i1)xy+ (T + i1)xy Ky+C (S − i1)y .<br />

Since S is symmetric, (S − i1)y2 =Sy2 +y2 . Hence<br />

<br />

(T − i1)y (K + C) (S − i1)y for y ∈ D(S). (3.4)<br />

It is well known that (S ± i1)D(S) are closed subspaces of H and Ker(S ± i1) ={0}. Define<br />

an operator M from (S −i1)D(S) into (T −i1)D(T ) by M(S−i1)y = (T −i1)y,fory∈ D(S).<br />

By (3.4), M is boun<strong>de</strong>d. Simi<strong>la</strong>rly, the operator R from (T − i1)D(T ) into (S − i1)D(S) <strong>de</strong>fined<br />

by R(T − i1)y = (S − i1)y, fory∈D(T ), is boun<strong>de</strong>d. Hence R = M−1 .<br />

Simi<strong>la</strong>rly, there is a boun<strong>de</strong>d invertib<strong>le</strong> operator N from (S + i1)D(S) on (T + i1)D(T ) such<br />

that T + i1 = N(S + i1). Part (ii) is proved.<br />

For R = R∗ ∈ B(H), the *-<strong>de</strong>rivation δR(A) = i[R,A], A ∈ A, is boun<strong>de</strong>d. Hence δ + δR ∈<br />

Der(A) and S + R is its minimal imp<strong>le</strong>mentation.<br />

Conversely, <strong>le</strong>t σ − δ be boun<strong>de</strong>d. As D(S) = D(T ), the operator R = T − S is symmetric<br />

on D(S). There is C>0 such that σ(A)−δ(A) CA for all A ∈ A. Hence, for all x,y ∈<br />

D(S), we have from (3.3) that x ⊗ y ∈ A and<br />

<br />

<br />

σ(x⊗ y) − δ(x ⊗ y) = i[T − S,x ⊗ y] =x ⊗ Ry − Rx ⊗ y Cx ⊗ y=Cxy.<br />

Fix x with x=1. Then<br />

Ry=x ⊗ Ry x ⊗ Ry − Rx ⊗ y+Rx ⊗ y Cxy+Rxy.<br />

Hence R is boun<strong>de</strong>d on D(S), so it extends to a selfadjoint boun<strong>de</strong>d operator. ✷


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 617<br />

4. Diffeomorphisms of C*-algebras containing C(H)<br />

Each *-automorphism φ of C(H) is imp<strong>le</strong>mented by a unitary operator U: φ(B) = UBU ∗ ,<br />

for B ∈ C(H) (see [8]). This is also true for all *-automorphisms φ of C*-subalgebras A of<br />

B(H) containing C(H). In<strong>de</strong>ed, for x,y,∈ H ,setR = φ(x ⊗ y). By (3.1), for all z, u ∈ H ,<br />

R ∗ z ⊗ Ru = R(z ⊗ u)R = φ (x ⊗ y)φ −1 (z ⊗ u)(x ⊗ y) = φ −1 (z ⊗ u)y, x R.<br />

Hence R is a rank one operator, so φ and φ −1 map finite rank operators into finite rank operators.<br />

Thus φ(C(H)) = C(H) and there is a unitary U such that φ(B) = UBU ∗ ,forB ∈ C(H).For<br />

A ∈ A and all x,y ∈ H ,<br />

Ux ⊗ UAy = U(x⊗ Ay)U ∗ = φ A(x ⊗ y) = φ(A)φ(x⊗ y)<br />

= φ(A)U(x⊗y)U ∗ = φ(A)(Ux ⊗ Uy) = Ux ⊗ φ(A)Uy.<br />

Hence φ(A)Uy = UAy for all y ∈ H ,soφ(A)= UAU∗ for all A ∈ A.<br />

Recall that we <strong>de</strong>note by US the group of all unitary operators in the algebra AS:<br />

US = U ∈ B(H): U is unitary, UD(S)= D(S) and [S,U]|D(S) is boun<strong>de</strong>d <br />

and set<br />

ZS = U ∈ US: δS(U) = λU for some λ ∈ C .<br />

Theorem 4.1. Let A be a domain in A, C(H) ⊆ A ⊆ B(H), and <strong>le</strong>t φ ∈ Dif(A). Let, as above,<br />

a unitary U ∈ B(H) imp<strong>le</strong>ments φ: φ(A)= UAU ∗ for all A ∈ A. Then<br />

(i) if a symmetric operator S is a minimal imp<strong>le</strong>mentation of δ ∈ Der(A), then UD(S)= D(S)<br />

and U ∗ SU is a minimal imp<strong>le</strong>mentation of the ∗ -<strong>de</strong>rivation δφ;<br />

(ii) φ ∈ B(δ) if and only if U ∈ US;<br />

(iii) φ ∈ Z(δ) if and only if U ∈ ZS.<br />

Proof. By Lemma 3.1, x ⊗ x ∈ A for x ∈ D(S). Hence<br />

φ(x ⊗ x) = U(x ⊗ x)U ∗ = Ux ⊗ Ux ∈ A,<br />

so Ux ∈ D(S). Thus UD(S) ⊆ D(S). Since φ −1 ∈ Dif(A) and imp<strong>le</strong>mented by U ∗ ,wehave<br />

U ∗ D(S) ⊆ D(S). Therefore UD(S)= D(S).<br />

Let T imp<strong>le</strong>ment δ. For all A ∈ A,wehaveUAU ∗ ∈ A, soUAU ∗ D(T ) ⊆ D(T ). By (1.1),<br />

δφ(A)|U ∗ D(T ) = φ −1 δ φ(A) U ∗ D(T ) = U ∗ δ UAU ∗ U U ∗ D(T )<br />

= U ∗ δ UAU ∗ D(T ) = U ∗ i T,UAU ∗ D(T ) = i U ∗ TU,A U ∗ D(T ) . (4.1)<br />

Thus U ∗ TU imp<strong>le</strong>ments δφ. Simi<strong>la</strong>rly, if R imp<strong>le</strong>ments δφ, URU ∗ imp<strong>le</strong>ments δ. Hence T →<br />

U ∗ TU is a one-to-one correspon<strong>de</strong>nce between the sets of imp<strong>le</strong>mentations of δ and δφ. Since S<br />

is a minimal imp<strong>le</strong>mentation of δ, U ∗ SU is a minimal imp<strong>le</strong>mentation of δφ. Part (i) is proved.


618 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

It follows from (i) and from Theorem 3.3(iii) that δφ − δ is a boun<strong>de</strong>d <strong>de</strong>rivation if and only<br />

if K = U ∗ SU − S is a boun<strong>de</strong>d operator on D(S). Hence φ ∈ B(δ), if and only if the operator<br />

[S,U]=UK is boun<strong>de</strong>d on D(S). Since UD(S)= D(S), wehavethatφ ∈ B(δ) if and only if<br />

U ∈ US. Part (ii) is proved.<br />

Let φ ∈ Z(δ). Then U ∈ US and, by (i), U ∗ SU is a minimal imp<strong>le</strong>mentation of δφ and<br />

D(U ∗ SU) = D(S). Since δφ = δ and S is a minimal imp<strong>le</strong>mentation of δ, there is λ ∈ C<br />

such that U ∗ SU = S + λ1|D(S). Hence δS(U) = iλU, soU ∈ ZS. Conversely, if U ∈ ZS then<br />

U ∗ SU = S + λ1|D(S). AsU ∗ SU is a minimal imp<strong>le</strong>mentation of δφ, wehaveδφ = δ. ✷<br />

Let A be a domain in A, C(H) ⊆ A ⊆ B(H). It follows from Theorem 4.1 that one can i<strong>de</strong>ntify<br />

(mo<strong>du</strong>lo sca<strong>la</strong>rs from the unit circ<strong>le</strong>) the group Dif(A) with the group of all unitary operators<br />

U on H whose action A → UAU ∗ preserve A. Forδ ∈ Der(A), we will also i<strong>de</strong>ntify the subgroups<br />

B(δ) and Z(δ) with the corresponding subgroups of unitary operators. By Theorem 4.1,<br />

if S is a minimal imp<strong>le</strong>mentation of δ then<br />

B(δ) = U ∈ US: UAU ∗ = A = Dif(A) ∩ US,<br />

Z(δ) = U ∈ ZS: UAU ∗ = A = Dif(A) ∩ ZS. (4.2)<br />

Proposition 4.2. Let A be a domain in A, C(H) ⊆ A ⊆ B(H), and <strong>le</strong>t S be a minimal symmetric<br />

imp<strong>le</strong>mentation of δ ∈ Der(A). Then<br />

(i) B(δ) is closed in (AS, ·δS ) and Z(δ) is closed in (B(H ), ·);<br />

(ii) if A is an i<strong>de</strong>al of AS, then B(δ) = US and Z(δ) = ZS.<br />

Proof. Let a sequence {Un} of unitaries in B(δ) converge to U in (AS, ·δS ). Then<br />

U − Un →0 and δS(U) − δS(Un) →0. Hence U is unitary. For each A ∈ A, wehave<br />

UnAU ∗ n ∈ A, UAU∗ ∈ AS and UAU∗ − UnAU ∗ n →0. Hence<br />

<br />

∗<br />

δS UAU − δ UnAU ∗ <br />

n<br />

= δS(U)AU ∗ + Uδ(A)U ∗ ∗<br />

+ UAδS U − δS(Un)AU ∗ n − Unδ(A)U ∗ n<br />

δS(U)AU ∗ − δS(Un)AU ∗ <br />

<br />

n + Uδ(A)U ∗<br />

− Unδ(A)U ∗ <br />

<br />

n<br />

+ <br />

UAδS<br />

∗<br />

U <br />

− UnAδS → 0.<br />

U ∗ n<br />

<br />

∗<br />

− UnAδS U <br />

n<br />

Since δ is closed, UAU∗ ∈ A. Thus U ∈ Dif(A).AsU ∈ US, it follows from (4.2) that U ∈ B(δ).<br />

Let Un ∈ Z(δ), δS(Un) = λnUn, and <strong>le</strong>t U ∈ B(H) and U − Un →0. If λn →∞, then<br />

Un/λn → 0 and δS(Un/λn) = Un → U. Since δS is a closed <strong>de</strong>rivation, U = 0. This contradiction<br />

shows that {λn} is boun<strong>de</strong>d. Choose a subsequence converging to some λ and <strong>de</strong>note it also<br />

by {λn}. Then δS(Un) = λnUn → λU. Since δS is a closed <strong>de</strong>rivation, U ∈ US and δS(U) = λU.<br />

Hence Un converge to U in ·δS<br />

and, as above, U ∈ Dif(A). By (4.2), U ∈ Z(δ). Part (i) is<br />

proved.<br />

If A is an i<strong>de</strong>al of AS then, for each U in US, themapA → UAU ∗ preserves A. Hence<br />

Dif(A) ⊇ US ⊇ ZS and (ii) follows from (4.2). ✷


5. Structure of the group ZS<br />

E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 619<br />

Let U ∈ ZS and δS(U) = λU, forλ∈C. Then U ∗ ∈ US and δS(U ∗ ) = δS(U) ∗ = λU ∗ .As<br />

∗ ∗ ∗<br />

0 = δS(1) = δS U U = U δS(U) + δS U U = (λ + λ)U ∗ U = (λ + λ)1,<br />

we have Re(λ) = 0. For each t ∈ R,set<br />

ZS(t) = U ∈ ZS: δS(U) = itU <br />

and ΓS = t ∈ R: ZS(t) =∅ .<br />

Then ZS(t)ZS(s) ⊆ ZS(t + s), fort,s ∈ ΓS, soUZS(0) ⊆ ZS(t) and U ∗ ZS(t) ⊆ ZS(0), for<br />

U ∈ ZS(t). Hence<br />

ZS(t) = UZS(0) = ZS(0)U for each U ∈ ZS(t),<br />

ZS(−t)= ZS(t) ∗ , ZS(t + s) = ZS(t)ZS(s) and ZS = <br />

t∈ΓS<br />

ZS(t). (5.1)<br />

All sets ZS(t) are norm closed and ZS(0) is a selfadjoint normal subgroup of the group ZS; ΓS is<br />

a subgroup of R by addition, isomorphic to the quotient group ZS/ZS(0).<br />

Denote by Λ(S) and Λ(S ∗ ) the sets of all eigenvalues of operators S and S ∗ and by Hλ(S)<br />

and Hλ(S ∗ ) the corresponding eigenspaces of S and S ∗ . For a selfadjoint S, <strong>le</strong>tES(λ) be the<br />

spectral resolution of the i<strong>de</strong>ntity of S. Then<br />

ES−t1(λ) = ES(λ + t). (5.2)<br />

Let t ∈ R −{0}. We say that a unitary operator U on H is an (S,t)-shift if<br />

Theorem 5.1.<br />

UD(S)= D(S) and UES(λ)U ∗ = ES(λ + t) for all λ ∈ R.<br />

(i) If ZS(t) ={0} then the map λ → λ + t is an isomorphism of the sets Sp(S), Λ(S), Sp(S∗ ),<br />

Λ(S∗ ).ForU∈ ZS(t) and all λ ∈ Λ(S) and μ ∈ Λ(S∗ ),<br />

∗<br />

Hλ+t(S) = UHλ(S) and Hμ+t S ∗<br />

= UHμ S .<br />

(ii) If S is selfadjoint then U ∈ ZS(t), t = 0, if and only if U is an (S, t)-shift.<br />

Proof. We have UD(S)= D(S), U ∗ D(S) = D(S) and<br />

Hence<br />

(SU − US)|D(S) = tU|D(S) and SU ∗ − U ∗ S D(S) =−tU ∗ |D(S). (5.3)<br />

U ∗ S − (λ + t)1 U D(S) = (S − λ1)|D(S) for each λ ∈ C.<br />

Therefore λ ∈ Sp(S) if and only if λ + t ∈ Sp(S). Hence λ → λ + t is an isomorphism of Sp(S).


620 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

By (5.3), for x ∈ D(S) and y ∈ D(S ∗ ),<br />

(Sx, Uy) = U ∗ Sx,y = SU ∗ x,y + tU ∗ x,y = x, US ∗ + tU y .<br />

Hence Uy ∈ D(S ∗ ) and (S ∗ U −US ∗ )y = tUy. Simi<strong>la</strong>rly, U ∗ y ∈ D(S ∗ ) and (S ∗ U ∗ −U ∗ S ∗ )y =<br />

−tU ∗ y. Therefore UD(S ∗ ) = D(S ∗ ), U ∗ D(S ∗ ) = D(S ∗ ) and<br />

S ∗ U − US ∗ D(S ∗ ) = tU|D(S ∗ ) and S ∗ U ∗ − U ∗ S ∗ D(S ∗ ) =−tU ∗ |D(S ∗ ). (5.4)<br />

Hence, as above, we have that λ → λ + t is an isomorphism of Sp(S ∗ ).<br />

For λ ∈ Λ(S), we have from (5.3) that UHλ ⊆ Hλ+t and U ∗ Hλ ⊆ Hλ−t . Therefore<br />

UHλ ⊆ Hλ+t = UU ∗ Hλ+t ⊆ UHλ.<br />

Hence UHλ = Hλ+t and λ → λ + t is an isomorphism of Λ(S). Using (5.4), we obtain that the<br />

same is true for Λ(S ∗ ). Part (i) is proved.<br />

For any unitary U, the operator USU ∗ is selfadjoint,<br />

D USU ∗ = UD(S) and EUSU∗(λ) ∗<br />

= UES(λ)U<br />

for all λ ∈ R. (5.5)<br />

Let U ∈ ZS(t). By (5.3), UD(S) = U ∗ D(S) = D(S) and USU ∗ |D(S) = S − t1| D(S). Hence it<br />

follows from (5.2) that<br />

EUSU ∗(λ) = ES−t1(λ) = ES(λ + t) for all λ ∈ R.<br />

Taking into account (5.5), we have UES(λ)U ∗ = ES(λ + t). Thus U is an (S,t)-shift.<br />

Conversely, <strong>le</strong>t U be an (S,t)-shift. Then UD(S) = D(S) and UES(λ)U ∗ = ES(λ + t) for<br />

all λ ∈ R. Hence it follows from (5.2) and (5.5) that<br />

so USU ∗ |D(S) = S − t1|D(S). Thus<br />

Therefore U ∈ ZS(t). ✷<br />

EUSU ∗(λ) = UES(λ)U ∗ = ES−t1(λ),<br />

δS(U)|D(S) = i(SU − US)|D(S) = itU|D(S).<br />

Theorem 5.1 has an especially simp<strong>le</strong> form when S is diagonal, that is, H = <br />

λ∈Λ(S) Hλ.<br />

Corol<strong>la</strong>ry 5.2. Let S be a diagonal selfadjoint operator. Then t ∈ ΓS if and only if λ → λ + t is<br />

an isomorphism of Λ(S) and dim Hλ = dim Hλ+t for all λ ∈ Λ(S).<br />

From Corol<strong>la</strong>ry 5.2 it follows that, for any subgroup Γ of R, there is a diagonal S with<br />

ΓS = Γ .<br />

If for each t ∈ ΓS, there is Ut ∈ ZS(t) such that U ={Ut: t ∈ ΓS} is a group, then U is cal<strong>le</strong>d<br />

a resolving subgroup of ZS. It is commutative and consists of unitary operators Ut , t ∈ ΓS,<br />

satisfying<br />

UtD(S) = D(S) and (SUt − UtS)|D(S) = tUt|D(S).


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 621<br />

This re<strong>la</strong>tion is cal<strong>le</strong>d the infinitesimal Weyl re<strong>la</strong>tion for the group U and the operator S (see [4]).<br />

It follows from (5.1) that ZS is the <strong>semi</strong>-direct pro<strong>du</strong>ct of U and the normal subgroup ZS(0).<br />

Proposition 5.3. If ΓS has a minimal positive e<strong>le</strong>ment μ, then ΓS ={nμ: n ∈ Z} and, for each<br />

U ∈ ZS(μ), U ={U n : n ∈ Z} is a resolving subgroup of ZS.<br />

Proof. We only need to show that ΓS ={nμ: n ∈ Z}. If there is λ ∈ ΓS such that λ = nμ, for<br />

all n ∈ Z, then mμ


622 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

It was proved in [7] that AS = CS + (AS ∩ C(H)) if all dim(Hi)0.<br />

For λ ∈ Λ(S), <strong>le</strong>tPλbe the projection on the eigenspace Hλ of S. Ifλ= μ, PλPμ = 0. Let<br />

A ∈ C(H). Each sequence {xλn }, xλn ∈ Hλn with xλn=1 and distinct λn, weakly converges<br />

to 0. Hence Axλn→0. This implies that ΛA ={λ∈ Λ(S): APλ <br />

= 0} is a finite or countab<strong>le</strong><br />

set: ΛA ={λi}, and APλi →0asi →∞. Thus the series λi∈ΛA<br />

Pλi APλi converges in norm,<br />

so<br />

ρ : A → <br />

PλAPλ = <br />

(5.8)<br />

λ∈Λ(S)<br />

λi∈ΛA<br />

Pλi APλi<br />

is a map from C(H) into C(H) and ρ=1. Let Q = 1 − <br />

λ∈Λ(S) Pλ and set<br />

DS = A ∈ C(H): AQ = QA = 0 and PλA = APλ for all λ ∈ Λ(S) .<br />

Then DS is a C*-subalgebra of C(H). For each A ∈ C(H), ρ(A) ∈ DS and, for each A ∈ DS,<br />

Lemma 5.5.<br />

<br />

A = Q + <br />

(i) CS is a W ∗ -algebra and<br />

λ∈Λ(S)<br />

Pλ<br />

<br />

A = ρ(A), so DS = ρ(A): A ∈ C(H) .<br />

CS = A ∈ B(H): AD(S) ⊆ D(S), A ∗ D(S) ⊆ D(S), [S,A]|D(S) = 0 <br />

= A ∈ B(H): AD(S) ⊆ D(S), AD S ∗ ⊆ D S ∗ , [S,A]|D(S) = S ∗ ,A <br />

D(S∗ = 0 ) .<br />

(ii) All Pλ ∈ CS ∩ C ′ S ,forλ ∈ Λ(S), and DS = CS ∩ C(H).<br />

Proof. The first equality in (i) follows from (1.2) and (5.6). For A ∈ CS, A ∗ ∈ AS and δS(A ∗ ) =<br />

δS(A) ∗ = 0. Thus A ∗ ∈ CS, soCS is a *-algebra. Denote by Π the <strong>la</strong>st set in (i). Let A ∈ CS. For<br />

all x ∈ D(S ∗ ) and y ∈ D(S),<br />

(Ax, Sy) = x,A ∗ Sy = x,SA ∗ y = AS ∗ x,y .<br />

Hence AD(S∗ ) ⊆ D(S∗ ) and AS∗ |D(S∗ ) = S∗A|D(S∗ ),so[S∗ ,A]|D(S∗ ) = 0. Thus CS ⊆ Π.<br />

Conversely, <strong>le</strong>t A ∈ Π. Then, for all x ∈ D(S∗ ) and y ∈ D(S),<br />

∗ ∗ ∗ ∗ ∗<br />

S x,A y = AS x,y = S Ax, y = x,A Sy .<br />

Hence A ∗ y ∈ D(S ∗∗ ) = D(S), soA ∗ D(S) ⊆ D(S). Thus Π ⊆ CS, soΠ = CS.


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 623<br />

Let CS ∋ Aλ → A in the weak operator topology (wot). Then A∗ wot<br />

λ → A∗ .AsS∗∗ = S,wehave,<br />

for all x ∈ D(S∗ ) and y ∈ D(S),<br />

∗ ∗ ∗ ∗<br />

S x,Ay = A S x,y = limλ AλS ∗ x,y ∗ ∗<br />

= lim S Aλx,y λ<br />

<br />

= lim λ<br />

A ∗ λ x,Sy = A ∗ x,Sy = (x, ASy).<br />

Hence Ay ∈ D(S ∗∗ ) = D(S) and SAy = ASy. Thus AD(S) ⊆ D(S) and [S,A]|D(S) = 0. Simi<strong>la</strong>rly,<br />

A ∗ D(S) ⊆ D(S). Therefore A ∈ CS, soCS is a W*-algebra. Part (i) is proved.<br />

Since PλD(S) ⊆ Hλ ⊆ D(S) and<br />

PλS|D(S) = SPλ|D(S) = λPλ|D(S), (5.9)<br />

we have δS(Pλ) = 0, so Pλ ∈ CS. LetA ∈ CS. Forx ∈ Hλ, wehaveSAx = ASx = λAx, so<br />

Ax ∈ Hλ. Hence PλAPλ = APλ. Since CS is a *-algebra, PλA = APλ, soPλ ∈ CS ∩ C ′ S .<br />

Let A ∈ C(H). For each λ ∈ Λ(S), PλAPλ ∈ C(H) and PλAPλD(S) ⊆ Hλ ⊆ D(S). By (5.9),<br />

δS(PλAPλ)|D(S) = i(SPλAPλ|D(S) − PλAPλS|D(S)) = 0.<br />

Hence PλAPλ ∈ CS ∩ C(H).Asρ(A) is the norm limit of sums of the operators PλAPλ, λ ∈ ΛA,<br />

we have ρ(A) ∈ CS ∩ C(H). Thus DS ⊆ CS ∩ C(H).<br />

Conversely, <strong>le</strong>t A = A ∗ ∈ CS ∩ C(H). Then A = <br />

i αiPi, where Pi are finite-dimensional<br />

mutually orthogonal projections from CS ∩ C(H) and |αi|→0. Each subspace PiH lies in D(S)<br />

and the operator S|PiH is selfadjoint. Hence PiH = <br />

j Kij , where S|Kij = λij PKij . Therefore<br />

λij ∈ Λ(S). AsPλ∈ C ′ S , each Pi commutes with all Pλ, soPKij = Pλij PiPλij ∈ DS. Hence<br />

Pi = <br />

j PKij ∈ DS.AsDS is norm closed, A ∈ DS. Thus DS = CS ∩ C(H). ✷<br />

Recall that a closed subspace L of H (the projection Q on L) re<strong>du</strong>ces a symmetric operator<br />

S if<br />

QD(S) ⊆ D(S) and SQ|D(S) = QS|D(S). (5.10)<br />

The operator S is cal<strong>le</strong>d simp<strong>le</strong> if it has no re<strong>du</strong>cing subspaces where it in<strong>du</strong>ces a selfadjoint<br />

operator; it is cal<strong>le</strong>d irre<strong>du</strong>cib<strong>le</strong> if it has no re<strong>du</strong>cing subspaces.<br />

Denote by H (n) ,1 n ∞, the orthogonal sum of n copies of H and by S (n) the orthogonal<br />

sum of n copies of S. Lemma 5.5(i) and (5.10) yield<br />

Lemma 5.6.<br />

(i) A projection Q re<strong>du</strong>ces a symmetric operator S if and only Q ∈ CS.<br />

(ii) S is irre<strong>du</strong>cib<strong>le</strong> if and only if CS = C1.<br />

(iii) If S is irre<strong>du</strong>cib<strong>le</strong>, C S (n) consists of all block-matrix boun<strong>de</strong>d operators (λij 1H ) on H (n)<br />

with λij ∈ C.<br />

Let S ∗ be the adjoint of S. The <strong>de</strong>ficiency subspaces N±(S) ={x ∈ D(S ∗ ): S ∗ x =±ix} of S<br />

areclosedinH and n±(S) = dim N±(S) are cal<strong>le</strong>d the <strong>de</strong>ficiency indices of S. The operator S<br />

is selfadjoint if n−(S) = n+(S) = 0; it is maximal symmetric if either n−(S) = 0orn+(S) = 0.


624 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

Recall that symmetric operators R on K and S on H are isomorphic if<br />

UD(R)= D(S) and UR|D(R) = SU|D(R), (5.11)<br />

for some unitary operator U from K on H .IfS and R are isomorphic, ZS = ZR and ΓS = ΓR.<br />

The operator T = i d dt on H = L2(0, ∞) with<br />

D(T ) = h ∈ H : h are absolutely continuous, h ′ ∈ H and h(0) = 0 <br />

is simp<strong>le</strong> and maximal symmetric with n−(T ) = 1, n+(T ) = 0. For each r ∈ R, the multiplication<br />

operator<br />

Vrh(t) = e −irt h(t) (5.12)<br />

on H is unitary. It is easy to check that Vr ∈ UT and δT (Vr) = irVr for all r ∈ R,soVr ∈ ZT (r).<br />

It is well known (see [1]) that each simp<strong>le</strong> maximal symmetric operator S is isomorphic<br />

either to T (k)<br />

if n−(S) = k, n+(S) = 0; or to − T (k)<br />

if n−(S) = 0, n+(S) = k. (5.13)<br />

Using Lemma 5.6, we have the following <strong>de</strong>scription of ZS for simp<strong>le</strong> maximal symmetric operators.<br />

Theorem 5.7. Let S be a simp<strong>le</strong> maximal symmetric operator satisfying (5.13), forsomek, and<br />

<strong>le</strong>t Vr, r ∈ R, be the unitary operators <strong>de</strong>fined in (5.12). Then ΓS = R, {V(r) (k) : r ∈ R} is a<br />

resolving subgroup of ZS and ZS(0) consists of all unitary block-matrix operators (λij 1H ) on<br />

H (k) with λij ∈ C.<br />

We consi<strong>de</strong>r now the following criteria for a symmetric operator to be irre<strong>du</strong>cib<strong>le</strong>.<br />

Lemma 5.8. Let S be a symmetric operator. Let {λn} be eigenvalues of S ∗ with one-dimensional<br />

eigenspaces: Hn = Chn and <strong>le</strong>t the linear span of all hn be <strong>de</strong>nse in H . Suppose that all<br />

hn /∈ D(S). If there are μn ∈ C such that hn − μnh1 ∈ D(S), for all n, then S is irre<strong>du</strong>cib<strong>le</strong>.<br />

Proof. Let a projection Q belong to CS. It commutes with S ∗ ,soS ∗ Qhn = QS ∗ hn = λnQhn.<br />

Since Hn are one-dimensional, Qhn = αnhn, where αn = 0 or 1. Since Q preserves D(S),<br />

Q(hn − μnh1) = αnhn − μnα1h1 = αn(hn − μnh1) + (αn − α1)μnh1 ∈ D(S).<br />

Hence (αn −α1)μnh1 ∈ D(S) for all n. Since all μn = 0, all αn = α1. Thus Q is either 1 or 0. ✷<br />

We shall now consi<strong>de</strong>r an irre<strong>du</strong>cib<strong>le</strong> non-maximal symmetric operator with a resolving subgroup.<br />

The symmetric operator S = i d dt on H = L2(0, 2π) with<br />

D(S) = h ∈ H : h is absolutely continuous, h ′ ∈ H and h(0) = h(2π)= 0 <br />

has n−(S) = n+(S) = 1 (see [1]). It is irre<strong>du</strong>cib<strong>le</strong>. In<strong>de</strong>ed, S ∗ = i d dt and<br />

D(S ∗ ) ={h ∈ H : h is absolutely continuous and h ′ ∈ H }.


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 625<br />

The functions hn(t) = e int , −∞


626 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

(i) δS|A is the maximal closed ∗ -<strong>de</strong>rivation of A imp<strong>le</strong>mented by S, that is, (6.1) holds;<br />

(ii) there exists a boun<strong>de</strong>d linear map θ from C(H) onto A ∩ C(H) such that<br />

and θ commutes with δ max<br />

S :<br />

θ(A)= A for all A ∈ A ∩ C(H), (6.2)<br />

θ(A)∈ JS and δ max<br />

max<br />

S θ(A) = θ δS (A) <br />

for all A ∈ JS. (6.3)<br />

Then B = A+JS is a domain of A+C(H), δS|B ∈ Der(B) and S is its minimal imp<strong>le</strong>mentation.<br />

Proof. Set δ = δS|B. Since S imp<strong>le</strong>ments δ and FS ⊆ JS ⊆ B, it is easy to see that S is a minimal<br />

imp<strong>le</strong>mentation of δ. Thus we only need to prove that δ is closed.<br />

By (6.2), C(H) = (A ∩ C(H))∔ Ker(θ) and Ker(θ) is a closed subspace of C(H). Therefore<br />

A + C(H) = A ∔ Ker(θ) (6.4)<br />

is the direct sum of A and Ker(θ).LetA ∈ JS. Since θ(A)∈ A ∩ C(H), we have from (6.3) that<br />

θ(A)∈ A ∩ JS and δS<br />

max<br />

max<br />

θ(A) = δS θ(A) = θ δS (A) ∈ A ∩ C(H). (6.5)<br />

Hence θ(A) ⊕ δS(θ(A)) belongs to A ⊕ A and to G(δS). Since δS|A satisfies (6.1), we have<br />

θ(A)∈ A. Therefore A = θ(A)+ (A − θ(A)) and A − θ(A)∈ JS ∩ Ker(θ). Thus, by (6.4),<br />

B = A + JS = A ∔ JS ∩ Ker(θ) .<br />

Let An ∈ A, Bn ∈ JS ∩ Ker(θ), A,T ∈ A and B,R ∈ Ker(θ), <strong>le</strong>tAn + Bn → A + B ∈<br />

A + C(H) and <strong>le</strong>t δ(An + Bn) → T + R. By (6.5), θ(δmax S (Bn)) = δmax S (θ(Bn)) = 0. Hence<br />

δmax S (Bn) ∈ Ker(θ). Thus<br />

δ(An + Bn) = δS(An) + δ max<br />

S (Bn) → T + R, where δS(An) ∈ A and δ max<br />

S (Bn) ∈ Ker(θ).<br />

If a sequence in the direct sum of closed subspaces converges, the components of its e<strong>le</strong>ments<br />

also converge. Hence, by (6.4), An → A, Bn → B, δS(An) → T and δmax S (Bn) → R. As<br />

δS|A is a closed <strong>de</strong>rivation, we have A ∈ A and δS(A) = T .Asδmax S is a closed <strong>de</strong>rivation,<br />

B ∈ JS ∩ Ker(θ) and δmax S (B) = R. Thus A + B ∈ B and δ(A + B) = T + R, soδ is a closed<br />

*-<strong>de</strong>rivation. ✷<br />

Corol<strong>la</strong>ry 6.2. Let A be a domain of A ⊆ B(H) and δS|A ∈ Der(A) satisfy (6.1). If<br />

A ∩ C(H) ={0}, then B = A + JS is a domain of A + C(H), δS|B ∈ Der(B) and S is its<br />

minimal imp<strong>le</strong>mentation.<br />

Let A be a C*-subalgebra of CS. Then δS|A = 0 and it satisfies (6.1). For λ ∈ Λ(S), <strong>le</strong>tPλbe the projection on the eigenspace Hλ of S. By Lemma 5.5(ii), PλCSPλ ⊆ CS. Assume also that<br />

<br />

Pλ A ∩ C(H) Pλ ⊂ A for all λ ∈ Λ(S). (6.6)


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 627<br />

Then all Pλ(A ∩ C(H ))Pλ are C*-algebras. Since δS(A) = 0, for A ∈ CS ∩ C(H),<br />

CS ∩ C(H) ⊆ CS ∩ JS. (6.7)<br />

Let A ∈ C(H). Recall (see (5.8)) that APλ = 0 for a finite or countab<strong>le</strong> subset ΛA ={λi} of<br />

Λ(S),<br />

APλi →0 and ρ(A) =<br />

<br />

PλAPλ =<br />

λ∈Λ(S)<br />

<br />

Pλi<br />

λi∈ΛA<br />

APλi ∈ CS ∩ C(H) ⊆ CS ∩ JS,<br />

where the series converges in norm.<br />

We will now construct a boun<strong>de</strong>d linear map θ from C(H) onto A∩C(H) satisfying (6.2) and<br />

(6.3). We will use for this the well-known result that, for each C ∗ -subalgebra B of C(H), there is<br />

a conditional expectation from C(H) onto B. Thus, for each λ ∈ Λ(S), there is a conditional expectation<br />

θλ from the algebra PλC(H)Pλ which is isomorphic to C(Hλ) onto the C*-subalgebra<br />

Pλ(A ∩ C(H ))Pλ of PλC(H)Pλ. Set<br />

θ(A)= <br />

λ∈Λ(S)<br />

θλ(PλAPλ) = <br />

λi∈ΛA<br />

θλi (Pλi APλi ) for all A ∈ C(H). (6.8)<br />

Since θλi (Pλi APλi ) APλi →0 and since θλi (Pλi APλi ) belong to Pλi C(H)Pλi and,<br />

hence, mutually orthogonal, the series in (6.8) is norm convergent. Hence we have from (6.7)<br />

that<br />

θ(A)∈ A ∩ C(H) ⊆ CS ∩ C(H) ⊆ CS ∩ JS for all A ∈ C(H), (6.9)<br />

so θ maps C(H) into A ∩ C(H). Moreover, θ is linear and boun<strong>de</strong>d, since<br />

<br />

θ(A) = supθλi (Pλi APλi ) sup Pλi APλi A.<br />

Since projections Pλ, λ ∈ Λ(S), are mutually orthogonal, Pλρ(A)Pλ = PλAPλ (see (5.8)).<br />

Hence<br />

θ ρ(A) = <br />

λ∈Λ(S)<br />

<br />

θλ Pλρ(A)Pλ =<br />

λi∈ΛA<br />

θλi (Pλi APλi ) = θ(A).<br />

Since θλ(PλAPλ) ∈ Pλ(A ∩ C(H ))Pλ,wehavePλθ(A)Pλ = θλ(PλAPλ), so (see (5.8))<br />

ρ θ(A) = <br />

Pλθ(A)Pλ = <br />

θλi (Pλi APλi ) = θ(A).<br />

Thus<br />

λ∈Λ(S)<br />

λi∈ΛA<br />

θ ρ(A) = ρ θ(A) = θ(A) for all A ∈ C(H). (6.10)


628 E. Kissin / Journal of Functional Analysis 236 (2006) 609–629<br />

Let A ∈ A ∩ C(H). Then A ∈ CS ∩ C(H) and we have from Lemma 5.5(ii) that A = ρ(B)<br />

for some B ∈ C(H). Since PλAPλ ∈ Pλ(A ∩ C(H ))Pλ and θλ are conditional expectations,<br />

θλ(PλAPλ) = PλAPλ = Pλρ(B)Pλ = PλBPλ. Thus (6.2) holds, since<br />

θ(A)= <br />

θλ(PλAPλ) = <br />

PλBPλ = ρ(A) = A.<br />

λ∈Λ(S)<br />

λ∈Λ(S)<br />

Let now A ∈ JS. Since PλH ⊂ D(S), for all λ ∈ Λ(S), we have from (5.9)<br />

Hence<br />

Pλδ max<br />

S (A)Pλ = PλδS(A)Pλ = Pλi[S,A]Pλ = i(PλSAPλ − PλASPλ) = 0.<br />

ρ δ max<br />

S (A) = <br />

λ∈Λ(S)<br />

As δ max<br />

S (A) ∈ C(H), we have from (6.10) that<br />

Pλδ max<br />

S (A)Pλ = 0.<br />

θ δ max<br />

S (A) = θ ρ δ max<br />

S (A) = 0.<br />

By (6.9), θ(C(H))⊆ CS ∩ JS, so that δ max<br />

S (θ(A)) = 0. Thus<br />

δ max<br />

max<br />

S θ(A) = 0 = θ δS (A) <br />

Therefore (6.3) holds and Theorem 6.1 yields<br />

for all A ∈ JS.<br />

Theorem 6.3. Let a C ∗ -subalgebra A of CS satisfy (6.6). Then B = A + JS is a domain of the<br />

C ∗ -algebra A + C(H), δS|B is a closed ∗ -<strong>de</strong>rivation of A + C(H) with minimal imp<strong>le</strong>mentation<br />

S.<br />

Finally, we consi<strong>de</strong>r <strong>de</strong>rivations δ with Z(δ) = ZS, where S is a minimal imp<strong>le</strong>mentation of δ.<br />

Proposition 6.4.<br />

(i) Let T be a minimal symmetric imp<strong>le</strong>mentation of δ ∈ Der(FS). Then B(δ) = UT and<br />

Z(δ) = ZT .<br />

(ii) B(δmax S ) = US and Z(δmax S ) = ZS.<br />

(iii) Let δ = δS|(CS + JS). Then Z(δ) = ZS.<br />

Proof. As δmin S ,δ∈ Der(FS), it follows from Corol<strong>la</strong>ry 3.2 that<br />

FS = F FS,δ min<br />

S = F(FS,δ)= FT .<br />

Since FT is an i<strong>de</strong>al of AT , Proposition 4.2(ii) yields (i).<br />

As JS is an i<strong>de</strong>al of AS, Proposition 4.2(ii) also yields (ii).<br />

Let U ∈ ZS(t) and A ∈ CS. Since ZS(t) ⊆ AS and CS ⊆ AS,wehaveUAU ∗ ∈ AS. Further<br />

∗<br />

δS UAU = δS(U)AU ∗ + UδS(A)U ∗ ∗<br />

+ UAδS U = itUAU ∗ + UA −itU ∗ = 0.


E. Kissin / Journal of Functional Analysis 236 (2006) 609–629 629<br />

Hence UAU ∗ ∈ CS, soUCSU ∗ = CS.AsJS is an i<strong>de</strong>al of AS,<br />

UD(δ)U ∗ = U(CS + JS)U ∗ ⊆ D(δ).<br />

Thus U ∈ Z(δ), soZS ⊆ Z(δ). Since always Z(δ) ⊆ ZS, wehaveZ(δ) = ZS. ✷<br />

Let S be a selfadjoint operator on H = ∞ i=−∞ Hi and <strong>le</strong>t S|Hi = λi1Hi with all distinct λi.<br />

The group ΓS is <strong>de</strong>scribed in Corol<strong>la</strong>ry 5.2 and CS consists of boun<strong>de</strong>d operators commuting<br />

with all Pλi . By Theorem 6.3 and Proposition 6.4, δ = δS|(CS + JS) is a closed *-<strong>de</strong>rivation of<br />

the C*-algebra CS + C(H) and Z(δ) = ZS.<br />

Acknow<strong>le</strong>dgment<br />

We are very grateful to Victor Shulman for his valuab<strong>le</strong> suggestions about this paper.<br />

References<br />

[1] N.I. Ahiezer, I.M. G<strong>la</strong>zman, The Theory of Linear Operators in Hilbert Spaces, Ungar, New York, 1961.<br />

[2] O. Bratteli, D.W. Robinson, Unboun<strong>de</strong>d <strong>de</strong>rivations of C*-algebras, Comm. Math. Phys. 42 (1975) 253–268.<br />

[3] J. Dixmier, Les C*-algebras et <strong>le</strong>urs representations, Gauthier–Vil<strong>la</strong>rs, Paris, 1969.<br />

[4] P.E.T. Jorgensen, P.S. Muhly, Selfadjoint extensions satisfying the Weyl operator commutation re<strong>la</strong>tions, J. Anal.<br />

Math. 37 (1980) 46–99.<br />

[5] E. Kissin, V.S. Shulman, Representations on Krein Spaces and Derivations of C*-algebras, Addison–<br />

Wes<strong>le</strong>y/Longman, Harlow, 1997.<br />

[6] E. Kissin, V.S. Shulman, Differential Banach *-algebras of compact operators associated with symmetric operators,<br />

J. Funct. Anal. 156 (1998) 1–29.<br />

[7] E. Kissin, V.S. Shulman, Dual spaces and isomorphisms of some differential Banach *-algebras of operators, Pacific<br />

J. Math. 190 (1999) 329–360.<br />

[8] M.A. Naimark, Normed Rings, Nauka, Moscow, 1968.<br />

[9] S. Sakai, Operator Algebras in Dynamical Systems, Cambridge Univ. Press, Cambridge, 1991.


Journal of Functional Analysis 236 (2006) 630–633<br />

www.elsevier.com/locate/jfa<br />

AnewprimeC ∗ -algebra that is not primitive<br />

M.J. Crabb<br />

Department of Mathematics, University of G<strong>la</strong>sgow, G<strong>la</strong>sgow G12 8QW, Scot<strong>la</strong>nd, UK<br />

Received 9 January 2006; accepted 28 February 2006<br />

Communicated by G. Pisier<br />

Abstract<br />

An explicit prime non-primitive C∗-algebra is constructed.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: C ∗ -algebra; Primitivity<br />

In [7], N. Weaver gives an examp<strong>le</strong> of a C ∗ -algebra that is prime but not primitive, solving a<br />

long-standing prob<strong>le</strong>m. Here we give a simp<strong>le</strong>r examp<strong>le</strong>.<br />

An algebra is said to be prime if the pro<strong>du</strong>ct of any two nonzero two-si<strong>de</strong>d i<strong>de</strong>als is nonzero.<br />

An algebra is primitive if it has a faithful irre<strong>du</strong>cib<strong>le</strong> representation. In the case of a C ∗ -algebra<br />

this representation can be assumed to be a ∗-representation on a Hilbert space [6, Corol<strong>la</strong>ry<br />

2.9.6]. A primitive algebra is always prime. J. Dixmier [5] proved that a separab<strong>le</strong> prime<br />

C ∗ -algebra is primitive.<br />

Throughout, X is an uncountab<strong>le</strong> set and G the free group on X with i<strong>de</strong>ntity e. Forw in G<br />

in re<strong>du</strong>ced form, <strong>de</strong>fine w! to be the set of all prefixes of w, including e and w. For examp<strong>le</strong>,<br />

(xyz)! ={e,x,xy,xyz}. Denote by P the set of all e<strong>le</strong>ments of G that in re<strong>du</strong>ced form have<br />

no negative exponents, and put Q = P −1 , the set of e<strong>le</strong>ments with no positive exponents. Thus<br />

P ∩ Q ={e}. Define L by<br />

L := {p!∪q!: p ∈ P, q ∈ Q}.<br />

For a = x1x2 ···xn in P , with xi ∈ X, we say that a has <strong>de</strong>gree n and a −1 has <strong>de</strong>gree −n; we<br />

<strong>de</strong>fine the content of a, con(a) := {x1,x2,...,xn}=con(a −1 ). We also <strong>de</strong>fine the <strong>de</strong>gree of e to<br />

E-mail address: mjc@maths.g<strong>la</strong>.ac.uk.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.02.018


M.J. Crabb / Journal of Functional Analysis 236 (2006) 630–633 631<br />

be 0 and take con(e) := ∅.ForA⊆P ∪ Q, <strong>de</strong>fine con(A) := <br />

g∈A con(g). The cardinal of a set<br />

S will be <strong>de</strong>noted by |S| and the symbol ⊃ will <strong>de</strong>note strict containment.<br />

Lemma.<br />

(L1) A ∈ L, g ∈ A ⇒ g −1 A ∈ L.<br />

(L2) A,B,C ∈ L, g ∈ A, h ∈ B, A ∪ gB ⊆ ghC ⇒ A ∪ gB ∈ L.<br />

(L3) For A ∈ L, A has at most one e<strong>le</strong>ment of each <strong>de</strong>gree.<br />

(L4) A ∈ L, h ∈ P ∪ Q, hA ⊆ A ⇒ h = e.<br />

Proof. (L1) Suppose that A = (a −1 )!∪b!, where a,b ∈ P . Then we have that A = a −1 (ab)!=<br />

a −1 c!, where c = ab. Ifg ∈ A then g = a −1 d for some d ∈ c!, sayc = df with f ∈ P . Then<br />

g −1 A = d −1 c!=d −1 (df )!=(d −1 )!∪f !∈L.<br />

(L2) We first show that<br />

A1,A2,A3 ∈ L, A1 ∪ A2 ⊆ A3 ⇒ A1 ∪ A2 ∈ L. (1)<br />

Suppose that Ai = pi!∪qi! (pi ∈ P , qi ∈ Q, i = 1, 2, 3). Then p1,p2 ∈ p3! and q1,q2 ∈ q3!.<br />

C<strong>le</strong>arly, A1 ∪ A2 = p!∪q!, where p is p1 or p2 and q is q1 or q2.<br />

Now assume that A,B,C ∈ L, g ∈ A, h ∈ B and A ∪ gB ⊆ ghC. Then g −1 A ∪ B ⊆ hC.<br />

By (L1), g −1 A ∈ L and since e ∈ hC, alsoh −1 ∈ C, hC ∈ L. By(1),g −1 A ∪ B ∈ L. Since<br />

g −1 ∈ g −1 A ∪ B, (L1) shows that A ∪ gB = g(g −1 A ∪ B) ∈ L.<br />

(L3), (L4) C<strong>le</strong>ar. ✷<br />

Define M := {(A, g): A ∈ L, g∈ A} and H := l 2 (M), a Hilbert space with inner pro<strong>du</strong>ct 〈|〉.<br />

We i<strong>de</strong>ntify e<strong>le</strong>ments of M with basis unit vectors of l 2 (M). For(A, g) ∈ M, <strong>de</strong>fine a linear<br />

operator Ag on H by the ru<strong>le</strong> that, for (C, k) ∈ M,<br />

<br />

(gC,gk) if A ⊆ gC,<br />

Ag(C, k) =<br />

0 otherwise.<br />

If A ⊆ gC here then e ∈ gC,g−1 ∈ C and gC ∈ L, by(L1),whichgives(gC,gk) ∈ M. Itis<br />

easily verified that Ag is a partial isometry. Write M := {Ag: (A, g) ∈ M}.IfalsoBh∈M then<br />

for (C, k) ∈ M we have that<br />

<br />

(ghC, ghk) if A ∪ gB ⊆ ghC,<br />

AgBh(C, k) =<br />

(2)<br />

0 otherwise.<br />

If here A ∪ gB ⊆ ghC then, by (L2), A ∪ gB ∈ L and (A ∪ gB)gh ∈ M. IfA ∪ gB ∈ L then<br />

(2) gives AgBh = (A ∪ gB)gh. IfA ∪ gB /∈ L then always A ∪ gB ghC and AgBh = 0. For<br />

Ag ∈ M,also(g −1 A) g −1 ∈ M,by(L1).For(C, k) and (D, f ) in M,<br />

Ag(C, k) = (D, f ) ⇔ A ⊆ gC = D, gk = f<br />

⇔ g −1 A ⊆ g −1 D = C, g −1 f = k<br />

⇔ g −1 A <br />

g −1(D, f ) = (C, k).


632 M.J. Crabb / Journal of Functional Analysis 236 (2006) 630–633<br />

Since these are partial isometries it follows that (Ag) ∗ = (g −1 A) g −1. In particu<strong>la</strong>r, for A ∈ L,<br />

Ae is a projection. The above gives, for A,B ∈ L,<br />

<br />

(A ∪ B)e if A ∪ B ∈ L,<br />

AeBe =<br />

0 otherwise.<br />

It follows that CM ≡ lin(M) is a ∗-algebra of boun<strong>de</strong>d linear operators on H . Denote by C ∗ M<br />

its clo<strong>sur</strong>e, a C ∗ -algebra with i<strong>de</strong>ntity {e}e.<br />

Theorem. C ∗ M is prime but not primitive.<br />

Proof. First consi<strong>de</strong>r primeness. Let J and K be nonzero two-si<strong>de</strong>d i<strong>de</strong>als of C ∗ M.Wefirst<br />

prove that J contains a projection of M. There exists T ∈ J with T 0 and v = (A, g) ∈ M<br />

with 〈Tv|v〉 > 0. By rep<strong>la</strong>cing T by a sca<strong>la</strong>r multip<strong>le</strong> we can assume that 〈Tv|v〉=1. We have<br />

that Aev = Ae(A, g) = v. There is a sequence (Tn) in CM such that Tn → T . Put S := AeTAe ∈<br />

J and Sn := AeTnAe; then Sn → S. ForBh ∈ M, AeBhAe is 0 or (A ∪ B ∪ hA)h = Ch (say),<br />

where C ⊇ A and C = A only if hA ⊆ A and so h = e by (L4). We may therefore write Sn =<br />

αnAe + Rn, where αn ∈ C and Rn is a linear combination of e<strong>le</strong>ments Ch ∈ M with C ⊃ A.For<br />

each such Ch, we have that |C| > |hA|=|A|, C hA and so Chv = Ch(A, g) = 0. Therefore<br />

Rnv = 0 and Snv = αnAev = αnv. Hence αn =〈αnv|v〉=〈Snv|v〉→〈Sv|v〉=〈AeTAev|v〉=<br />

〈TAev|Aev〉=〈Tv|v〉=1asn →∞.<br />

Suppose that A = p!∪q!, where p ∈ P , q ∈ Q. Choose z in X not in con(C) for any Ch which<br />

appears in the linear sum expressions for the Rn. Define J ∈ L by J = (pz)!∪(qz −1 )!. IfCh<br />

appears in the sum for Rn then C ⊃ A and so C contains an e<strong>le</strong>ment of the <strong>de</strong>gree of pz or qz −1<br />

but which cannot equal pz or qz −1 .By(L3),J ∪ C/∈ L and so JeCh = 0. Hence JeRn = 0 and<br />

JeSn = αnJeAe = αnJe → Je, since J ∪ A = J .Also,JeSn → JeS. Therefore Je = JeS ∈ J .<br />

For j in J , (j −1 J)e = (j −1 J) j −1JeJj ∈ J .Takej = qz −1 which gives j −1 J ⊂ P .The<br />

i<strong>de</strong>al K contains a projection Ke of M. We likewise choose k ∈ K such that k −1 K ⊂ Q. Then<br />

(k −1 K)e ∈ K and JKcontains (j −1 J)e(k −1 K)e = (j −1 J ∪k −1 K)e = 0. This shows that C ∗ M<br />

is prime.<br />

Now consi<strong>de</strong>r primitivity. For m, n = 0, 1, 2, 3,...,<br />

(p!∪q!)e: p ∈ P, q ∈ Q, p has <strong>de</strong>gree m, q has <strong>de</strong>gree −n <br />

consists of orthogonal projections ((L3) and (3)). Let φ be a positive linear functional on C ∗ M.<br />

Then φ is positive only on a countab<strong>le</strong> subset of the above set. Therefore φ is positive only on a<br />

countab<strong>le</strong> set of Ae, A in L. Choose x in X not in con(A) for any A with φ(Ae)>0. If B ∈ L<br />

has x ∈ con(B) then φ(Be) = 0. If also h ∈ B then, since Bh = BeBh, the Cauchy–Schwarz<br />

inequality gives φ(Bh) = 0. Hence φ vanishes on lin{Bh: x ∈ con(B)}, an i<strong>de</strong>al of CM. Since<br />

φ is continuous [2, Section 37, Corol<strong>la</strong>ry 9], φ vanishes on its clo<strong>sur</strong>e, an i<strong>de</strong>al of C ∗ M.<br />

Suppose that C ∗ M has an irre<strong>du</strong>cib<strong>le</strong> ∗-representation on a comp<strong>le</strong>x space V with inner<br />

pro<strong>du</strong>ct 〈|〉.Letv ∈ V \ 0 and <strong>de</strong>fine a positive linear functional φ on C ∗ M by φ(a) =〈av|v〉.<br />

From above, φ vanishes on a nonzero i<strong>de</strong>al I. Letb ∈ I \ 0. For all a,c ∈ C ∗ M, 〈bcv|av〉=<br />

〈a ∗ bcv|v〉 =φ(a ∗ bc) = 0. Since v is cyclic this gives bV = 0, and the representation is not<br />

faithful. Therefore C ∗ M has no faithful irre<strong>du</strong>cib<strong>le</strong> representation, i.e. is not primitive. ✷<br />

(3)


M.J. Crabb / Journal of Functional Analysis 236 (2006) 630–633 633<br />

Remark. The set M ∪{0} forms an inverse monoid un<strong>de</strong>r the multiplication given by<br />

<br />

(A ∪ gB)gh if A ∪ gB ∈ L,<br />

AgBh =<br />

0 otherwise.<br />

This is cal<strong>le</strong>d the McAlister monoid on X and is the Rees quotient of the free inverse monoid on<br />

X by the i<strong>de</strong>al generated by all e<strong>le</strong>ments xy −1 and x −1 y with x,y ∈ X and x = y [4]. Barnes [1]<br />

constructs for any inverse <strong>semi</strong>group a representation on Hilbert space; this is used here to<br />

generate C ∗ M. Weaver’s examp<strong>le</strong> in [7] is also generated by an inverse <strong>semi</strong>group of partial<br />

isometries. Re<strong>la</strong>ted results on l 1 -algebras are <strong>de</strong>scribed in [3,4].<br />

Acknow<strong>le</strong>dgment<br />

I thank Professor W.D. Munn for intro<strong>du</strong>cing me to the notion of the McAlister monoid and<br />

for help with setting out this paper.<br />

References<br />

[1] B.A. Barnes, Representations of the l 1 -algebra of an inverse <strong>semi</strong>group, Trans. Amer. Math. Soc. 218 (1976) 361–<br />

396.<br />

[2] F. Bonsall, J. Duncan, Comp<strong>le</strong>te Normed Algebras, Springer-Ver<strong>la</strong>g, Berlin, 1973.<br />

[3] M.J. Crabb, The l 1 -algebra of a free inverse monoid, G<strong>la</strong>sgow University, preprint.<br />

[4] M.J. Crabb, W.D. Munn, The contracted l 1 -algebra of a McAlister monoid, G<strong>la</strong>sgow University, preprint.<br />

[5]J.Dixmier,Sur<strong>le</strong>sC ∗ -algèbres, Bull. Soc. Math. France 88 (1960) 95–112.<br />

[6] J. Dixmier, C ∗ -algèbres et <strong>le</strong>urs représentations, Gauthier–Vil<strong>la</strong>rs, Paris, 1969.<br />

[7] N. Weaver, A prime C ∗ -algebra that is not primitive, J. Funct. Anal. 203 (2003) 356–361.


Journal of Functional Analysis 236 (2006) 634–681<br />

www.elsevier.com/locate/jfa<br />

Quasiregu<strong>la</strong>r representations of the infinite-dimensional<br />

nilpotent group<br />

Sergio Albeverio a,1 , A<strong>le</strong>xandre Kosyak b,∗<br />

a Institut für Angewandte Mathematik, Universität Bonn, Wege<strong>le</strong>rstr. 6, D-53115 Bonn, Germany<br />

b Institute of Mathematics, Ukrainian National Aca<strong>de</strong>my of Sciences, 3 Tereshchenkivs’ka, Kyiv 01601, Ukraine<br />

Received 7 February 2006; accepted 13 March 2006<br />

Communicated by Paul Malliavin<br />

Abstract<br />

In the present work an analog of the quasiregu<strong>la</strong>r representation which is well known for locally-compact<br />

groups is constructed for the nilpotent infinite-dimensional group BN 0 and a criterion for its irre<strong>du</strong>cibility is<br />

presented. This construction uses the infinite tensor pro<strong>du</strong>ct of arbitrary Gaussian mea<strong>sur</strong>es in the spaces<br />

Rm with m>1 extending in a rather subt<strong>le</strong> way previous work of the second author for the infinite tensor<br />

pro<strong>du</strong>ct of one-dimensional Gaussian mea<strong>sur</strong>es.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Infinite-dimensional groups; Nilpotent groups; Quasiregu<strong>la</strong>r representations; Irre<strong>du</strong>cibility; Infinite tensor<br />

pro<strong>du</strong>cts; Gaussian mea<strong>sur</strong>es; Ismagilov conjecture<br />

1. Intro<strong>du</strong>ction<br />

1.1. The setting and the main results<br />

Let (X, B) be a mea<strong>sur</strong>ab<strong>le</strong> space and <strong>le</strong>t Aut(X) <strong>de</strong>note the group of all mea<strong>sur</strong>ab<strong>le</strong><br />

automorphisms of the space X. With any mea<strong>sur</strong>ab<strong>le</strong> action α : G → Aut(X) of a group<br />

* Corresponding author. Fax: 38044 2352010.<br />

E-mail addresses: albeverio@uni.bonn.<strong>de</strong> (S. Albeverio), kosyak01@yahoo.com, kosyak@imath.kiev.ua<br />

(A. Kosyak).<br />

1 SFB 611, BIGS, IZKS, Bonn, BiBoS, Bie<strong>le</strong>feld–Bonn, Germany; CERFIM, Locarno; Acca<strong>de</strong>mia di Architettura,<br />

USI, Mendrisio, Switzer<strong>la</strong>nd.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.013


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 635<br />

G on the space X and a G-quasi-invariant mea<strong>sur</strong>e μ on X one can associate a unitary<br />

representation π α,μ,X : G → U(L2 (X, μ)), of the group G by the formu<strong>la</strong> (π α,μ,X<br />

t f )(x) =<br />

(dμ(αt−1(x))/dμ(x)) 1/2f(αt−1(x)), f ∈ L2 (X, μ). Let us set α(G) ={αt ∈ Aut(X) | t ∈ G}.<br />

Let α(G) ′ be the centralizer of the subgroup α(G) in Aut(X): α(G) ′ ={g ∈ Aut(X) |{g,αt}=<br />

gαtg−1α −1<br />

t = e ∀t ∈ G}. The following conjecture has been discussed in [23–25].<br />

Conjecture 1. The representation π α,μ,X : G → U(L 2 (X, μ)) is irre<strong>du</strong>cib<strong>le</strong> if and only if :<br />

(1) μ g ⊥ μ ∀g ∈ α(G) ′ \{e} (where ⊥ stands for singu<strong>la</strong>r),<br />

(2) the mea<strong>sur</strong>e μ is G-ergodic.<br />

We recall that a mea<strong>sur</strong>e μ is G-ergodic if f(αt(x)) = f(x)∀t ∈ G implies f(x)= const μ<br />

a.e. for all functions f ∈ L1 (X, μ).<br />

In this paper we shall prove Conjecture 1 in the case where G is the infinite-dimensional<br />

nilpotent group G = BN 0 of finite upper-triangu<strong>la</strong>r matrices of infinite or<strong>de</strong>r with unities on the<br />

diagonal, the space X = Xm being the set of <strong>le</strong>ft cosets Gm \ BN , Gm being suitab<strong>le</strong> subgroups<br />

of the group BN of all upper-triangu<strong>la</strong>r matrices of infinite or<strong>de</strong>r with unities on the diagonal,<br />

and μ an infinite tensor pro<strong>du</strong>ct of Gaussian mea<strong>sur</strong>es on the spaces Rm with some fixed m>1.<br />

A more <strong>de</strong>tai<strong>le</strong>d exp<strong>la</strong>nation of the concepts used here is given in the following sections.<br />

1.2. Regu<strong>la</strong>r and quasiregu<strong>la</strong>r representations of locally compact groups<br />

Let G be a locally compact group. The right ρ (respectively <strong>le</strong>ft λ) regu<strong>la</strong>r representation of<br />

the group G is a particu<strong>la</strong>r case of the representation π α,μ,X with the space X = G, the action<br />

α being the right action α = R (respectively the <strong>le</strong>ft action α = L), and the mea<strong>sur</strong>e μ being the<br />

right invariant Haar mea<strong>sur</strong>e on the group G (see, for examp<strong>le</strong>, [8,16,17,37]).<br />

A quasiregu<strong>la</strong>r representation of a locally compact group G is also a particu<strong>la</strong>r case of the<br />

representation π α,μ,X (see, for examp<strong>le</strong>, [37, p. 27]) with the space X = H \ G, where H is a<br />

subgroup of the group G, the action α being the right action of the group G on the space X and<br />

the mea<strong>sur</strong>e μ being some quasi-invariant mea<strong>sur</strong>e on the space X (this mea<strong>sur</strong>e is unique up to<br />

a sca<strong>la</strong>r multip<strong>le</strong>). We remark that in [16,17] this representation has also been cal<strong>le</strong>d geometric<br />

representation.<br />

1.3. Analogs of the regu<strong>la</strong>r and quasiregu<strong>la</strong>r representations of infinite-dimensional groups and<br />

the Ismagilov conjecture<br />

In the present artic<strong>le</strong> we will consi<strong>de</strong>r the approach which <strong>de</strong>als with analogs for infinitedimensional<br />

groups of the regu<strong>la</strong>r and quasiregu<strong>la</strong>r representations of finite-dimensional groups.<br />

Let G be an infinite-dimensional topological group. To <strong>de</strong>fine an analog of the regu<strong>la</strong>r representation,<br />

<strong>le</strong>t us consi<strong>de</strong>r some topological group G, containing the initial group G as a <strong>de</strong>nse<br />

subgroup, i.e. G = G (G being the clo<strong>sur</strong>e of G). Suppose we have some quasi-invariant mea<strong>sur</strong>e<br />

μ on X = G with respect to the right action of the group G, i.e.α = R, Rt(x) = xt −1 .Inthis<br />

case we shall call the representation π α,μ,G an analog of the regu<strong>la</strong>r representation. We shall<br />

<strong>de</strong>note this representation by T R,μ , and the Conjecture 1 is re<strong>du</strong>ced to the following Ismagilov<br />

conjecture.


636 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Conjecture 2. (Ismagilov, 1985) The right regu<strong>la</strong>r representation T R,μ : G → U(L 2 (G, μ)) is<br />

irre<strong>du</strong>cib<strong>le</strong> if and only if :<br />

(1) μ Lt ⊥ μ ∀t ∈ G \{e},<br />

(2) the mea<strong>sur</strong>e μ is G-ergodic.<br />

Remark 3. In the case of the right regu<strong>la</strong>r representation, the group α(G) ′ = R(G) ′ ⊂ Aut(G)<br />

obviously contains the group L(G), the image of the group G with respect to the <strong>le</strong>ft action.<br />

The work [11] initiated the study of representations of current groups, i.e. groups C(X,U) of<br />

continuous mappings X ↦→ U, where X is a finite-dimensional Riemannian manifold and U is a<br />

finite-dimensional Lie group.<br />

The regu<strong>la</strong>r representation of infinite-dimensional groups, in the case of current groups, was<br />

studied firstly in [1,4,5,14] (see also the book [6]). An analog of the regu<strong>la</strong>r representation for an<br />

arbitrary infinite-dimensional group G, usingaG-quasi-invariant mea<strong>sur</strong>e on some comp<strong>le</strong>tion<br />

G of such a group, is <strong>de</strong>fined in [18,20].<br />

For X = S1 , U a compact or non-compact connected Lie group, Wiener mea<strong>sur</strong>es on the loop<br />

groups G = C(X,U) were constructed and their quasi-invariance were proved in [1,4–6,28–32].<br />

and the mea<strong>sur</strong>e μ<br />

Conjecture 2 was formu<strong>la</strong>ted by R.S. Ismagilov for the group G = B N 0<br />

being the pro<strong>du</strong>ct of arbitrary one-dimensional centered Gaussian mea<strong>sur</strong>es on the group G =<br />

B N and was proved for this case in [18,19].<br />

The first result in this direction was proved in [33]. For the comp<strong>le</strong>x infinite-dimensional Borel<br />

group Bor c,N<br />

0 and the standard Gaussian mea<strong>sur</strong>e on its comp<strong>le</strong>tion Bor c,N the irre<strong>du</strong>cibility of<br />

the corresponding regu<strong>la</strong>r representation was proved there. Here Bor c,N<br />

0 (respectively Bor c,N )is<br />

the group of matrices of the form x = exp t + s where t is a diagonal matrix with a finite number<br />

of nonzero real e<strong>le</strong>ments (respectively arbitrary real e<strong>le</strong>ments) and s is a finite (respectively<br />

arbitrary) comp<strong>le</strong>x strictly upper-triangu<strong>la</strong>r matrix.<br />

For the pro<strong>du</strong>ct of arbitrary one-dimensional mea<strong>sur</strong>es on the group B N Conjecture 2 was<br />

proved in [21] un<strong>de</strong>r some technical assumptions on the mea<strong>sur</strong>e.<br />

In [20] Conjecture 2 was proved for the groups of the interval and circ<strong>le</strong> diffeomorphisms. For<br />

the group of the interval diffeomorphisms the Shavgulidze mea<strong>sur</strong>e [35] was used, the image of<br />

the c<strong>la</strong>ssical Wiener mea<strong>sur</strong>e with respect to some bijection. For the group of circ<strong>le</strong> diffeomorphisms<br />

the Malliavin mea<strong>sur</strong>e [30] was used.<br />

Whether Conjecture 2 holds in the general case is an open prob<strong>le</strong>m.<br />

In [25] it was shown that Conjecture 1 holds for the in<strong>du</strong>ctive limit G = SL0(2∞, R) =<br />

lim<br />

−→n SL(2n − 1, R), of the special linear groups (simp<strong>le</strong> groups) acting on a strip of <strong>le</strong>ngth m ∈ N<br />

in the space of real matrices which are infinite in both directions, the mea<strong>sur</strong>e μ being a pro<strong>du</strong>ct<br />

Gaussian mea<strong>sur</strong>e.<br />

Let us consi<strong>de</strong>r the special case of a G-space, namely the homogeneous space X = H \ G,<br />

where H is a subgroup of the group G and μ is some quasi-invariant mea<strong>sur</strong>e on X (if it exists)<br />

with respect to the right action R of the group G on the homogeneous space H \ G. In this case<br />

we call the corresponding representation π R,μ,H\G an analog of the quasiregu<strong>la</strong>r or geometric<br />

representation of the group G (see [22]).<br />

In [2] Conjecture 1 was proved for the solvab<strong>le</strong> infinite-dimensional real Borel group G =<br />

Bor N 0 acting on G-spaces Xm , m ∈ N, where X m is the set of <strong>le</strong>ft cosets Gm \ Bor N , and Gm<br />

is some subgroups of the group Bor N of all upper-triangu<strong>la</strong>r matrices of infinite or<strong>de</strong>r with non-


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 637<br />

zero e<strong>le</strong>ments on the diagonal. The mea<strong>sur</strong>e μ on Xm is the pro<strong>du</strong>ct of infinitely many onedimensional<br />

Gaussian mea<strong>sur</strong>es on R.<br />

In [23,24] Conjecture 1 was proved for the nilpotent group G = BN 0 and some G-spaces Xm ,<br />

m ∈ N, being the set of <strong>le</strong>ft cosets Gm \B N , where Gm are some subgroups of the group BN .Here<br />

the mea<strong>sur</strong>e μ on Xm is the infinite pro<strong>du</strong>ct of arbitrary one-dimensional Gaussian mea<strong>sur</strong>es<br />

on R. In this case the variab<strong>le</strong>s xpq,1 p


638 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Obviously, BN 0 = lim −→n<br />

B(n,R) is the in<strong>du</strong>ctive limit of the group B(n,R) of real uppertriangu<strong>la</strong>r<br />

matrices with units on the principal diagonal<br />

<br />

B(n,R) = I + <br />

<br />

xkr ∈ R<br />

1k


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 639<br />

Proof. The right action Rt for t ∈ B N 0<br />

the point x ∈ X m . ✷<br />

changes linearly only a finite number of coordinates of<br />

Now we can <strong>de</strong>fine the representation associated with the right action<br />

in the natural way, i.e.<br />

T R,μm B : B N 0 → UL 2 X m ,μ m B<br />

R,μ<br />

T m B<br />

t f (x) = dμ m −1<br />

B Rt (x) dμ m B (x) 1/2 −1<br />

f Rt (x) .<br />

Theorem 5. For the mea<strong>sur</strong>e μm B the following four statements are equiva<strong>le</strong>nt:<br />

(i) the representation T R,μm B is irre<strong>du</strong>cib<strong>le</strong>;<br />

(ii) (μm B )Lt ⊥ μm B ∀t ∈ B(m,R) \{e};<br />

(iii) (μm B )Lexp(tEpq) ⊥ μm B ∀t ∈ R \{0} ∀1 pm), from the fact that the right action Rt for t ∈ BN 0<br />

finite number of coordinates of the point x ∈ Xm , and that the group Gm 0 = Gm ∩ BN 0 ⊂ Xm acts<br />

transitively on itself. In fact it is shown that the mea<strong>sur</strong>e is ergodic with respect to the action of<br />

the subgroup Gm 0 ⊂ BN 0 . ✷


640 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

3. I<strong>de</strong>a of the proof of irre<strong>du</strong>cibility<br />

Proof of Theorem 5. The proof of Theorem 5 is organized as follows:<br />

(i) ⇒ (ii) ⇒ (iii) ⇒ (iv) ⇒ (i).<br />

The parts (i) ⇒ (ii) ⇒ (iii) are evi<strong>de</strong>nt. The part (iii) ⇔ (iv) follows from Lemma 8, which is<br />

based on the Kakutani criterion [15].<br />

The i<strong>de</strong>a of the proof of irre<strong>du</strong>cibility, i.e. the part (iv) ⇒ (i). Let us <strong>de</strong>note by A m the von<br />

Neumann algebra generated by the representation T R,μm B<br />

A m = T R,μm B<br />

t<br />

| t ∈ G ′′ .<br />

We show that (iv) ⇒[(Am ) ′ ⊂ L∞ (Xm ,μm B )]⇒(i). Let the inclusion (Am ) ′ ⊂ L∞ (Xm ,μm B )<br />

holds. Using the ergodicity of the mea<strong>sur</strong>e μm B (Lemma 6) this proves the irre<strong>du</strong>cibility. In<strong>de</strong>ed<br />

in this case an operator A ∈ (Am ) ′ should be the operator of multiplication (since (Am ) ′ ⊂<br />

L∞ (Xm ,μm B )) by some essentially boun<strong>de</strong>d function a ∈ L∞ (Xm ,μm B ). The commutation re<strong>la</strong>tion<br />

[A,T R,μm B<br />

t ]=0 ∀t ∈ BN 0 implies a(R−1 t (x)) = a(x) (mod μm B ) ∀t ∈ BN 0 , so by ergodicity of<br />

the mea<strong>sur</strong>e μ m B with respect to the right action of the group BN 0 on the space Xm we conclu<strong>de</strong><br />

that A = a = const (mod μm B ). This then proves the irre<strong>du</strong>cibility in Theorem 5, i.e. the part<br />

[(Am ) ′ ⊂ L∞ (Xm ,μm B )]⇒(i).<br />

The proof of the remaining part, i.e. the implication (iv) ⇒[(Am ) ′ ⊂ L∞ (Xm ,μm B )] is based<br />

on the fact that the operators of multiplication by in<strong>de</strong>pen<strong>de</strong>nt variab<strong>le</strong>s xpq, 1 p m, p


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 641<br />

and the series SL pq (μm ) and Σr pq (m) are <strong>de</strong>fined in Lemmas 8 and 15 (see also (18)). This is done<br />

in Appendices A–C.<br />

In Appendix A we <strong>de</strong>fine the generalization of the characteristic polynomial for matrix C and<br />

establish some its properties. These properties are used then in Appendices B and C. For a matrix<br />

C ∈ Mat(k, C) we set<br />

Gk(λ) = <strong>de</strong>t Ck(λ), where Ck(λ) = C +<br />

k<br />

λrErr, λ= (λ1,...,λk) ∈ C k .<br />

Lemma A. (See Appendix A, Lemma A.7) For a positive <strong>de</strong>finite matrix C ∈ Mat(k, C), λ ∈ R k<br />

with λr 0, r = 1,...,k, we have<br />

r=1<br />

∂ Gk(λ)<br />

0,<br />

∂λp Gl(λ)<br />

where Gl(λ) = M 12...l<br />

12...l (Ck(λ)) and 1 p l k.<br />

The proof of Lemma A is based on the following inequality (see Lemma A.6).<br />

Lemma B. (Hadamard–Ficher’s inequality [12,13], see also [27]) Let C ∈ Mat(m, R) be a positive<br />

<strong>de</strong>finite matrix and ∅⊆α, β ⊆{1,...,m}. Then<br />

<br />

<br />

<strong>de</strong>t Cα <strong>de</strong>t<br />

<strong>de</strong>t Cα∪β<br />

Cα∩β<br />

<strong>de</strong>t Cβ<br />

<br />

<br />

<br />

=<br />

<br />

<br />

<br />

M(α) M(α∩ β) <br />

<br />

M(α∪ β) M(β) 0,<br />

where Cα for α ={α1,...,αs} <strong>de</strong>notes the matrix which entries lie on the intersection of<br />

α1,...,αs rows and α1,...,αs columns of the matrix C and M(α) = M α α (C) = <strong>de</strong>t Cα are corresponding<br />

minors of the matrix C.<br />

The “best” approximation of xpq by the generators A R,m<br />

kn is based on the exact computation<br />

of the matrix e<strong>le</strong>ments<br />

φp(t) = T R,μm B<br />

t 1, 1 , t = I +<br />

p<br />

r=1<br />

trErn, (tr) p<br />

r=1 ∈ Rp ,<br />

of the representation T R,μm B and their generalization (see Appendix B, Lemma B.1), and on<br />

the finding the appropriate combinations of operator functions of the generators A R,m<br />

kn (see Remark<br />

13) to approximate the operators of multiplication by xpq.<br />

Finally the proof of the inequality Σm >CSm, is based on Lemmas A, B and 16 <strong>de</strong>aling with<br />

some inequalities involving the generalized characteristic polynomials. Lemma 16 is proved in<br />

Appendix C.


642 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Remark 7. We shall firstly prove the approximation of xkn in the above sense for the one vector<br />

1 ∈ L2 (Xm ,μm B ). Secondly, the approximation also holds for some <strong>de</strong>nse set D of analytic<br />

vectors in the space L2 (Xm ,μm B )<br />

<br />

D = X α =<br />

<br />

1km, k


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 643<br />

<br />

<br />

<strong>de</strong>t C<br />

= exp −<br />

(2π) m 1<br />

exp(tEpq)<br />

2<br />

∗ C exp(tEpq)x, x <br />

dx = dμBpq(t)(x),<br />

where (Bpq(t)) −1 = Cpq(t) = exp(tEpq) ∗ C exp(tEpq) (we note that <strong>de</strong>t C = <strong>de</strong>t Cpq(t)).<br />

Hence, using (2) we get<br />

We shall prove that<br />

H μ LI+tEpq <br />

B ,μB =<br />

<strong>de</strong>t Cpq(t) + C<br />

2<br />

<strong>de</strong>t Cpq(t) <strong>de</strong>t C<br />

<strong>de</strong>t 2 Cpq(t)+C<br />

2<br />

1/4<br />

<br />

=<br />

<strong>de</strong>t C<br />

<strong>de</strong>t Cpq(t)+C<br />

2<br />

1/2<br />

. (3)<br />

= <strong>de</strong>t C + t2<br />

4 cppA q q(C), (4)<br />

where A p q (C), 1 p,q m, <strong>de</strong>note the cofactors of the matrix C corresponding to the row p<br />

and the column q. Wehave<br />

hence<br />

<strong>de</strong>t Cpq(t)+C<br />

2<br />

<strong>de</strong>t C<br />

<br />

and finally, using (3) we get<br />

where<br />

H μ mLI+tEpq m<br />

B ,μB =<br />

B (n) = <br />

1r,sm<br />

= <strong>de</strong>t C + t2<br />

<strong>de</strong>t C<br />

<strong>de</strong>t Cpq(t)+C<br />

2<br />

∞<br />

n=q+1<br />

4 cppA q q(C)<br />

= 1 +<br />

<strong>de</strong>t C<br />

t2<br />

4 cppbqq,<br />

1/2<br />

<br />

= 1 + t2<br />

4 cppbqq<br />

−1/2 H μ LI+tEpq<br />

B (n)<br />

<br />

,μB (n) =<br />

∞<br />

n=q+1<br />

b (n)<br />

rs Ers and C (n) := B (n) −1 = <br />

<br />

1 + t2<br />

4 c(n)<br />

1r,sm<br />

pp b(n) qq<br />

c (n)<br />

rs Ers.<br />

−1/2<br />

,<br />

So using the properties of the Hellinger integral for two Gaussian mea<strong>sur</strong>es we conclu<strong>de</strong> that<br />

m LI+tEpq m<br />

μB ⊥ μB ∀t ∈ R \{0} ⇔<br />

∞<br />

n=q+1<br />

<br />

1 + t2<br />

4 c(n)<br />

⇔ S L m<br />

pq μB =∞.<br />

pp b(n) qq<br />

−1/2<br />

= 0


644 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

To prove (4) we set Cpq(t) = exp(tEpq) ∗ C exp(tEpq). Wehaveform ∈ N and 1 p 0, k= 1, 2,...,n.<br />

We use the same result in a slightly different form with bk = 0, k = 1, 2,...,n,<br />

min<br />

x∈R n<br />

n<br />

k=1<br />

akx 2 k<br />

<br />

<br />

<br />

<br />

<br />

n<br />

<br />

n<br />

xkbk = 1 =<br />

k=1<br />

b 2 k<br />

ak<br />

k=1<br />

−1<br />

. (5)


The minimum is realized for<br />

S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 645<br />

xk = bk<br />

<br />

n<br />

ak<br />

k=1<br />

For any subset I ⊂ N <strong>le</strong>t us <strong>de</strong>note as before by 〈fn | n ∈ I〉 the clo<strong>sur</strong>e of the linear space<br />

generated by the set of vectors (fn | n ∈ I) in a Hilbert space H .<br />

We note that the distance d(fn+1;〈f1,...,fn〉) of the vector fn+1 in H from the hyperp<strong>la</strong>ne<br />

〈f1,...,fn〉 may be calcu<strong>la</strong>ted in terms of the Gram <strong>de</strong>terminants Γ(f1,f2,...,fk) corresponding<br />

to the set of vectors f1,f2,...,fk (see [10]):<br />

d fn+1;〈f1,...,fn〉 = min<br />

t=(tk)∈R n<br />

<br />

<br />

<br />

<br />

fn+1 +<br />

b 2 k<br />

ak<br />

n<br />

k=1<br />

−1<br />

.<br />

tkfk<br />

<br />

<br />

<br />

<br />

<br />

2<br />

= Γ(f1,f2,...,fn+1)<br />

, (6)<br />

Γ(f1,f2,...,fn)<br />

where the Gram <strong>de</strong>terminant is <strong>de</strong>fined by Γ(f1,f2,...,fn) = <strong>de</strong>t γ(f1,f2,...,fn) and<br />

γ(f1,f2,...,fn) =: γn is the Gram matrix<br />

Lemma 10. We have<br />

⎛<br />

(f1,f1) (f1,f2) ...<br />

⎞<br />

(f1,fn)<br />

⎜ (f2,f1)<br />

γ(f1,f2,...,fn) = ⎜<br />

⎝<br />

(f2,f2) ...<br />

. ..<br />

(f2,fn) ⎟<br />

⎠<br />

(fn,f1) (fn,f2) ... (fn,fn)<br />

.<br />

d fn+1;〈f1,...,fn〉 =<br />

<strong>de</strong>t γn+1<br />

<strong>de</strong>t γn<br />

where dn+1 = ((f1,fn+1), (f2,fn+1),...,(fn,fn+1)) ∈ R n .<br />

Proof. We may write<br />

<br />

n<br />

<br />

<br />

<br />

k=1<br />

tkfk − fn+1<br />

<br />

2<br />

<br />

=<br />

<br />

n<br />

k,m=1<br />

tktm(fk,fk) − 2<br />

= (fn+1,fn+1) − γ −1<br />

n dn+1,dn+1<br />

<br />

,<br />

n<br />

k=1<br />

= (γnt,t)− 2(t, dn+1) + (fn+1,fn+1),<br />

tk(fk,fn+1) + (fn+1,fn+1)<br />

where t = (t1,t2,...,tn) ∈ Rn . Using (58) for An = γn we get<br />

(γnt,t)− 2(t, dn+1) = γn(t − t0), (t − t0) − γ −1<br />

n dn+1,dn+1<br />

<br />

,<br />

where t0 = γ −1<br />

n dn. Hence we get (see (6))<br />

min<br />

t=(tk)∈R n<br />

<br />

<br />

<br />

<br />

fn+1 −<br />

n<br />

k=1<br />

tkfk<br />

<br />

<br />

<br />

<br />

<br />

2<br />

= min<br />

t=(tk)∈Rn <br />

(γnt,t)− 2(t, dn+1) + (fn+1,fn+1)


646 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

= (fn+1,fn+1) − γ −1<br />

n dn+1,dn+1<br />

<br />

+ min<br />

t=(tk)∈Rn <br />

γn(t − t0), (t − t0) <br />

= (fn+1,fn+1) − γ −1<br />

n dn+1,dn+1<br />

<br />

. ✷<br />

Remark 11. In fact a more general result holds. Let us <strong>de</strong>note by An+1 the real non-necessarily<br />

symmetric matrix in R n+1 and by An its n × n block after crossing the e<strong>le</strong>ment in the <strong>la</strong>st column<br />

and row, by vn+1 = (a1n+1,a2n+1,...,ann+1), hn+1 = (an+11,an+12,...,an+1n) vectors vn+1,<br />

hn+1 ∈ R n . If <strong>de</strong>t An = 0 then we have<br />

an+1n+1 − A −1<br />

n vn+1,hn+1<br />

<strong>de</strong>t An+1<br />

= . (7)<br />

<strong>de</strong>t An<br />

Proof. It is sufficient to use the i<strong>de</strong>ntity (Schur–Frobenius <strong>de</strong>composition)<br />

The generators<br />

<br />

An<br />

An+1 =<br />

v t n+1<br />

hn+1 an+1n+1<br />

<br />

=<br />

Akn := A R,m<br />

kn<br />

<br />

An<br />

<br />

0 Id A−1 0 1<br />

= d<br />

dt T R,μm B<br />

n vt n+1<br />

hn+1 an+1n+1<br />

<br />

<br />

<br />

I+tEkn<br />

t=0<br />

<br />

. ✷<br />

of the one-parameter groups I + tEkn have the following form (on smooth functions of compact<br />

support):<br />

where<br />

k−1<br />

Akn =<br />

r=1<br />

xrkDrn + Dkn, 1 k m, k < n, Akn =<br />

m<br />

r=1<br />

xrkDrn, m


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 647<br />

is the image of the standard Fourier transform F m in the space L 2 (R m ,dx),i.e.Fmn =<br />

U(C (n) ) −1 F m U(B (n) ), where<br />

U B (n) <br />

L<br />

1/2<br />

dμB (n)(x)<br />

=<br />

dx<br />

2 (Rm ,μB (n))<br />

L 2 (R m ,dx)<br />

Fmn<br />

F m<br />

L2 (Rm ,μC (n))<br />

U C (n) <br />

dμC (n)(x)<br />

=<br />

dx<br />

L 2 (R m ,dx).<br />

Since the standard Fourier transform F m is <strong>de</strong>fined as follows:<br />

<br />

m 1<br />

F f (y) = √ exp i(y,x)f(x)dx,<br />

(2π) m<br />

and, for D = B (n) respectively D = C (n)<br />

<br />

dμD(x)<br />

U(D)=<br />

dx<br />

we have finally for Fmn:<br />

R m<br />

1/2<br />

=<br />

R m<br />

1<br />

((2π) m <br />

exp<br />

<strong>de</strong>t D) 1/4<br />

− 1<br />

4<br />

D −1 x,x <br />

,<br />

(Fmnf )(y) = U C (n)−1 m (n)<br />

F U B f (y)<br />

1<br />

=<br />

((2π) m <strong>de</strong>t C (n) <br />

1<br />

(n)<br />

exp C<br />

) 1/4 4<br />

−1 <br />

y,y <br />

1<br />

√<br />

(2π) m<br />

<br />

× exp i(y,x)f(x) (2π) m <strong>de</strong>t B (n) <br />

1/4<br />

exp − 1<br />

(n)<br />

B<br />

4<br />

−1 <br />

x,x <br />

dx<br />

= exp( 1<br />

<br />

4 ((C(n) ) −1y,y)) <br />

(2π) m <strong>de</strong>t C (n)<br />

Rm <br />

exp<br />

i(y,x) − 1<br />

4<br />

Using Fourier transform Fm we obtain for Akn = FmAkn(Fm) −1 :<br />

Akn = i<br />

where<br />

k−1<br />

<br />

r=1<br />

xrkyrn + ykn<br />

B (n) −1 x,x <br />

f(x)dx.<br />

1/2<br />

<br />

m<br />

, 1 k m


648 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

For a function f : X m → C we set<br />

<br />

Mf =<br />

X m<br />

f(x)dμ m B (x).<br />

To approximate the variab<strong>le</strong>s xpq, 1 p


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 649<br />

4. To make the expression Σr pq (s,α,m) in (11) <strong>la</strong>rger (to apply then the criterium in<br />

Lemma 12) we chose s (n) ∈ Rr such that<br />

<br />

rp<br />

(n)<br />

Mξn s 2 <br />

= max<br />

rp<br />

Mξn (s) 2 s∈R r<br />

(which is possib<strong>le</strong>, |Mξ rp<br />

n (s)| 2 being continuous and boun<strong>de</strong>d).<br />

5. With the same aim we chose α (n)<br />

k in such a way that<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Aqn − xpqDpn +<br />

m<br />

k=1,k=q<br />

α (n)<br />

k Akn<br />

2<br />

<br />

<br />

1<br />

= min<br />

<br />

(tk)∈R m−1<br />

<br />

<br />

<br />

<br />

Aqn − xpqDpn +<br />

6. The right-hand si<strong>de</strong> of the previous expression is equal (see (6)) to<br />

where<br />

Γ(g1,g2,...,g p q ,...,gm)<br />

Γ(g1,g2,...,gq−1,gq−1,...,gm) ,<br />

m<br />

k=1,k=q<br />

tkAkn<br />

<br />

2<br />

<br />

1<br />

.<br />

<br />

gk := gkn := Akn1, 1 k m, k = q, g p q := g p qn := (Aqn − xpqDpn)1. (12)<br />

Proof of Lemma 12. If we put rp<br />

n tnMξn (s (n) ) = 1 we get<br />

<br />

<br />

<br />

<br />

<br />

r<br />

tn exp s<br />

<br />

n<br />

l=1<br />

(n)<br />

l Aln<br />

<br />

m<br />

α<br />

k=1<br />

(n)<br />

k Akn<br />

<br />

2<br />

<br />

− xpq 1<br />

<br />

<br />

<br />

<br />

<br />

r<br />

= tn exp s<br />

<br />

n<br />

l=1<br />

(n)<br />

l Aln<br />

<br />

m<br />

Aqn − xpqDpn + xpqDpn + α<br />

k=1,k=q<br />

(n)<br />

k Akn<br />

<br />

2<br />

<br />

− xpq 1<br />

<br />

<br />

<br />

<br />

<br />

r<br />

= tn xpq Dpn exp s<br />

<br />

n<br />

l=1<br />

(n)<br />

l Aln<br />

<br />

− Mξ rp<br />

(n)<br />

n s <br />

<br />

r<br />

+ exp s (n)<br />

l Aln<br />

<br />

m<br />

Aqn − xpqDpn + α (n)<br />

k Akn<br />

<br />

2<br />

<br />

1<br />

<br />

l=1<br />

k=1,k=q<br />

= <br />

t<br />

n<br />

2 <br />

n xpq 2<br />

<br />

<br />

<br />

r<br />

<br />

Dpn exp s<br />

<br />

l=1<br />

(n)<br />

l Aln<br />

<br />

− Mξ rp<br />

(n)<br />

n s <br />

<br />

2<br />

<br />

1<br />

<br />

<br />

<br />

<br />

+ <br />

exp<br />

<br />

r<br />

s (n)<br />

l Aln<br />

<br />

m<br />

Aqn − xpqDpn + α (n)<br />

k Akn<br />

<br />

2<br />

<br />

1<br />

<br />

= <br />

n<br />

t 2 n<br />

l=1<br />

<br />

xpq 2 c (n)<br />

pp − Mξ rp<br />

(n)<br />

n s 2 k=1,k=q


650 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

<br />

<br />

<br />

+ <br />

exp<br />

<br />

r<br />

l=1<br />

s (n)<br />

l Aln<br />

<br />

Aqn − xpqDpn +<br />

m<br />

k=1,k=q<br />

α (n)<br />

k Akn<br />

where we have used the equality ξ − Mξ 2 =ξ 2 −|Mξ| 2 :<br />

<br />

<br />

<br />

r<br />

<br />

Dpn exp<br />

<br />

l=1<br />

s (n)<br />

l Aln<br />

<br />

− Mξ rp<br />

(n)<br />

n s <br />

<br />

<br />

<br />

1<br />

<br />

<br />

2<br />

<br />

1<br />

,<br />

<br />

=Dpn1 2 − rp<br />

(n)<br />

Mξn s 2 = c (n)<br />

pp − rp<br />

(n)<br />

Mξn s 2 .<br />

Definition. We shall say that two series <br />

n an and <br />

n bn with positive members are equiva<strong>le</strong>nt<br />

and shall <strong>de</strong>note this by <br />

n an ∼ <br />

n bn if they are convergent or divergent simultaneously. We<br />

note that if an > 0, bn > 0, n ∈ N, then we have<br />

an<br />

∼ an<br />

. (13)<br />

min<br />

t∈R N<br />

an + bn<br />

n∈N<br />

bn<br />

n∈N<br />

Using (5) we get, setting b = (Mξ rp<br />

n (s (n) )) m+1+N<br />

n=m+1 ∈ RN , N ∈ N,<br />

∼<br />

m+1+N<br />

<br />

n=m+1<br />

m+1+N<br />

<br />

n=m+1<br />

<br />

r<br />

tn exp s<br />

l=1<br />

(n)<br />

l Aln<br />

<br />

m<br />

α<br />

k=1<br />

(n)<br />

k Akn<br />

<br />

2<br />

<br />

<br />

<br />

− xpq 1<br />

(t, b) =−1<br />

<br />

|Mξ rp<br />

n (s (n) )| 2<br />

c (n)<br />

pp −|Mξ rp<br />

n (s (n) )| 2 +(Aqn − xpqDpn + m k=1,k=p α (n)<br />

2<br />

k<br />

Akn)1 2<br />

−1<br />

. ✷<br />

DuetoRemark9weshallwriteC (respectively Ĉ) instead of C (n) (respectively Ĉ (n) ), where<br />

⎛<br />

Ĉ (n) ⎜<br />

= ⎜<br />

⎝<br />

c (n)<br />

11<br />

c (n)<br />

12<br />

c (n)<br />

1m<br />

⎛<br />

C (n) ⎜<br />

= ⎜<br />

⎝<br />

c (n)<br />

11<br />

c (n)<br />

11<br />

c (n)<br />

12<br />

c (n)<br />

1m c(n)<br />

2m<br />

c (n)<br />

12 ... c (n)<br />

1m<br />

c (n)<br />

22 ... c (n)<br />

⎞<br />

⎟<br />

2m ⎟<br />

.<br />

⎟<br />

.. ⎠ ,<br />

... c(n)<br />

mm<br />

c (n)<br />

12 ... c (n)<br />

+ c(n)<br />

22 ...<br />

1m<br />

c (n)<br />

. ..<br />

2m<br />

c (n)<br />

2m ... c (n)<br />

11 + c(n)<br />

22 +···+c(n)<br />

⎞<br />

⎟<br />

⎠<br />

mm<br />

.<br />

Remark 14. To simplify the further computations we assume that the mea<strong>sur</strong>es μ B (n) for 2 <br />

n m are standard: B (n) = I . Since μ m B = ∞ n=2 μ B (n) this assumption, which only concerns<br />

finitely many of the μ B (n)’s, does not change the equiva<strong>le</strong>nce c<strong>la</strong>ss of the initial mea<strong>sur</strong>e μ m B and<br />

the equiva<strong>le</strong>nce c<strong>la</strong>ss of the corresponding representation T R,μm B .


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 651<br />

Using this remark, notation (12) and Fourier transforms we conclu<strong>de</strong> that<br />

Γ(g1,g2,...,gm) = <strong>de</strong>t Ĉ, i.e. Γ(g1n,g2n,...,gmn) = <strong>de</strong>t Ĉ (n) , (14)<br />

since (gq,gp) = (Ĉ)pq, 1 p,q m. In<strong>de</strong>ed for p = q we have<br />

(gqn,gpn) = (gpn,gqn) =<br />

(gpn,gpn) =<br />

<br />

p−1<br />

<br />

<br />

<br />

<br />

r=1<br />

p−1<br />

<br />

r=1<br />

xrpyrn + ypn<br />

<br />

<br />

<br />

<br />

<br />

q−1<br />

xrpyrn + ypn, xsqysn + yqn<br />

2<br />

p−1 <br />

=<br />

r=1<br />

s=1<br />

(we reinserted here the upper in<strong>de</strong>x n in c (n)<br />

pq for c<strong>la</strong>rity).<br />

<br />

xrp 2 yrn 2 +ypn 2 =<br />

= (ypn,yqn) = c (n)<br />

pq ,<br />

p<br />

r=1<br />

c (n)<br />

rr = Ĉ (n)<br />

pp<br />

In the following we shall need a variant of Lemma 12 using Remark 13 rep<strong>la</strong>cing the<br />

|Mξ rp<br />

n (s)| by its maximum Ξ rp<br />

n . Let us set (see (10) for <strong>de</strong>finition of ξ rp<br />

n (s))<br />

<br />

n = max<br />

rp<br />

Mξn (s) 2 . (15)<br />

Ξ rp<br />

s∈R r<br />

Now we see that using s and α as in parts 4 and 5 of Remark 13 we have<br />

Σ r pq (s,α,m)<br />

= <br />

n<br />

(13)<br />

∼ <br />

n<br />

(15)<br />

= <br />

n<br />

max s (n) ∈R r |Mξ rp<br />

n (s (n) )| 2<br />

c (n)<br />

pp − maxs (n) ∈Rr |Mξ rp<br />

n (s (n) )| 2 +(Aqn − xpqDpn + m k=1,k=p α (n)<br />

c (n)<br />

max s (n) ∈R r |Mξ rp<br />

n (s (n) )| 2<br />

pp +(Aqn − xpqDpn + m k=1,k=p α (n)<br />

c (n)<br />

pp +<br />

Remark 9<br />

= <br />

n<br />

Ξ rp<br />

n<br />

Γ(g1n,g2n,...,g p qn,...,gmn)<br />

Γ(g1n,g2n,...,gq−1n,gq+1n,...,gmn)<br />

k<br />

Akn)1 2<br />

Ξ rp Γ(g1,g2,...,gq−1,gq+1,...,gm)<br />

cppΓ(g1,g2,...,gq−1,gq+1,...,gm) + Γ(g1,g2,...,g p q ,...,gm)<br />

k<br />

Akn)1 2<br />

= Σ r Ξ<br />

pq (m) :=<br />

n<br />

rpΓ(g1,g2,...,gq−1,gq+1,...,gm) (14) Ξ<br />

=<br />

Γ(g1,g2,...,gm)<br />

n<br />

rp<br />

n A q q(Ĉ (n) )<br />

<strong>de</strong>t Ĉ (n)<br />

.<br />

For the <strong>la</strong>tter equality we have used the fact that<br />

cppΓ(g1,g2,...,gq−1,gq+1,...,gm) + Γ g1,g2,...,g p <br />

q ,...,gm = Γ(g1,g2,...,gm),<br />

which follows from (26). In<strong>de</strong>ed it is sufficient to take in (26) C = Ĉ − cppEqq and λq = cpp.<br />

Then we have


652 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Γ(g1,g2,...,gm) = <strong>de</strong>t Ĉ = <strong>de</strong>t(Ĉ − cppEqq + cppEqq)<br />

So we have proved the following <strong>le</strong>mma.<br />

= <strong>de</strong>t(Ĉ − cppEqq) + cppA q q(Ĉ − cppEqq)<br />

= Γ g1,g2,...,g p <br />

q ,...,gm + cppΓ(g1,g2,...,gq−1,gq+1,...,gm).<br />

Lemma 15. Let 1 r p


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 653<br />

(see (40)) and theirs generalizations (see (42)). We cannot calcu<strong>la</strong>te explicitly<br />

<br />

n = max<br />

pq<br />

Mξn (t) 2 ,<br />

Ξ pq<br />

t∈R p<br />

but we are ab<strong>le</strong> by Lemmas B.1 and B.2 to obtain the estimation Ξ pq<br />

n >Ψ pq<br />

n ,<br />

Ψ pq<br />

n :=<br />

(M 12...p−1p<br />

12...p−1q (C(n)<br />

p,q)) 2 exp(−1)<br />

M 12...p−1<br />

12...p−1 (C(n) p )M 12...p<br />

12...p (C(n) p ) + p k=2<br />

ˆλk(A<br />

p<br />

k (C(n) p )) 2<br />

(see (47) and (48)). The crucial for proving (17) is Lemma 16 <strong>de</strong>aling with some inequalities<br />

involving the generalized characteristic polynomials. This <strong>le</strong>mma is proved in Appendix C.<br />

We use the notations of Lemma 8 (see Remark 9):<br />

Let<br />

S L m<br />

pq μB =<br />

∞<br />

n=q+1<br />

c (n)<br />

pp b(n)<br />

qq =<br />

∞<br />

n=q+1<br />

c (n)<br />

pp A q q(C (n)<br />

m )<br />

<strong>de</strong>t C (n)<br />

m<br />

=<br />

∞<br />

n=q+1<br />

cppA q q(Cm)<br />

<strong>de</strong>t Cm<br />

λ = (λk) m k=1 ∈ Rm , ˆλ = (ˆλk) m k=1 , k−1<br />

ˆλ1 = 0, ˆλk = crr, 2 k m, (19)<br />

fq = e <br />

Ĉm =<br />

1rp


654 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

We have rep<strong>la</strong>ced the series<br />

with the equiva<strong>le</strong>nt one<br />

S L m<br />

pq μB =<br />

S L m<br />

pq μB ∼<br />

If we use the equality Ĉm = Cm(ˆλ), we get<br />

Σm :=<br />

<br />

1rp


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 655<br />

Corol<strong>la</strong>ry 17. If SL kn (μm B ) =∞for some 1 k


656 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Proof. Probably Lemma A.1 is known in the literature, but since we do not know any precise<br />

reference, we provi<strong>de</strong> here a direct proof. Obviously Gm(λ) is a polynomial in the variab<strong>le</strong>s<br />

λ = (λ1,λ2,...,λm) ∈ C m of or<strong>de</strong>r m. A direct calcu<strong>la</strong>tion gives us<br />

hence<br />

∂r <br />

<br />

1 0 ... 0 0 ... 0 <br />

<br />

<br />

0 1 ... 0 0 ... 0 <br />

<br />

<br />

. ..<br />

. <br />

..<br />

<br />

Gm(λ)<br />

<br />

<br />

<br />

= 0 0 ... 1 0 ... 0<br />

<br />

,<br />

∂λ1∂λ2 ...∂λr <br />

c1r+1 c2r+1 ... crr+1 cr+1r+1 + λr+1 ...<br />

<br />

cr+1m <br />

<br />

<br />

. ..<br />

. <br />

..<br />

<br />

<br />

<br />

<br />

c1m c2m ... crm cr+1m ... cmm +<br />

<br />

λm<br />

∂r <br />

Gm <br />

<br />

∂λ1∂λ2 ...∂λr<br />

λ=0<br />

= A 12...r<br />

12...r (C).<br />

Simi<strong>la</strong>rlywehavefor1i1


For m = 3wehave<br />

S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 657<br />

G2(λ) = <strong>de</strong>t C2 + λ1A 1 1 (C2)<br />

2<br />

+ λ2 A2 (C2) + λ1A 12 <br />

12 (C2)<br />

= <strong>de</strong>t C2 + λ1A 1 {1}<br />

1 C2 λ + λ2A 2 {2}<br />

2 C2 λ .<br />

G3(λ) = <strong>de</strong>t C3 + λ1A 1 1 (C3) + λ2A 2 2 (C3) + λ3A 3 3 (C3) + λ1λ2A 12<br />

12 (C3) + λ1λ3A 13<br />

13 (C3)<br />

+ λ2λ3A 23<br />

23 (C3) + λ1λ2λ3A 123<br />

123 (C3)<br />

1<br />

= <strong>de</strong>t C2 + λ1 A1 (C3) + λ2A 12<br />

12 (C3) + λ3A 13<br />

13 (C3) + λ2λ3A 123 <br />

123 (C3)<br />

2<br />

+ λ2 A2 (C3) + λ3A 23<br />

23 (C3) + λ3A 3 3 (C3)<br />

= <strong>de</strong>t C3 + λ1A 1 [1]<br />

1 C3 λ + λ2A 2 [2]<br />

2 C3 λ + λ3A 3 [3]<br />

3 C3 λ ,<br />

G3(λ) = <strong>de</strong>t C3 + λ1A 1 1 (C3)<br />

2<br />

+ λ2 A2 (C3) + λ1A 12 <br />

12 (C3)<br />

1<br />

+ λ3 A1 (C3) + λ1A 12<br />

13 (C3) + λ2A 23<br />

23 (C3) + λ1λ2A 123 <br />

123 (C3)<br />

= <strong>de</strong>t C3 + λ1A 1 {1}<br />

1 C3 λ + λ2A 2 {2}<br />

2 C3 λ + λ3A 3 {3}<br />

3 C3 λ .<br />

For m>3 the proof of (28) and (30) is the same. The i<strong>de</strong>ntity (29) follows from (28) and (31)<br />

follows from (30). ✷<br />

The proof of Lemma 16 is based on Lemmas A.4, A.6 and A.7 concerning the properties of a<br />

positive matrices.<br />

Lemma A.4. (Sylvester [10, Chapter II, Section 3]) Let C ∈ Mat(n, R) and 1 p


658 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

where M(α) = Mα α (C), A(α) = Aαα (C) and ˆα ={1,...,m}\α.<br />

More precisely, see [12, p. 573]; [13, Chapter 2.5, Prob<strong>le</strong>m 36]. See also [27, Corol<strong>la</strong>ry 3.2,<br />

p. 34].<br />

Let us set as before (see (25)) for λ = (λ1,...,λk) ∈ C k and C ∈ Mat(k, C)<br />

Gk(λ) = <strong>de</strong>t Ck(λ), where Ck(λ) = C +<br />

In the following <strong>le</strong>mma we use the notation for λ = (λ1,...,λk) ∈ C k :<br />

k<br />

r=1<br />

λrErr.<br />

λ ]l[ = (λ1,...,λl−1, 0,λl+1,...,λk), 1 l k,<br />

and Gl(λ) = M 12...l<br />

12...l (Ck(λ)), 1 l k.Forα and β such that ∅⊆α ⊆{1, 2,...,l} and ∅⊆β ⊆<br />

{l + 1,...,k}, with l


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 659<br />

Simi<strong>la</strong>rly, we have<br />

]p[<br />

Gl(λ) = Gl λ ]p[<br />

+ λpGl λ p (35) ]p[<br />

= Gk λ p<br />

l+1...k ]p[<br />

+ λpGk λ l+1...k pl+1...k , 1 p l.<br />

pl+1...k<br />

Finally we get (36). Using the following formu<strong>la</strong>:<br />

′<br />

a + bx<br />

=<br />

c + dx<br />

bc − ad<br />

(c + dx) 2<br />

we conclu<strong>de</strong> that (36) implies the i<strong>de</strong>ntity in (37).<br />

To prove the inequality in (36) we get<br />

<br />

Gk(λ<br />

<br />

<br />

]p[ ) p p Gk(λ ]p[ )<br />

Gk(λ ]p[ ) pl+1...k<br />

pl+1...k Gk(λ ]p[ ) l+1...k<br />

<br />

<br />

<br />

<br />

l+1...k<br />

=<br />

<br />

<br />

<br />

<br />

A pl+1...k<br />

<br />

A<br />

= <br />

<br />

α α (C) Aα∩β<br />

A α∪β<br />

α∪β (C) Aβ<br />

β (C)<br />

where C = Ck(λ ]p[ ), α ={p} and β ={l + 1,l+ 2,...,k}. ✷<br />

A p p(Ck(λ ]p[ )) A∅ ∅ (Ck(λ ]p[ ))<br />

pl+1...k (Ck(λ ]p[ )) A l+1...k<br />

l+1...k (Ck(λ ]p[ ))<br />

α∩β (C)<br />

<br />

<br />

<br />

<br />

(34)<br />

0,<br />

Appendix B. Calcu<strong>la</strong>tion of the matrix e<strong>le</strong>ments φp(t) for t ∈ R p , their generalizations<br />

and Ξ pq<br />

n<br />

Let us recall (see (10) and (19)) that ˆλk = k−1<br />

r=1 crr,2 r m, ˆλ1 = 0 and<br />

To estimate<br />

<br />

n = max<br />

pq<br />

Mξn (t) 2 , 1 p q m. (39)<br />

Ξ pq<br />

<br />

max<br />

t∈R p<br />

t∈Rp Mξ pq<br />

n (t) 2 <br />

= max<br />

ξ pq<br />

n (t)1, 1 2 ,<br />

t∈R p<br />

where ξ pq<br />

n (t) = iyqn exp( p<br />

r=1 tr Arn) we shall find the exact formu<strong>la</strong>s for the matrix e<strong>le</strong>ments<br />

φp(t) = φ (n)<br />

p (t) = T R,μm B<br />

exp( p<br />

r=1 tr Ern) 1, 1 , t = (tr) p<br />

r=1 ∈ Rp , 1 p m, (40)<br />

of the restriction of the representation T R,μm B on the commutative subgroup (exp( p r=1 trErn) |<br />

t ∈ Rp ) Rp of the group BN 0 and theirs generalization <strong>de</strong>fined below. We note that<br />

exp( p r=1 trErn) = I + p r=1 trErn.<br />

For 1 p q, p,q ∈ N we get<br />

ξ pq<br />

p<br />

n (t) = iyqn exp<br />

r=1<br />

tr Arn<br />

p<br />

= iyqn exp i<br />

r=1<br />

tr<br />

r−1<br />

<br />

k=1<br />

xkrykn + yrn<br />

<br />

<br />

<br />

<br />

<br />

; (41)<br />

we have used the expression Arn = r−1<br />

k=1 xkrykn + yrn = r k=1 xkrykn (see (9)). We have


660 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

T R,μm B<br />

exp( p r=1 tr<br />

= exp<br />

Ern)<br />

= exp i<br />

p<br />

r=1<br />

tr Arn<br />

p p<br />

k=1<br />

To obtain ξ pp (t) we generalize the function<br />

r=k<br />

p<br />

= exp i<br />

xkrtr<br />

<br />

T R,μm B<br />

exp( p<br />

r=1 tr Ern)<br />

ykn<br />

r=1<br />

<br />

.<br />

tr<br />

r<br />

k=1<br />

xkrykn<br />

in the following way. We rep<strong>la</strong>ce in the <strong>la</strong>tter i<strong>de</strong>ntity the vectors (tr,...,tr) ∈ R p−k+1 by<br />

(trk) p<br />

r=k ∈ Rp−k+1 and <strong>de</strong>note the result by ξpp(t):<br />

⎛<br />

⎜<br />

ξpp(t) = ξpp ⎝<br />

t11<br />

t21 t22<br />

t31 t32 ...<br />

tp1 tp2 ... tpp<br />

⎞<br />

p p<br />

⎟<br />

⎠ := exp i<br />

k=1<br />

r=k+1<br />

<br />

xkrtrk + tkk<br />

To obtain ξ pq (t) we consi<strong>de</strong>r the function ξpq(t; tqq) = ξpp(t) exp(itqqyqn). Wehave<br />

Finally we have<br />

ξ pp (t) = ∂ξpp(t)<br />

∂tpp<br />

⎛<br />

⎜<br />

ξpq(t; tqq) = ξpq ⎝<br />

<br />

<br />

<br />

:= exp i<br />

tkr=tk, 1rkp<br />

t11<br />

t21 t22<br />

t31 t32 ...<br />

tp1 tp2 ... tpp; tqq<br />

p p<br />

k=1<br />

r=k+1<br />

xkrtrk + tkk<br />

<br />

⎞<br />

⎟<br />

⎠<br />

ykn + tqqyqn<br />

and ξ pq (t) = ∂ξpq(t;<br />

<br />

tqq) <br />

<br />

∂tqq<br />

<br />

.<br />

<br />

ykn<br />

<br />

. (42)<br />

tqq=0,tkr=tk, 1rkp<br />

Let us <strong>de</strong>fine φp(t) = ξpp(t)dμ(x,y), φpq(t) = ξpq(t)dμ(x,y), where μ(x,y) = μI (x) ⊗<br />

( ∞ n=m+1 μ C (n)(y)) and μI (x) is the standard Gaussian mea<strong>sur</strong>e in R × R 2 ×···×R m .<br />

Using <strong>de</strong>finition (39) and the previous equalities we have finally<br />

Ξ pp <br />

<br />

= max<br />

∂φp(t) <br />

<br />

<br />

t∈R p<br />

∂tpp<br />

Ξ pq <br />

<br />

= max<br />

∂φpq(t) <br />

<br />

<br />

t∈R p<br />

∂tqq<br />

2<br />

2<br />

tkr=tk, 1rkp<br />

tqq=0,tkr=tk, 1rkp<br />

,<br />

. (43)<br />

Our aim is to estimate Ξ pq . We shall use the notation Ck := C{1,2,...,k} for Mat(m, C) and 1 <br />

k m (see notation Cα for ∅⊆α ⊆{1,...,m} in Lemma B of Section 3).<br />

.


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 661<br />

Lemma B.1. For 1 p q m and φpq(t) = ξpq(t)dμ(x,y) we have<br />

φpq<br />

=<br />

=<br />

⎛<br />

⎜<br />

⎝<br />

t11<br />

t21 t22<br />

t31 t32 t33<br />

...<br />

tp1 tp2 tp3 ... tpp; tqq<br />

<br />

R (p−1)(p−2)<br />

2 +p<br />

exp i<br />

1<br />

√ <strong>de</strong>t C1(t) exp<br />

<br />

p p<br />

k=1<br />

r=k<br />

⎞<br />

⎟<br />

⎠<br />

xkrtrk<br />

<br />

ykn + tqqyqn<br />

<br />

dμ(x,y)<br />

− 1<br />

<br />

(CT , T ) − C1(t)<br />

2<br />

−1 d,d <br />

, (44)<br />

where we set T = (t11,t22,t33,...,tpp; tqq) ∈ R p+1 , C ∈ Mat(p + 1, C) is <strong>de</strong>fined by<br />

⎛<br />

c11 c12 c13 ... c1p c1q<br />

⎜ c12 c22 c23 ... c2p c2q<br />

⎜ c13 c23 c33 ... c3p c3q<br />

C := Cp,q := C{1,2,...,p,q} := ⎜<br />

.<br />

⎜<br />

..<br />

⎝<br />

c1p c2p c3p ... cpp cpq<br />

c1q c2q c3q ... cpq cqq<br />

d = d21(t), d31(t), . . . , dp1(t); d32(t), d42(t),...,dp2(t); ...; dpp−1(t) ∈ R (p−1)(p−2)<br />

2 ,<br />

drs(t) = trses(t), 1 s


662 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

where D(t) = diag(t21,...,tp1; t32,...,tp2; t43,...,tp3; ...; tpp−1). We have<br />

<strong>de</strong>t C1(t) = 1 +<br />

p<br />

<br />

r=1 1i1


hence<br />

S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 663<br />

Using the <strong>la</strong>tter inequality we get (see (48))<br />

M 12...p−1 12...p<br />

12...p−1 (Cp)M12...p (Cp) +<br />

= A p p(Cp)A ∅ ∅ (Cp) +<br />

A p<br />

k (Cp) 2


664 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Fourier transform for the Gaussian mea<strong>sur</strong>e μC in the space Rm is:<br />

<br />

1<br />

√<br />

(2π) m<br />

Rm <br />

exp i(y,x)dμC(x) = exp − 1<br />

2 (Cy,y)<br />

<br />

, y ∈ R m .<br />

Let p = 1. Using (55)–(57) we have<br />

<br />

φ1(t11) =<br />

R<br />

<br />

exp(it11y1n)dμ(y)= exp − 1<br />

2 c11t 2 <br />

11 ;<br />

<br />

φ1q(t11; tqq) =<br />

R2 <br />

exp i(t11y1n + tqqyqn)dμ(y)= exp − 1<br />

c11t<br />

2<br />

2 11 + 2c1qt11tqq + cqqt 2 <br />

qq<br />

<br />

;<br />

Mξ 1q <br />

(t11) = iyqn exp(it11y1n)dμ(y)=<br />

R<br />

∂φ1q(t11;<br />

<br />

<br />

tqq) <br />

<br />

∂tqq<br />

=−c1qt11 exp −<br />

tqq=0<br />

1<br />

2 c2 11t2 <br />

11 ,<br />

<br />

Mξ 1,q (t11) 2 = c 2 1qt2 11 exp−c11t 2 <br />

11 ;<br />

Ξ 1q <br />

= maxMξ<br />

1q (t11) 2 = c2 1q exp(−1)<br />

= Ψ 1q ,<br />

we have used the obvious result<br />

t11∈R<br />

c11<br />

<br />

1<br />

max f(x)= f =<br />

x∈R a<br />

1<br />

, where f(x)= x exp(−ax), a > 0. (59)<br />

ea<br />

This proves (47) for (p, q) = (1,q).<br />

To prove (44) in the general case we note that<br />

where<br />

p<br />

p<br />

k=1<br />

r=k+1<br />

xkrtrk + tkk<br />

<br />

ykn + tqqyqn = a(x) + T,y <br />

R p+1,<br />

y = (y1n,y2n,...,ypn; yqn), T = (t11,t22,...,tpp; tqq) ∈ R p+1 ,<br />

a(x) = a1(x), a2(x), . . . , ap(x); 0 ∈ R p+1 , ak(x) =<br />

x = <br />

1


φpq(t; tqq) =<br />

Since<br />

<br />

=<br />

S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 665<br />

R p+1<br />

exp i<br />

p p<br />

k=1<br />

r=k<br />

xkrtrk<br />

<br />

ykn + tqqyqn<br />

<br />

dμ(x,y)<br />

exp i a(x) + T,y <br />

dμ(x,y) = exp − 1<br />

<br />

C a(x) + T ,a(x)+ T<br />

2<br />

<br />

dμI (x).<br />

C a(x) + T ,a(x)+ T = Ca(x), a(x) + 2 a(x),CT + (CT , T ),<br />

we have<br />

<br />

φpq(t; tqq) = exp − 1<br />

<br />

(CT , T )<br />

2<br />

<br />

exp − 1<br />

<br />

Ca(x), a(x) + 2 a(x),CT<br />

2<br />

<br />

To calcu<strong>la</strong>te the <strong>la</strong>tter integral we use (56). Let us intro<strong>du</strong>ce the notation<br />

We show that<br />

for some<br />

We have<br />

where<br />

Further<br />

a(x),CT =<br />

d(t) = drk(t) <br />

X = (x12; x13,x23; ...; x1p,...; xp−1p) ∈ R (p−1)(p−2)<br />

2 .<br />

Ca(x), a(x) + 2 a(x),CT = C(t)X,X + 2 d(t),X <br />

d(t) ∈ R (p−1)(p−2)<br />

<br />

(p − 1)(p − 2)<br />

2 and C(t) ∈ Mat<br />

, R .<br />

2<br />

p<br />

ak(x)(CT )k =<br />

k=1<br />

= <br />

1k


666 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Ca(x), a(x) = <br />

1k,np<br />

= <br />

cknak(x)an(x) = <br />

<br />

1k


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 667<br />

Let e1(t) = c11t11 + c12t22 = 0sot11 =−c12t22/c11. In this case<br />

Finally<br />

(CT , T ) = c11t 2 11 + 2c12t11t22 + c22t 2 22 =<br />

<br />

c2 12<br />

− 2<br />

c11<br />

c2 <br />

12<br />

+ c22 t<br />

c11<br />

2 22<br />

<br />

c12t11 + c22t22 = − c12c12<br />

<br />

+ c22 t22 =<br />

c11<br />

c11c22 − c2 12<br />

c11<br />

M12<br />

12<br />

=<br />

c11<br />

= M12<br />

12<br />

c11<br />

<br />

Mξ 22 (t) 2 = Miy2nexp it11 + it22(x12y1n + y2n) 2 <br />

<br />

= <br />

∂φ2(t) <br />

<br />

<br />

=<br />

M 12<br />

12<br />

c11 t22<br />

We have used the inequality<br />

2 M<br />

exp − 12<br />

12<br />

c11 t2 <br />

22<br />

1 + c11t 2 22<br />

M12<br />

12<br />

c11<br />

Hence if we <strong>de</strong>note t = (t11,t22) ∈ R 2 we have using (43)<br />

∂t22<br />

2<br />

.<br />

t 2 22 ,<br />

e1(t)=0,t21=t22<br />

t 2 22 exp<br />

<br />

M12 <br />

12<br />

− + c11 t<br />

c11<br />

2 <br />

22 .<br />

1 + x exp x, x ∈ R. (62)<br />

Ξ 22 = max<br />

t∈R2 <br />

22<br />

Mξ (t) 2 22 (M<br />

>Ψ := 12<br />

12 )2 exp(−1)<br />

c11(M12 12 + c2 .<br />

11 )<br />

This proves (47) for (p, q) = (2, 2). For (2,q),2


668 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

<br />

<br />

∂φ2q(t; tqq)<br />

<br />

∂tqq<br />

<br />

tqq=0<br />

<br />

= −(c1qt11 + c2qt22 + cqqtqq) + (c11t11 + c12t22 + c1qtqq)c1qt 2 <br />

21<br />

∂φ2q(t; tqq)<br />

∂tqq<br />

<br />

tqq=0<br />

1 + c11t 2 21<br />

<br />

× exp − 1<br />

<br />

(CT , T ) − C1(t)<br />

2<br />

−1 d,d 1<br />

√<br />

<strong>de</strong>t C1(t) ,<br />

<br />

<br />

= −(c1qt11 + c2qt22) + (c11t11 + c12t22)c1qt 2 <br />

21<br />

1 + c11t 2 21<br />

<br />

× exp − 1<br />

<br />

1 <br />

(CT , T ) √ <br />

2 <strong>de</strong>t C1(t)<br />

tqq=0<br />

Let tqq = 0. We chose d(t) = 0sowehavec11t11 + c12t22 = 0 and t11 = −c12t22 . In this case<br />

c11<br />

(CT , T ) = c11t 2 11 + 2c12t11t22 + c22t 2 22 =<br />

<br />

c2 12<br />

− 2<br />

c11<br />

c2 <br />

12<br />

+ c22 t<br />

c11<br />

2 22<br />

<br />

c1qt11 + c2qt22 = − c12c1q<br />

<br />

+ c2q<br />

c11<br />

t22 = c11c2q − c12c1q<br />

c11<br />

Finally, if we <strong>de</strong>note t = (t11,t22) ∈ R2 ,wehave<br />

<br />

2q<br />

Mξ (t) 2 <br />

= Miyqn exp it11 + it22(x12y1n + y2n) 2 <br />

<br />

= <br />

<br />

=<br />

M 12<br />

1q<br />

c11 t22<br />

2 exp − M 12<br />

1 + c11t 2 22<br />

By (59) we conclu<strong>de</strong> using (43) that<br />

Ξ 2q <br />

= maxMξ<br />

2q (t) 2 <br />

<br />

max<br />

<br />

t∈R 2<br />

t22∈R<br />

12<br />

c11 t2 22<br />

This proves (47) for (p, q) = (2,q),2<br />

∂φ2q(t; tqq) 2<br />

∂tqq<br />

2 1q<br />

t22<br />

c11<br />

M 12<br />

<br />

<br />

<br />

tqq=0,e1(t)=0<br />

.<br />

M12<br />

12<br />

=<br />

c11<br />

t 2 22 ,<br />

t22 = M12<br />

1q<br />

t22.<br />

c11<br />

∂φ2q(t; tqq) 2<br />

∂tqq<br />

<br />

<br />

<br />

tqq=0,e1(t)=0<br />

<br />

M12 <br />

12<br />

exp − + c11 t<br />

c11<br />

2 <br />

22 .<br />

(M12<br />

1q )2 exp(−1)<br />

c11(M 12<br />

12 + c2 11 ) = Ψ 2q .<br />

<br />

1<br />

= √<br />

<strong>de</strong>t C1(t) exp<br />

<br />

− 1<br />

<br />

(CT , T ) − C1(t)<br />

2<br />

−1 d,d <br />

,<br />

T = (t11,t22,t33), d(t) = d21(t), d31(t), d32(t) ,<br />

d21(t) = t21e1(t), d31(t) = t31e1(t), d32(t) = t32e2(t),<br />

e1(t) = c11t11 + c12t22 + c13t33, e2(t) = c21t11 + c22t22 + c23t33,


hence<br />

C = C3 =<br />

S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 669<br />

c11 c12 c13<br />

c12 c22 c23<br />

c13 c23 c33<br />

⎛<br />

C1(t) = I + C(t) = ⎝<br />

= diag(t21,t31,t32)<br />

<br />

, C(t) (61)<br />

⎛<br />

= ⎝<br />

1 + c11t 2 21 c11t21t31 c12t21t32<br />

c11t21t31 1 + c11t 2 31 c12t31t32<br />

c11t 2 21 c11t21t31 c12t21t32<br />

c11t21t31 c11t 2 31 c12t31t32<br />

c12t21t32 c12t31t32 c22t 2 32<br />

⎞<br />

⎠<br />

c12t21t32 c12t31t32 1 + c22t 2 32<br />

⎛<br />

c11 + t<br />

⎝<br />

−2<br />

21<br />

c11 c12<br />

c11 c11 + t −2<br />

31<br />

c12<br />

c12 c12 c22 + t −2<br />

32<br />

⎞<br />

⎠ ,<br />

⎞<br />

⎠ diag(t21,t31,t32).<br />

We prove the following inequality for an operator C of or<strong>de</strong>r n such that I + C>0:<br />

<strong>de</strong>t(I + C) exp tr C. (63)<br />

In<strong>de</strong>ed by Hadamard inequality (see [7] or [13, Section 2.5.4]) we have for positive operator C<br />

of or<strong>de</strong>r n<br />

<strong>de</strong>t C <br />

Using the Hadamard inequality and (62) we have for an operator C such that I + C>0<br />

i=1<br />

n<br />

i=1<br />

cii.<br />

n<br />

<strong>de</strong>t(I + C) (1 + cii) (62) n<br />

<br />

n<br />

exp cii = exp<br />

i=1<br />

i=1<br />

cii<br />

<br />

= exp(tr C),<br />

where we <strong>de</strong>note by tr C the trace of an operator C in the space C n . Using (63) and (61) we<br />

conclu<strong>de</strong> that<br />

<strong>de</strong>t I + C(t) tr C(t) = exp<br />

p−1<br />

<br />

k=1<br />

where α2 k = p r=k+1 t2 rk since by (61) we have<br />

Using (26) we get<br />

tr C(t) = <br />

1k


670 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Finally we have<br />

<strong>de</strong>t C1(t) = 1 + c11α 2 1 + c22α 2 2<br />

= t 2 21t2 31t2 <br />

c11 c11 c12 <br />

<br />

32 c11 c11 c22 <br />

<br />

c12 c12 c22<br />

+<br />

<br />

1<br />

t2 +<br />

21<br />

1<br />

t2 <br />

c11 c22<br />

<br />

<br />

c12 c22<br />

<br />

31<br />

+ 1<br />

t2 21t2 <br />

1<br />

c22 +<br />

31 t2 21t2 +<br />

32<br />

1<br />

t2 31t2 <br />

1<br />

c11 +<br />

32 t2 21t2 31t2 <br />

32<br />

2<br />

= 1 + c11 t21 + t 2 <br />

31 + c22t 2 2<br />

32 + M12 12 t21 + t 2 2<br />

31 t32 .<br />

For general n we have by analogy (it proves thus (46))<br />

n−1<br />

<strong>de</strong>t C1(t) = 1 +<br />

<br />

r=1 1i1


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 671<br />

If we <strong>de</strong>note ek(t) = n r=1 ckrtrr we get<br />

(CT , T ) = <br />

1k,rn<br />

Un<strong>de</strong>r conditions (67) we have<br />

and<br />

en(t) =<br />

n<br />

r=1<br />

ckrtrrtkk =<br />

n<br />

ek(t)tkk,<br />

k=1<br />

A<br />

cnr<br />

n r (Cn)<br />

An n (Cn) tnn = M12...n<br />

(CT , T ) (69)<br />

=<br />

n<br />

k=1<br />

For n = 3 using (70) and (71) we can calcu<strong>la</strong>te<br />

12...n (Cn)<br />

M 12...n−1<br />

tnn,<br />

12...n−1 (Cn)<br />

e3(t) = M123<br />

123 (C3)<br />

M12 12 (C3) t33, (CT,T)= M123<br />

123 (C3)<br />

M12 12 (C3) t2 33 ,<br />

If, in addition, e1(t) = e2(t) = 0, we have (see (66))<br />

2<br />

tr C(t) = c11 t22 + t 2 <br />

33 + c22t 2 33 =<br />

<br />

M12 c11<br />

1 ∂(CT,T)<br />

= en(t). (69)<br />

2 ∂tnn<br />

∂(C1(t) −1d(t), d(t))<br />

= 0 (70)<br />

∂tnn<br />

ek(t)tkk = en(t)tnn = M12...n<br />

12...n (Cn)<br />

M 12...n−1<br />

12...n−1 (Cn) t2 nn . (71)<br />

13 (C3)<br />

M 12<br />

23 (C3)<br />

∂(C1(t) −1d(t), d(t))<br />

= 0.<br />

∂t33<br />

2 <br />

+ 1 + c22 t 2 33 .<br />

For n = 3wehaveife1(t) = e2(t) = 0, using the values for t22, e3(t) and (CT , T )<br />

<br />

<br />

<br />

∂φ3(t) 2<br />

<br />

∂t33<br />

= e2 3 (t) exp(−(CT , T )) (64)<br />

e<br />

<strong>de</strong>t C1(t)<br />

2 3 (t) exp−(CT , T ) − tr C(t) <br />

<br />

M123 123<br />

=<br />

(C3)<br />

M12 12 (C3)<br />

2 t 2 33 exp<br />

<br />

−t 2 <br />

M123 123<br />

33<br />

(C3)<br />

M12 12 (C3) + (c11<br />

<br />

M12 + c22) + c11<br />

We get by (59)<br />

<br />

M123 max<br />

t33∈R<br />

=<br />

=<br />

123 (C3)<br />

M 12<br />

12 (C3)<br />

2 t 2 33 exp<br />

<br />

−t 2 <br />

M123 123<br />

33<br />

(C3)<br />

M12 12 (C3) + (c11 + c22) + c11<br />

M 123<br />

123 (C3)<br />

M 12<br />

12 (C3)<br />

2 exp(−1)<br />

M 123<br />

123 (C3)<br />

M 12<br />

12 (C3) + (c11 + c22) + c11<br />

M 12<br />

13 (C3)<br />

M 12<br />

12 (C3)<br />

2<br />

13 (C3)<br />

M 12<br />

12 (C3)<br />

<br />

M12 13 (C3)<br />

M12 12 (C3)<br />

2 (M123 123 (C3)) 2 exp(−1)<br />

M12 12 (C3)M123 123 (C3) + c11(M12 13 (C3)) 2 + (c11 + c22)(M12 12 (C3)) 2 = Ψ 33 .<br />

Finally we have (see (43))<br />

2<br />

.


672 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Ξ 33 <br />

<br />

= max<br />

33<br />

Mξ (t) 2 <br />

max<br />

∂φ3(t) <br />

<br />

<br />

t∈R 2<br />

This proves (47) for (p, q) = (3, 3).<br />

By analogy we have for general n:<br />

<br />

= − 1<br />

t33∈R<br />

∂t33<br />

2<br />

e1(t)=e2(t)=0<br />

Ψ 33 .<br />

+ ∂(C1(t) −1 1<br />

d(t), d(t)) exp(− 2 [(CT , T ) − (C1(t)<br />

∂tnn<br />

−1d(t), d(t))])<br />

√ ,<br />

<strong>de</strong>t C1(t)<br />

∂φn(t) ∂(CT,T)<br />

∂tnn 2 ∂tnn<br />

<br />

∂φn(t)<br />

= −en(t) +<br />

∂tnn<br />

∂(C1(t) −1 1<br />

d(t), d(t)) exp(− 2 [(CT , T ) − (C1(t)<br />

∂tnn<br />

−1d(t), d(t))])<br />

√ .<br />

<strong>de</strong>t C1(t)<br />

When trk = trr, n r k 2, we have by (65)<br />

tr C(t) = <br />

1k


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 673<br />

where C = Cn,q and T are <strong>de</strong>fined in Lemma B.1. Moreover, the above conditions gives us the<br />

same solutions (68) as before, hence using the <strong>de</strong>composition of the minor M 12...n−1n<br />

12...n−1q (Cn,q) we<br />

have<br />

eq(t) = (Cn,qT)q =<br />

n<br />

r=1<br />

cqrtrr =<br />

n<br />

r=1<br />

Finally we get if e1(t) =···=en−1(t) = 0 and tqq = 0<br />

Ξ nq <br />

<br />

max<br />

<br />

tnn∈R<br />

∂φnq(t; tqq) 2<br />

∂tqq<br />

<br />

<br />

<br />

tqq=0<br />

12...n−1n<br />

M<br />

= max<br />

tnn∈R M 12...n−1<br />

12...n−1 (Cn)<br />

=<br />

12...n−1q (Cn,q)<br />

A<br />

cqr<br />

n r (Cn)<br />

An n (Cn) tnn = M12...n−1n<br />

An n (Cn)<br />

max<br />

tnn∈R e2 q (t) exp −(CT , T ) − tr C(t) <br />

2 t 2 nn exp<br />

<br />

−t 2 <br />

M12...n 12...n<br />

nn<br />

(Cn)<br />

(M 12...n−1n<br />

12...n−1q (Cn,q)) 2 exp(−1)<br />

M 12...n−1<br />

12...n−1<br />

12...n−1q (Cn,q)tnn<br />

nk=2 ˆλk(A<br />

+<br />

(Cn) n k<br />

(An n (Cn)) 2<br />

M 12...n−1<br />

12...n−1 (Cn)M12...n 12...n (Cn) + n ˆλk(A k=2 n k (Cn)) 2 = Ψ nq . ✷<br />

Appendix C. Proof of Lemma 16<br />

.<br />

(Cn)) 2<br />

Proof. Firstly, we prove by in<strong>du</strong>ction the inequalities I k k 0fork2. Secondly, we show that<br />

inequality I k k 0 and Lemma A.6 imply the inequality I k m 0formk where (see (24)):<br />

I k m := fkA k k Cm(ˆλ) − ˆλkA k <br />

k Cm ˆλ [k] 0, 2 k m.<br />

We shall show also that I 2 m<br />

= 0. In the case m = 2wehave<br />

I 2 2 = f2A 2 2 C2(ˆλ) − ˆλ2A 2 <br />

2 C2 ˆλ [2] = 0<br />

since f2 = ˆλ2 = c11 by (19), (20) and (49), and<br />

A 2 2 C2(ˆλ) = A 2 <br />

2 C2 ˆλ [2] = A 2 2 (C2) = c11,<br />

where<br />

<br />

c11<br />

C2(ˆλ) =<br />

c12<br />

c12 c11 + c22<br />

<br />

<br />

, C2ˆλ<br />

[2] = C2 =<br />

c11 c12<br />

c12 c22<br />

In the case m = 3 we prove the following inequalities:<br />

I 2 3 := f2A 2 2 C3(ˆλ) − ˆλ2A 2 <br />

2 C3 ˆλ [2] 0, (72)<br />

I 3 3 := f3A 3 3 C3(ˆλ) − ˆλ3A 3 <br />

3 C3 ˆλ [3] 0. (73)<br />

Since (see (21))<br />

<br />

.


674 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

C3(ˆλ) =<br />

c11 c12 c13<br />

c12 c11 + c22 c23<br />

c13 c23 c11 + c22 + c33<br />

<br />

<br />

, C3ˆλ<br />

[2] =<br />

c11 c12 c13<br />

c12 c22 c23<br />

c13 c23 c11 + c22 + c33<br />

and C3(ˆλ [3] ) = C3 we have by (26)<br />

A 2 2 C3(ˆλ) = A 2 <br />

2 C3 ˆλ [2] = A 2 2 (C3) + ˆλ3A 23<br />

23 (C3), A 3 <br />

3 C3 ˆλ [3] = A 3 3 (C3).<br />

The <strong>la</strong>tter equalities give us I 2 3 = 0. This proves (72). In<strong>de</strong>ed we have<br />

I 2 3 = <br />

ˆλ2<br />

2<br />

A2 (C3) + ˆλ3A 23<br />

23 (C3) <br />

− ˆλ2<br />

2<br />

A2 (C3) + ˆλ3A 23<br />

23 (C3) ≡ 0.<br />

Since f2 = ˆλ2 = c11 and ˆλ1 = 0wehaveA2 2 (Cm(ˆλ)) = A2 2 (Cm(ˆλ [2] )) hence<br />

I 2 m := f2A 2 2 Cm(ˆλ) − ˆλ2A 2 <br />

2 Cm ˆλ [2] ≡ 0, 2 m. (74)<br />

Remark C.1. In what follows we take λ = (λr) k 1 ∈ Rk with λ1 = 0.<br />

To prove (73) for k = 3 we use the i<strong>de</strong>ntity for λ = (0,λ2) ∈ R 2 , ˆλ = (0,c11),<br />

A 3 3<br />

A 3 3<br />

<br />

C3(ˆλ) = M 12<br />

12 C3(ˆλ) = M 12<br />

12 (C3) + c 2 11<br />

<br />

C3(λ) = M 12<br />

12 C3(λ) = M 12<br />

12 (C3) + λ2c11,<br />

= M12<br />

12 (C3) + ˆλ2c11,<br />

∂M 12<br />

12 (C3(λ))<br />

∂λ2<br />

= c11.<br />

We have<br />

I 3 3 := f3A 3 3 C3(ˆλ) − ˆλ3A 3 <br />

3 C3 ˆλ [3]<br />

<br />

= c11 + c2 12 (M<br />

+<br />

c11<br />

12<br />

12 (C3)) 2<br />

c11(M12 12 (C3) + c2 11 )<br />

<br />

M12 12 (C3) + c 2 <br />

11 − (c11 + c22)M 12<br />

12 (C3)<br />

<br />

= c11 + c2 12<br />

+<br />

c11<br />

(M12<br />

12 (C3)) 2<br />

c11M 12<br />

12 (C3(ˆλ))<br />

<br />

M 12<br />

12 C3(ˆλ) − (c11 + c22)M 12<br />

12 (C3),<br />

we use here the <strong>de</strong>finition of fq = e <br />

1rp


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 675<br />

Since I 3 3 = I 3 3 (ˆλ) it is sufficient to prove that I 3 3 (λ) > 0forλ2 > 0. We show that<br />

I 3 3 (0) = 0 and<br />

In<strong>de</strong>ed we have M 12<br />

12 (C3(0)) = M 12<br />

12 (C3) hence<br />

I 3 <br />

3 (0) = c11 + c2 12<br />

c11<br />

= M 12<br />

12 (C3)<br />

+ M12<br />

12 (C3)<br />

c11<br />

c 2 12 + M 12<br />

∂I3 3 (λ)<br />

<br />

=<br />

∂λ2<br />

∂I3 3 (λ)<br />

> 0.<br />

∂λ2<br />

<br />

M 12<br />

12 (C3) − (c11 + c22)M 12<br />

12 (C3)<br />

12 (C3)<br />

− c22<br />

c11<br />

c11 + c2 12<br />

c11<br />

<br />

= 0 and<br />

<br />

c11 > 0.<br />

Finally I 3 3 (λ) > 0forλ2 > 0soI 3 3 = I 3 3 (ˆλ) = I 3 3 (0,c11)>0 and (73) is proved. To prove that<br />

I k k 0 <strong>le</strong>t us <strong>de</strong>note f q = e q−1<br />

r=1 Ψ rq−1 . Using (20) we have<br />

fq = e <br />

1rpI4 4 (λ)| λ=ˆλ where I 4 4 (λ) is <strong>de</strong>fined by the formu<strong>la</strong><br />

r=1


676 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

I 4 <br />

(c11 + c22)M<br />

4 (λ) :=<br />

12<br />

12 (C4)<br />

where<br />

M12 12 (C4(λ))<br />

+ c2 13<br />

c11<br />

+ (M12<br />

13 (C4)) 2<br />

c11M 12<br />

12<br />

(C4(λ)) +<br />

M 12<br />

12<br />

(M123 123 (C4)) 2<br />

(C4)M 123<br />

123 (C4(λ))<br />

× M 123<br />

123 C4(λ) − (c11 + c22 + c33)M 123<br />

123 (C4)<br />

<br />

a2<br />

= a1 +<br />

M12 12 (C4(λ))<br />

<br />

M 123<br />

123 C4(λ) + b1 = a1M 123<br />

123 C4(λ) M<br />

+ a2<br />

123<br />

123 (C4(λ))<br />

M12 + b1,<br />

12 (C4(λ))<br />

a1 = c2 13<br />

> 0, a2 = (c11 + c22)M<br />

c11<br />

12<br />

12 (C4) + (M12<br />

13<br />

(C4)) 2<br />

c11<br />

b1 = (M123<br />

123 (C4)) 2<br />

M12 12 (C4) − (c11 + c22 + c33)M 123<br />

123 (C4).<br />

We prove that I 4 4 (λ) 0forλ = (0,λ2,λ3), when λ2 0, λ3 0. It then gives us I 4 4 <br />

I 4 4 (ˆλ) 0. We have (see below the proof of I k k (0) = 0, k 3)<br />

I 4 <br />

c2 13<br />

4 (0) =<br />

c11<br />

+ (M12<br />

13 (C4)) 2<br />

c11M 12<br />

12<br />

12<br />

> 0,<br />

M123<br />

123<br />

+<br />

(C4) (C4)<br />

M12 <br />

− c33 M<br />

(C4) 123<br />

123 (C4) = 0.<br />

Moreover, by inequality (37) of Lemma A.7 we have for λ2 0, λ3 0<br />

Let us consi<strong>de</strong>r the function<br />

We have<br />

∂I4 4 (λ) ∂M<br />

= a1<br />

∂λ2<br />

123<br />

123 (C4(λ)) ∂ M<br />

+ a2<br />

∂λ2 ∂λ2<br />

123<br />

123 (C4(λ))<br />

M12 0,<br />

12 (C4(λ))<br />

∂I4 4 (λ)<br />

<br />

a2<br />

= a1 +<br />

∂λ3 M12 12 (C4(λ))<br />

<br />

∂M123 123 (C4(λ))<br />

0.<br />

∂λ3<br />

i 4 4 (t) = I 4 4 (t ˆλ) = I 4 4 (0,tˆλ2,tˆλ3), t ∈ R.<br />

i 4 4 (0) = I 4 4 (0) = 0 and<br />

di4 4 (t)<br />

dt = ∂I4 4 (λ)<br />

∂λ2<br />

hence i4 4 (t) 0 by the previous inequalities for t>0. So<br />

To prove that I k k (ˆλ) 0 we show that<br />

I 4 4 >I4 4 (0, ˆλ2, ˆλ3) = i 4 4 (t) t=1 0.<br />

I k k (0) = 0, 2 k and<br />

ˆλ2 + ∂I4 4 (λ)<br />

ˆλ3 0<br />

∂λ3<br />

∂Ik k (λ)<br />

0, 2 p


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 677<br />

To <strong>de</strong>fine the function I k+1<br />

k+1 k+1<br />

k+1 (λ) with Ik+1 Ik+1 (ˆλ) we have<br />

I k+1<br />

k+1<br />

k+1<br />

= fk+1Ak+1 Ck+1(ˆλ) − ˆλk+1A k+1<br />

k+1 (Ck+1)<br />

<br />

= fk + f k+1 A k+1<br />

Ck+1(λ) − ˆλk+1A k+1<br />

(75)<br />

(76)<br />

<br />

(54)<br />

<br />

k+1<br />

ˆλkA kk+1<br />

kk+1 (Ck+1)<br />

A kk+1<br />

kk+1<br />

(Ck+1(λ)) + e<br />

ˆλkA kk+1<br />

kk+1 (Ck+1)<br />

A kk+1<br />

kk+1<br />

:= I k+1<br />

k+1 (ˆλ),<br />

(Ck+1(λ)) + e<br />

k<br />

r=1<br />

k<br />

r=1<br />

Ψ rk<br />

Ψ rk<br />

0<br />

<br />

<br />

A k+1<br />

k+1<br />

A k+1<br />

k+1<br />

k+1 (Ck+1) λ=ˆλ<br />

Ck+1(λ) − ˆλk+1A k+1<br />

Ck+1(λ) − ˆλk+1A k+1<br />

where the function I k+1<br />

pq<br />

k+1 (λ) is <strong>de</strong>fined by (see <strong>de</strong>finition (54) of Ψ0 ):<br />

r=2<br />

<br />

<br />

<br />

k+1 (Ck+1) <br />

<br />

λ=ˆλ<br />

<br />

<br />

<br />

k+1 (Ck+1) <br />

<br />

λ=ˆλ<br />

I k+1<br />

<br />

ˆλkM<br />

k+1 (λ) =<br />

12...k−1<br />

12...k−1 (Ck+1)<br />

M 12...k−1<br />

12...k−1 (Ck+1(λ)) + c2 k (M 1k<br />

+<br />

c11<br />

r=2<br />

12...r−1r<br />

12...r−1k<br />

(Ck+1))<br />

2<br />

M 12...r−1<br />

12...r−1<br />

(Ck+1)M12...r 12...r (Ck+1(λ))<br />

<br />

× M 12...k<br />

12...k Ck+1(λ) − ˆλk+1M 12...k<br />

12...k (Ck+1)<br />

<br />

ˆλkM<br />

=<br />

12...k−1<br />

12...k−1 (Ck+1)<br />

M 12...k−1<br />

12...k−1 (Ck+1(λ)) + c2 k−1<br />

(M 1k<br />

+<br />

c11<br />

12...r−1r<br />

12...r−1k<br />

(Ck+1))<br />

2<br />

M 12...r−1<br />

12...r−1<br />

(Ck+1)M12...r 12...r (Ck+1(λ))<br />

<br />

× M 12...k<br />

12...k Ck+1(λ) + (M12...k<br />

12...k<br />

(Ck+1))<br />

2<br />

M 12...k−1<br />

12...k−1 (Ck+1) − ˆλk+1M 12...k<br />

12...k (Ck+1).<br />

Finally we have the following expression for I k+1<br />

k+1 (λ) with corresponding positive constants ar,<br />

2 r k − 1 (<strong>de</strong>pending on k) and b1 ∈ R:<br />

I k+1<br />

k+1 (λ) =<br />

=<br />

<br />

<br />

k−1<br />

a1 +<br />

r=2<br />

k−1<br />

ar<br />

a1 +<br />

Gr(λ)<br />

r=2<br />

ar<br />

M12...r 12...r (Ck+1(λ))<br />

<br />

<br />

Gk(λ) + b1.<br />

M 12...k<br />

12...k<br />

By (37) of Lemma A.7 we conclu<strong>de</strong> that for λr 0, 2 r k, holds<br />

∂I k+1<br />

k+1 (λ)<br />

∂λp<br />

∂I k+1<br />

k+1 (λ)<br />

∂λk<br />

=<br />

<br />

∂Gk(λ)<br />

= a1<br />

∂λp<br />

k−1<br />

a1 +<br />

r=2<br />

k−1<br />

+<br />

r=2<br />

ar<br />

<br />

ar ∂Gk(λ)<br />

0,<br />

Gr(λ) ∂λk<br />

∂<br />

∂λp<br />

Ck+1(λ) + b1<br />

Gk(λ)<br />

0, 2 p k. (78)<br />

Gr(λ)


678 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Remark C.2. In fact ∂I k+1<br />

k+1 (λ)/∂λp > 0, 2 p k, forλ = (λr) k+1<br />

r=1 ∈ Rk+1 , λr 0, 1 r <br />

k + 1, since by (38) we have ∂Gk(λ)/∂λp = A p p(C(λ ]p[ )) > 0.<br />

Let us suppose that I k k (0) = 0, i.e.<br />

0 = I k k (0) = M12...k−1<br />

12...k−1<br />

For k = 3, k = 4 and k = 5wehave<br />

c 2 1k−1<br />

c11<br />

+ (M12...k−3k−2<br />

12...k−3k−1 )2<br />

M 12...k−3<br />

12...k−3 M12...k−2<br />

12...k−2<br />

I 3 3 (0) = M12 12<br />

I 4 4 (0) = M123 123<br />

I 5 5 (0) = M1234 1234<br />

c 2 14<br />

c11<br />

c 2 13<br />

c 2 12<br />

c11<br />

c11<br />

+ (M12<br />

14 )2<br />

c11M 12<br />

12<br />

We prove that I k+1<br />

k+1 (0) = 0. In<strong>de</strong>ed, we get<br />

I k+1<br />

<br />

c2 1k<br />

k+1 (0) = M12...k 12...k<br />

c11<br />

+ (M12<br />

1k−1 )2<br />

c11M 12<br />

12<br />

+ (M123<br />

12k−1 )2<br />

M 12<br />

12 M123<br />

123<br />

+ M12...k−1<br />

12...k−1<br />

M 12...k−2<br />

− ck−1k−1<br />

12...k−2<br />

+ M12<br />

<br />

12<br />

− c22 = 0,<br />

c11<br />

+ (M12<br />

13 )2<br />

c11M 12<br />

12<br />

+ (M123<br />

124 )2<br />

+ (M12<br />

1k )2<br />

c11M 12<br />

12<br />

+ (M12...k−2k−1<br />

12...k−2k ) 2<br />

M 12...k−2<br />

12...k−2 M12...k−1<br />

12...k−1<br />

+ M123<br />

123<br />

M12 <br />

− c33 ,<br />

12<br />

M 12<br />

12 M123<br />

123<br />

+···<br />

<br />

.<br />

+ M1234<br />

1234<br />

M123 − c44<br />

123<br />

+ (M123<br />

12k )2<br />

M 12<br />

12 M123<br />

123<br />

+···<br />

+ M12...k<br />

12...k<br />

M 12...k−1<br />

− ckk<br />

12...k−1<br />

Since by Corol<strong>la</strong>ry A.5 we have<br />

<br />

<br />

<br />

A<br />

<br />

k−1<br />

k−1 (Ck) A k−1<br />

k (Ck)<br />

Ak k−1 (Ck) Ak k (Ck)<br />

<br />

<br />

<br />

= A∅ k−1k<br />

∅ (Ck)Ak−1k (Ck) or<br />

<br />

<br />

<br />

A<br />

<br />

k−1<br />

k−1 (Ck) A k−1k<br />

k−1k (Ck)<br />

<br />

<br />

<br />

= A k k−1 (Ck) 2 ,<br />

A ∅ ∅ (Ck) A k k (Ck)<br />

we conclu<strong>de</strong> that<br />

<br />

M<br />

<br />

<br />

12...k−1<br />

12...k−1 (Ck) M 12...k−2<br />

12...k−2 (Ck)<br />

M12...k 12...k (Ck) M 12...k−2k<br />

12...k−2k (Ck)<br />

<br />

<br />

<br />

= M 12...k−2k−1<br />

12...k−2k (Ck) 2 .<br />

Hence<br />

(M 12...k−2k−1<br />

12...k−2k (Ck)) 2<br />

M 12...k−2 12...k−1<br />

12...k−2 (Ck)M12...k−1 12...k (Ck)<br />

M 12...k−1<br />

12...k−1<br />

(Ck) + M12...k<br />

<br />

.<br />

<br />

.<br />

12...k−2k (Ck)<br />

M12...k−2k<br />

=<br />

(Ck) M 12...k−2<br />

,<br />

(Ck)<br />

12...k−2


and<br />

S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 679<br />

I k+1<br />

<br />

c2 1k<br />

k+1 (0) = M12...k 12...k<br />

c11<br />

+ (M12<br />

1k )2<br />

c11M 12<br />

12<br />

+ (M12...k−3k−2<br />

12...k−3k ) 2<br />

M 12...k−3<br />

12...k−3 M12...k−2<br />

12...k−2<br />

+ (M123<br />

12k )2<br />

M 12<br />

12 M123<br />

123<br />

+ M12...k−2k<br />

12...k−2k<br />

M 12...k−2<br />

12...k−2<br />

+···<br />

− ckk<br />

If we change k with k − 1 in the <strong>la</strong>st expression we obtain the right-hand part (up to a positive<br />

factor) of the expression for I k k (0).<br />

Finally we have proved (77) for I k+1<br />

k+1 (λ). Let us consi<strong>de</strong>r the function<br />

We have<br />

i k+1<br />

k+1<br />

by (78) and Remark C.2. So<br />

k+1<br />

(0) = Ik+1 (0) = 0 and<br />

i k+1 k+1<br />

k+1 (t) = Ik+1 (t ˆλ), t ∈ R.<br />

di k+1<br />

k+1 (t)<br />

dt<br />

=<br />

k<br />

p=2<br />

I k k >Ik k (ˆλ) = i k k (t) t=1 0.<br />

∂I k+1<br />

k+1 (λ)<br />

∂λp<br />

<br />

.<br />

ˆλp > 0<br />

We recall (see (32)) that for λ = (λ1,...,λm) ∈ C m and 1 k m we <strong>de</strong>note<br />

Using (27)<br />

we get<br />

λ [k] = (0,...,0,λk+1,...,λm), λ {k} = (λ1,...,λk, 0,...,0).<br />

Gm(λ) = A ∅ ∅ Cm(λ) =<br />

A k k Cm(λ) =<br />

<br />

∅⊆δ⊆{1,2,...,m}<br />

<br />

∅⊆δ⊆{1,2,...,k−1,k+1,...m}<br />

If we put Cm(λ [k] ) = Cm + m r=k+1 λrErr in (79) we get<br />

A k [k]<br />

k Cm λ <br />

=<br />

∅⊆δ⊆{k+1,k+2,...,m}<br />

Simi<strong>la</strong>rly, if we put Cm(λ) = Cm(λ {k} ) + m r=k+1 λrErr we get<br />

A k k Cm(λ) =<br />

<br />

∅⊆δ⊆{k+1,k+2,...,m}<br />

λδA δ δ (C),<br />

λδA k∪δ<br />

k∪δ (Cm). (79)<br />

λδA k∪δ<br />

k∪δ (Cm). (80)<br />

λδA k∪δ<br />

{k}<br />

k∪δ Cm λ . (81)


680 S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681<br />

Using (76) we have<br />

fk ˆλkA k k (Ck) A k k Ck(ˆλ) −1 = ˆλkA kk+1...m<br />

kk+1...m (Cm) A kk+1...m<br />

kk+1...m Cm(ˆλ) −1 hence I k m = fkA k k (Cm(ˆλ)) − ˆλkA k k (Cm(λ [k] )) I k m (ˆλ), where the function I k m (ˆλ) is <strong>de</strong>fined by<br />

kk+1...m<br />

Akk+1...m Cm(ˆλ) −1 kk+1...m<br />

A<br />

∅⊆δ⊆{k+1,k+2,...,m}<br />

<br />

Cm(ˆλ) − ˆλkA k <br />

k Cm ˆλ [k]<br />

<br />

<br />

I k m (ˆλ) := ˆλk<br />

kk+1...m (Cm)A k k<br />

<br />

= ˆλk<br />

kk+1...m<br />

Akk+1...m Cm(ˆλ) <br />

−1 A<br />

<br />

<br />

kk+1...m<br />

kk+1...m (Cm) Ak k (Cm(ˆλ [k] ))<br />

A kk+1...m<br />

kk+1...m (Cm(ˆλ)) Ak k (Cm(ˆλ))<br />

<br />

<br />

(80), (81) <br />

= ˆλk<br />

kk+1...m<br />

Akk+1...m Cm(ˆλ) −1 <br />

<br />

A<br />

×<br />

ˆλδ <br />

<br />

kk+1...m<br />

kk+1...m (Cm) Ak∪δ k∪δ (Cm)<br />

A kk+1...m<br />

kk+1...m (Cm(ˆλ)) Ak∪δ k∪δ (Cm(ˆλ {k} <br />

<br />

<br />

))<br />

.<br />

Using (26) or (27) we conclu<strong>de</strong> for λ = (0,λ2,...,λm) ∈ C m<br />

Finally we obtain<br />

I k m (ˆλ) = ˆλk<br />

A kk+1...m<br />

kk+1...m Cm(λ) =<br />

A k∪δ<br />

{k}<br />

k∪δ Cm λ =<br />

×<br />

kk+1...m<br />

Akk+1...m Cm(ˆλ) −1 <br />

∅⊆γ ⊆{2,3,...,k−1}<br />

<br />

∅⊆γ ⊆{2,3,...,k−1}<br />

<br />

∅⊆γ ⊆{2,3,...,k−1}<br />

<br />

<br />

ˆλγ<br />

<br />

<br />

<br />

γ ∪{k,k+1,...m}<br />

λγ Aγ ∪{k,k+1,...m} (Cm),<br />

<br />

∅⊆δ⊆{k+1,k+2,...,m}<br />

A kk+1...m<br />

kk+1...m (Cm) A<br />

A k∪δ<br />

k∪δ (Cm) A<br />

γ ∪{k}∪δ<br />

λγ Aγ ∪{k}∪δ (Cm).<br />

ˆλδ<br />

γ ∪{k,k+1,...m}<br />

γ ∪{k,k+1,...m} (Cm)<br />

γ ∪{k}∪δ<br />

γ ∪{k}∪δ (Cm)<br />

<br />

<br />

<br />

0<br />

<br />

<strong>du</strong>e to the Hadamard–Fisher’s inequality (Lemma A.6), for α ={k,k + 1,...,m} and β = γ ∪<br />

{k}∪δ. This comp<strong>le</strong>tes the proof of Lemma 16. ✷<br />

References<br />

[1] S. Albeverio, R. Höegh-Krohn, The energy representation of Sobo<strong>le</strong>v–Lie group, Compos. Math. 36 (1978) 37–52.<br />

[2] S. Albeverio, A. Kosyak, Quasiregu<strong>la</strong>r representations of the infinite-dimensional Borel group, J. Funct.<br />

Anal. 218 (2) (2005) 445–474.<br />

[3] S. Albeverio, A. Kosyak, Group action, quasi-invariant mea<strong>sur</strong>es and quasiregu<strong>la</strong>r representations of the infinitedimensional<br />

nilpotent group, Contemp. Math. 385 (2005) 259–280.<br />

[4] S. Albeverio, R. Höegh-Krohn, D. Testard, Irre<strong>du</strong>cibility and re<strong>du</strong>cibility for the energy representation of a group<br />

of mapping of a Riemannian manifold into a compact Lie group, J. Funct. Anal. 41 (1981) 378–396.<br />

[5] S. Albeverio, R. Höegh-Krohn, D. Testard, A. Vershik, Factorial representations of path groups, J. Funct. Anal. 51<br />

(1983) 115–131.<br />

[6] S. Albeverio, R. Höegh-Krohn, J. Marion, D. Testard, B. Torrésani, Noncommutative Distributions, Unitary Representation<br />

of Gauge Groups and Algebras, Monogr. Textbooks Pure Appl. Math., vol. 175, Dekker, New York,<br />

1993.


S. Albeverio, A. Kosyak / Journal of Functional Analysis 236 (2006) 634–681 681<br />

[7] E.F. Beckenbach, R. Bellmann, Inequalities, Springer, Berlin, 1961.<br />

[8]J.Dixmier,LesC∗-algèbres et <strong>le</strong>ur représentation, Gautier–Vil<strong>la</strong>rs, Paris, 1969.<br />

[9] J. Dixmier, Les algèbres d’opérateurs dans l’espace hilbertien, second ed., Gauthier–Vil<strong>la</strong>rs, Paris, 1969.<br />

[10] R.F. Gantmacher, Matrizenrechnung, Teil 1, VEB, Berlin, 1958.<br />

[11] I.M. Gel’fand, A.M. Vershik, M.I. Graev, Representations of SL2(R),whereRis a ring of functions, Uspekhi Mat.<br />

Nauk 28 (1973) 83–128.<br />

[12] R.A. Horn, C.R. Jonson, Matrix Analysis, Cambridge Univ. Press, Cambridge, 1989.<br />

[13] R.A. Horn, C.R. Jonson, Topics in Matrix Analysis, Cambridge Univ. Press, Cambridge, 1991.<br />

[14] R.S. Ismagilov, Representations of the group of smooth mappings in a compact Lie group, Funkts. Anal. i<br />

Priložhen. 15 (2) (1981) 73–74 (in Russian).<br />

[15] S. Kakutani, On equiva<strong>le</strong>nce of infinite pro<strong>du</strong>ct mea<strong>sur</strong>es, Ann. of Math. 4 (1948) 214–224.<br />

[16] A.A. Kirillov, E<strong>le</strong>ments of the Theory of Representations, Grund<strong>le</strong>hren Math. Wiss., vol. 220, Springer, Berlin,<br />

1976 (trans<strong>la</strong>ted from the Russian).<br />

[17] A.A. Kirillov, Intro<strong>du</strong>ction to the theory of representations and noncommutative harmonic analysis, in: Representation<br />

Theory and Noncommutative Harmonic Analysis, I, in: Encyclopaedia Math. Sci., vol. 22, Springer, Berlin,<br />

1994, pp. 1–156.<br />

[18] A.V. Kosyak, Irre<strong>du</strong>cibility criterion for regu<strong>la</strong>r Gaussian representations of group of finite upper triangu<strong>la</strong>r matrices,<br />

Funkts. Anal. i Priložhen. 24 (3) (1990) 82–83 (in Russian); transl. in: Funct. Anal. Appl. 24 (3) (1990)<br />

243–245.<br />

[19] A.V. Kosyak, Criteria for irre<strong>du</strong>cibility and equiva<strong>le</strong>nce of regu<strong>la</strong>r Gaussian representations of group of finite upper<br />

triangu<strong>la</strong>r matrices of infinite or<strong>de</strong>r, Se<strong>le</strong>cta Math. Soviet. 11 (1992) 241–291.<br />

[20] A.V. Kosyak, Irre<strong>du</strong>cib<strong>le</strong> regu<strong>la</strong>r Gaussian representations of the group of the interval and the circ<strong>le</strong> diffeomorphisms,<br />

J. Funct. Anal. 125 (1994) 493–547.<br />

[21] A.V. Kosyak, Regu<strong>la</strong>r representations of the group of finite upper-triangu<strong>la</strong>r matrices, corresponding to pro<strong>du</strong>ct<br />

mea<strong>sur</strong>es, and criteria for their irre<strong>du</strong>cibility, Methods Funct. Anal. Topology 6 (2000) 43–65.<br />

[22] A.V. Kosyak, The generalized Ismagilov conjecture for the group BN 0 . I, Methods Funct. Anal. Topology 8 (2)<br />

(2002) 33–49.<br />

[23] A.V. Kosyak, The generalized Ismagilov conjecture for the group BN 0 . II, Methods Funct. Anal. Topology 8 (3)<br />

(2002) 27–45.<br />

[24] A.V. Kosyak, A criterion for irre<strong>du</strong>cibility of quasiregu<strong>la</strong>r representations of the group of finite upper-triangu<strong>la</strong>r<br />

matrices, Funkts. Anal. i Priložhen. 37 (1) (2003) 78–81 (in Russian); transl. in: Funct. Anal. Appl. 37 (1) (2003)<br />

65–68.<br />

[25] A.V. Kosyak, Quasi-invariant mea<strong>sur</strong>es and irre<strong>du</strong>cib<strong>le</strong> representations of the in<strong>du</strong>ctive limit of the special linear<br />

groups, Funkts. Anal. i Priložhen. 38 (1) (2004) 82–84 (in Russian); transl. in: Funct. Anal. Appl. 38 (1) (2004)<br />

67–68.<br />

[26] H.H. Kuo, Gaussian Mea<strong>sur</strong>es in Banach Spaces, Lecture Notes in Math., vol. 463, Springer, Berlin, 1975.<br />

[27] S.M. Ma<strong>la</strong>mud, A converse to the Jensen inequality, its matrix extensions and inequalities for minors and eigenvalues,<br />

Linear Algebra Appl. 322 (1–3) (2001) 19–41.<br />

[28] M.P. Malliavin, Naturality of quasi-invariance of some mea<strong>sur</strong>es, in: A.B. Cruzeiro, J.-C. Zambrini (Eds.), Stochastic<br />

Analysis and Applications, Lisbon, 1989, in: Progress in Probability, vol. 26, Birkhäuser Boston, Boston, MA,<br />

1991, pp. 144–154.<br />

[29] M.P. Malliavin, Probability and geometry, in: Taniguchi Conference on Mathematics, Nara’98, in: Adv. Stud. Pure<br />

Math., vol. 31, Math. Soc. Japan, Tokyo, 2001, pp. 179–209.<br />

[30] M.P. Malliavin, P. Malliavin, Mea<strong>sur</strong>es quasi invariantes <strong>sur</strong> certain <strong>groupe</strong>s <strong>de</strong> dimension infinie, C. R. Acad. Sci.<br />

Paris Sér. I Math. 311 (1990) 765–768.<br />

[31] M.P. Malliavin, P. Malliavin, Integration on loop groups, I, J. Funct. Anal. 93 (1990) 207–237.<br />

[32] M.P. Malliavin, P. Malliavin, Integration on loop groups, III: Asymptotic Peter–Weyl orthogonality, J. Funct.<br />

Anal. 108 (1992) 13–46.<br />

[33] N.I. Nessonov, Examp<strong>le</strong>s of factor-representations of the group GL(∞), in: Mathematical Physics, Functional<br />

Analysis, Naukova Dumka, Kiev, 1986, pp. 48–52 (in Russian).<br />

[34] M. Reed, B. Simon, Methods of Mo<strong>de</strong>rn Mathematical Physics, vol. I, Aca<strong>de</strong>mic Press, New York, 1972.<br />

[35] E.T. Shavgulidze, Distributions on infinite-dimensional spaces and second quantization in string theories, II, in:<br />

V International Vilnius Conference on Probability Theory and Math. Statistics, Abstracts of Comm., Vilnius,<br />

June 26–July 1, 1989, pp. 359–360.<br />

[36] G.E. Shilov, Fan Dik Tun’, Integral, Mea<strong>sur</strong>e, and Derivative on Linear Spaces, Nauka, Moscow, 1967 (in Russian).<br />

[37] D.P. Zhelobenko, A.I. Shtern, Representations of Lie Groups, Nauka, Moscow, 1983.


Journal of Functional Analysis 236 (2006) 682–711<br />

www.elsevier.com/locate/jfa<br />

Finite range <strong>de</strong>compositions of positive-<strong>de</strong>finite<br />

functions<br />

David Brydges a,∗ , Anna Ta<strong>la</strong>rczyk b<br />

a Department of Mathematics, University of British Columbia, Vancouver, BC V6T 1Z2, Canada<br />

b Institute of Mathematics, University of Warsaw, ul. Banacha 2, 02-097 Warszawa, Po<strong>la</strong>nd<br />

Received 9 February 2006; accepted 10 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 2 May 2006<br />

Communicated by L. Gross<br />

Abstract<br />

We give sufficient conditions for a positive-<strong>de</strong>finite function to admit <strong>de</strong>composition into a sum of<br />

positive-<strong>de</strong>finite functions which are compactly supported within disks of increasing diameters Ln .More<br />

generally we consi<strong>de</strong>r positive-<strong>de</strong>finite bilinear forms f → v(f,f ) <strong>de</strong>fined on C∞ 0 .Wesayvhas a finite<br />

range <strong>de</strong>composition if v can be written as a sum v = Gn of positive-<strong>de</strong>finite bilinear forms Gn such<br />

that Gn(f, g) = 0 when the supports of the test functions f,g are separated by a distance greater or equal<br />

to Ln . We prove that such <strong>de</strong>compositions exist when v is <strong>du</strong>al to a bilinear form ϕ → |Bϕ| 2 where B is<br />

a vector valued partial differential operator satisfying some regu<strong>la</strong>rity conditions.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Positive-<strong>de</strong>finite; Generalized Gaussian field; Renormalisation group; Elliptic operator; Green’s function<br />

1. Intro<strong>du</strong>ction<br />

Let Λ be an open set in Rd , which may be all of Rd , and <strong>le</strong>t D(Λ) = C∞ 0 (Λ). Letf,g ↦→<br />

v(f,g) be a bilinear form <strong>de</strong>fined for f,g ∈ D(Λ). The form v is said to be positive-<strong>de</strong>finite if<br />

v(f,f ) > 0 for every non-vanishing f ∈ D(Λ).<br />

* Corresponding author.<br />

E-mail addresses: db5d@math.ubc.ca (D. Brydges), annatal@mimuw.e<strong>du</strong>.pl (A. Ta<strong>la</strong>rczyk).<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.008


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 683<br />

Given positive-<strong>de</strong>finite continuous bilinear forms, v and Gj <br />

,j ∈ N, onD(Λ), we write v =<br />

j1 Gj when<br />

v(f,f ) = <br />

Gj (f, f )<br />

j1<br />

holds for all f ∈ D(Λ).<br />

The forms v, Gj are said to be trans<strong>la</strong>tion invariant if Λ = Rd and v(Ttf,Ttg) = v(f,g) for<br />

all t ∈ Rd , where Ttf(x)= f(x+ t). We often encounter the case where the bilinear form arises<br />

from a function ˜v(x − y) such that<br />

<br />

v(f,g) = f(x)˜v(x − y)g(y)dx dy.<br />

˜v is said to be a positive-<strong>de</strong>finite function when the form v is positive-<strong>de</strong>finite.<br />

Definition 1.1. Let v be a trans<strong>la</strong>tion invariant bilinear form. We say that v admits a trans<strong>la</strong>tion<br />

invariant finite range <strong>de</strong>composition if, for some L>1, there exist positive-<strong>de</strong>finite forms Gj<br />

such that<br />

(1) v = <br />

j1 Gj ;<br />

(2) Gj (f, g) = 0ifdist(supp f,supp g) Lj ;<br />

(3) for j ∈ N, Gj is trans<strong>la</strong>tion invariant.<br />

This paper is concerned with the question of when a bilinear form has a finite range <strong>de</strong>composition.<br />

Our main result on the existence of such <strong>de</strong>compositions is given in Theorem 2.6. It<br />

concerns bilinear forms associated to the Green’s functions of a <strong>la</strong>rge c<strong>la</strong>ss of elliptic partial<br />

differential operators. It also gives a <strong>de</strong>composition if v is not trans<strong>la</strong>tion invariant and then con-<br />

dition (3) is rep<strong>la</strong>ced by a kind of uniformity of convergence of <br />

j1 Gj . Theorem 2.8 and<br />

Proposition 2.9 give a more explicit form of this <strong>de</strong>composition. In Section 4 we give concrete<br />

examp<strong>le</strong>s based on the construction used to prove existence in Section 3.<br />

Our interest is motivated by an equiva<strong>le</strong>nt question concerning Gaussian random fields. It<br />

is well known (e.g., see [9]) that for any continuous bilinear form v there exists a generalized<br />

Gaussian random field, i.e. a distribution valued random variab<strong>le</strong> φ such that for any<br />

test functions ϕ1,...,ϕn ∈ D(Λ) the vector (〈φ,ϕ1〉,...,〈φ,ϕn〉) is centered Gaussian and<br />

Cov(〈φ,ϕ〉, 〈φ,ψ〉) = v(ϕ,ψ). The existence of a finite range <strong>de</strong>composition of v is equiva-<br />

<strong>le</strong>nt to the existence of a <strong>de</strong>composition φ = <br />

j1 ζj , where ζj are in<strong>de</strong>pen<strong>de</strong>nt generalized<br />

Gaussian random fields with covariance functionals Gj respectively and such that 〈ζj ,ϕ〉 and<br />

〈ζj ,ψ〉 are in<strong>de</strong>pen<strong>de</strong>nt if dist(supp ϕ,supp ψ) L j .<br />

Many mo<strong>de</strong>ls in statistical mechanics have the form of an expectation EZ of a nonlinear<br />

functional Z of a Gaussian random field φ, where φ has long range power <strong>la</strong>w corre<strong>la</strong>tions and<br />

the functional Z <strong>de</strong>pends on the field φ in a <strong>la</strong>rge region Λ ⊂ R d . The <strong>la</strong>rge size of Λ and the<br />

long range corre<strong>la</strong>tions make it difficult to get accurate estimates on EZ. The c<strong>la</strong>ss of methods<br />

known as the Renormalisation Group (RG) was originated by K.G. Wilson [12,13] in or<strong>de</strong>r to<br />

address this prob<strong>le</strong>m.<br />

A very convenient version of Wilson’s i<strong>de</strong>a was intro<strong>du</strong>ced in rigorous mathematical arguments<br />

by Gal<strong>la</strong>votti et al. in [1,2]. These authors write the field φ = <br />

j1 ζj as a sum of


684 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

in<strong>de</strong>pen<strong>de</strong>nt increments ζj so that Z becomes a function Z = Z1(ζ1 + ζ2 +···) of all the increments<br />

and then they <strong>de</strong>fine the conditional expectation Ej to be the integration over increment ζj .<br />

E.g., <strong>le</strong>t Zj+1 = Ej Zj so that EZ = lim Zj . The paper [1] is a good intro<strong>du</strong>ction to the RG.<br />

In essentially all papers based on [2] and re<strong>la</strong>ted <strong>de</strong>compositions intro<strong>du</strong>ced by other authors,<br />

e.g., [6,7], the <strong>de</strong>cay of the ζj covariance is exponential on sca<strong>le</strong> L j but not finite range as in<br />

property (2) above. The machinery known as “cluster expansions” was <strong>de</strong>veloped to control this<br />

<strong>la</strong>ck of exact in<strong>de</strong>pen<strong>de</strong>nce and these expansions have become the workhorse of rigorous work<br />

on the RG. Thus finite range <strong>de</strong>compositions are certainly not essential for progress in the RG,<br />

but by removing the need for cluster expansions they rep<strong>la</strong>ce a major technical prerequisite of<br />

this subject. Wave<strong>le</strong>t <strong>de</strong>compositions are also used in the RG [5]. These <strong>de</strong>compositions can<br />

have the finite range property but not in<strong>de</strong>pen<strong>de</strong>nce of ζi and ζj for j = i because they are not a<br />

<strong>de</strong>composition of the covariance into a sum of positive <strong>de</strong>finite forms.<br />

These consi<strong>de</strong>rations have motivated us to write this paper to prove that finite range <strong>de</strong>compositions<br />

exist for a wi<strong>de</strong> c<strong>la</strong>ss of Gaussian fields. In the case that the covariance of the Gaussian<br />

field φ is homogeneous, e.g., |x − y| −α , it is easy to establish existence of these <strong>de</strong>compositions<br />

and we have ma<strong>de</strong> this point and used them, for examp<strong>le</strong>, in [3]. In [4] we proved that the resolvent<br />

of the Lap<strong>la</strong>cian (aI − ) −1 with a 0 admits finite range <strong>de</strong>compositions and also<br />

showed that these <strong>de</strong>compositions exist when is the finite difference Lap<strong>la</strong>cian on the simp<strong>le</strong><br />

cubic <strong>la</strong>ttice Z d .<br />

Finite range <strong>de</strong>compositions for radial functions have appeared for different reasons in the<br />

context of the stability of matter. In [8] are given necessary and sufficient conditions on the<br />

<strong>de</strong>rivatives of a function f that <strong>de</strong>fines a radially symmetric kernel f(x − y) so that the bilinear<br />

form associated to f(x − y) has an expansion with non-negative coefficients into tent<br />

functions.<br />

We <strong>de</strong>fer precise <strong>de</strong>finitions and give a litt<strong>le</strong> outline of our results. Two bilinear forms v and<br />

E are said to be <strong>du</strong>al if the Hilbert space comp<strong>le</strong>tions of C ∞ 0 (Rd ) in the two bilinear forms<br />

are <strong>du</strong>al re<strong>la</strong>tive to the L 2 (R d ) inner pro<strong>du</strong>ct. Our main condition for v to admit a trans<strong>la</strong>tion<br />

invariant finite range <strong>de</strong>composition is that the <strong>du</strong>al bilinear form E is associated with a constant<br />

coefficient partial differential operator B by<br />

<br />

E (ϕ, ϕ) =<br />

R d<br />

|Bϕ| 2 dx,<br />

where B can be vector valued. B need not be first or<strong>de</strong>r. As an examp<strong>le</strong> consi<strong>de</strong>r<br />

so that<br />

<br />

E (ϕ, ϕ) =<br />

B = (∂1,...,∂d,λI)<br />

R d<br />

|∂ϕ| 2 + λ 2 |ϕ| 2 dx. (1.1)<br />

By integration by parts this form is associated to the elliptic partial differential operator B ′ B =<br />

− + λ 2 . The <strong>du</strong>ality of v,E means that the distribution kernel of v is a Green’s function<br />

(B ′ B) −1 for the partial differential operator B ′ B.


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 685<br />

Our condition does not characterise the forms that have trans<strong>la</strong>tion invariant finite range <strong>de</strong>compositions<br />

because one can <strong>de</strong><strong>du</strong>ce from it that the kernel of (B ′ B) −α with α ∈ (0, 1] also has<br />

a finite range <strong>de</strong>composition. We know of no examp<strong>le</strong>s of positive-<strong>de</strong>finite forms that are not in<br />

this wi<strong>de</strong>r c<strong>la</strong>ss.<br />

The construction in our proof achieves much more because it also creates <strong>de</strong>compositions<br />

satisfying items (1), (2) when B has non-constant coefficients and Λ need not be all of R d .<br />

In this case v is not trans<strong>la</strong>tion invariant, hence (3) cannot be satisfied and is rep<strong>la</strong>ced by the<br />

following. If the partial differential operator B has constant coefficients, but the domain Λ is not<br />

all of R d then the <strong>de</strong>composition satisfies:<br />

• Trans<strong>la</strong>tion invariance away from ∂Λ.Forj ∈ N small enough such that 2L j < diam Λ,<br />

Gj (Ttf,Ttf) is in<strong>de</strong>pen<strong>de</strong>nt of t and Λ for f,t such that the support of Ttf is in Λ but<br />

separated from the boundary of Λ by a distance greater than L j .<br />

In a few examp<strong>le</strong>s like the ones discussed in Section 4 where we can make explicit computations<br />

of norms, we find that the terms in our <strong>de</strong>compositions <strong>de</strong>cay with a scaling that correctly<br />

ref<strong>le</strong>cts the dimensional analysis of the operator B ′ B. If the coefficients of B are not constant,<br />

we do not know very much about the rate of convergence of the <strong>de</strong>composition. We only have<br />

the following estimate which says that the <strong>de</strong>composition converges uniformly with respect to<br />

trans<strong>la</strong>tion of the argument f :<br />

• Uniformity. There exists a constant c such for all L>cthere exist finite range <strong>de</strong>compositions<br />

such that for all f = B ′ Bϕ in B ′ BD(Λ),<br />

0 v(f,f ) − <br />

jn<br />

(n−p)∨0<br />

c<br />

Gj (f, f ) <br />

v(f,f ), (1.2)<br />

L<br />

where p is the smal<strong>le</strong>st integer such that diam supp ϕ L p . The c<strong>la</strong>ss B ′ BD(Λ) is <strong>de</strong>nse in<br />

the Hilbert space with inner pro<strong>du</strong>ct v.<br />

We examine these <strong>de</strong>compositions for the simp<strong>le</strong> case of the Lap<strong>la</strong>cian in Section 4 in or<strong>de</strong>r<br />

to un<strong>de</strong>rstand these <strong>de</strong>compositions more concretely and in particu<strong>la</strong>r to examine their smoothness.<br />

In Section 5 we continue these calcu<strong>la</strong>tions for the Lap<strong>la</strong>cian to construct a finite range<br />

<strong>de</strong>composition with C ∞ smoothness.<br />

2. Notation and main result on existence<br />

The proofs of the results in this section are found in Section 3.<br />

Let B = (B1,...,Bn) be an n-vector of partial differential operators,<br />

Bi = <br />

ci,α(x)∂ α .<br />

We call<br />

E (ϕ, ψ) =<br />

i=1<br />

a Dirich<strong>le</strong>t form. We impose the following assumptions.<br />

α<br />

n<br />

<br />

BiϕBiψdx= (Bϕ, Bψ) L2 (2.1)


686 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Assumptions.<br />

(1) For each i, α, ci,α ∈ C∞ (Λ).<br />

(2) For each ψ ∈ D(Λ) there exists c such that for all ϕ ∈ D(Λ)<br />

<br />

<br />

ψϕdx<br />

c E (ϕ, ϕ) 1/2 . (2.2)<br />

(3) The continuity Hypothesis 2.2 exp<strong>la</strong>ined below.<br />

By the first assumption Bϕ is <strong>de</strong>fined for all ϕ ∈ D(Λ) and<br />

B ′ ψ = <br />

i,α<br />

(−∂) α (ci,αψi)<br />

is <strong>de</strong>fined for ψ any vector-valued smooth function. Note that this implies that for any ϕ ∈ D(Λ)<br />

B ′ Bϕ ∈ D(Λ). By integration by parts (Bϕ, ψ) L 2 = (ϕ, B ′ ψ) L 2.<br />

By the second assumption, with ϕ = ψ, E (ψ, ψ) = 0 implies ψ = 0 and so E is an inner<br />

pro<strong>du</strong>ct <strong>de</strong>fined on D(Λ). LetH+(Λ) be the Hilbert space comp<strong>le</strong>tion of D(Λ) with inner<br />

pro<strong>du</strong>ct E . The corresponding norm will be <strong>de</strong>noted by ·+. For any open subset U ⊂ Λ, <strong>le</strong>t<br />

H+(U) be the closed subspace of H+(Λ) obtained by taking the clo<strong>sur</strong>e of D(U).<br />

Definition 2.1. We say that H+(U) is continuous at U if<br />

H+(V ): V ⊃ U = H+(U).<br />

The <strong>la</strong>st of the three hypotheses mentioned above is<br />

Hypothesis 2.2. Λ is convex and H+ is continuous at U for all sets U which are boun<strong>de</strong>d, convex<br />

and open.<br />

This is an implicit condition on the coefficients of the partial differential operator B. Asufficient<br />

condition for Hypothesis 2.2 to hold is provi<strong>de</strong>d by the following <strong>le</strong>mma.<br />

Lemma 2.3. Let U be a boun<strong>de</strong>d convex open set. H+ is continuous at U if, there exists x∗ ∈ U,<br />

δ>0, C>0 and a convex open Ũ ⊃ U, Ũ ⊂ Λ such that |ci,α(x∗ + λ(x − x∗))| C|ci,α(x)|<br />

for all x ∈ Ũ,alli, α and all λ ∈[1 − δ,1].<br />

Remark 2.4. In particu<strong>la</strong>r, Hypothesis 2.2 holds if the coefficients ci,α are boun<strong>de</strong>d away from<br />

zero in every boun<strong>de</strong>d subset of Λ.<br />

Let H−(Λ) be the abstract <strong>du</strong>al space of boun<strong>de</strong>d linear functionals on H+(Λ). H−(Λ) is<br />

contained in the space of distributions<br />

H−(Λ) ⊂ D ′ (Λ)


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 687<br />

because the topology on D(Λ) is stronger than the topology on H+(Λ) (a subsequence ϕn is<br />

convergent in the topology of D(Λ) if there exists a compact set K ⊂ Λ with supp ϕn ⊂ K and<br />

all <strong>de</strong>rivatives ∂ α ϕn converge uniformly). Therefore<br />

where<br />

H−(Λ) = f ∈ D ′ (Λ): f −,Λ < ∞ , (2.3)<br />

f −,Λ = sup 〈f,ϕ〉: ϕ ∈ D(Λ), E (ϕ, ϕ) 1 . (2.4)<br />

Since the <strong>du</strong>al of a Hilbert space is an isomorphic Hilbert space, the norm arises from an inner<br />

pro<strong>du</strong>ct.<br />

Definition 2.5. Define G(f, g) = (f, g)−,Λ to be the inner pro<strong>du</strong>ct on H−(Λ). In particu<strong>la</strong>r,<br />

G(f, f ) =f 2 −,Λ . (2.5)<br />

By Assumption (2), a function ψ ∈ D(Λ) <strong>de</strong>termines a linear functional fψ ∈ H−(Λ) by<br />

<br />

〈fψ,ϕ〉= ψϕdx<br />

and in this sense D(Λ) ⊂ H−(Λ) so that G is a positive-<strong>de</strong>finite bilinear form on D(Λ).<br />

The main result of this paper is the following.<br />

Theorem 2.6. Un<strong>de</strong>r Assumptions (1)–(3) given above, G admits a <strong>de</strong>composition satisfying<br />

(1), (2) in Definition 1.1 and the uniformity estimate (1.2). IfΛ = R d and if B has constant<br />

coefficients, then this <strong>de</strong>composition for G is a trans<strong>la</strong>tion invariant finite range <strong>de</strong>composition.<br />

If the partial differential operator B has constant coefficients, but the domain Λ is not all of R d ,<br />

then the <strong>de</strong>composition is trans<strong>la</strong>tion invariant away from ∂Λ as <strong>de</strong>fined above.<br />

To un<strong>de</strong>rstand why this is a <strong>de</strong>composition of the Green’s function for the differential operator<br />

B ′ B we use a standard argument cal<strong>le</strong>d the Friedrich’s extension [10, p. 278], [11, p. 177]. We<br />

will show that G is the form for a Green’s function for the partial differential operator B ′ B with<br />

zero boundary conditions on ∂Λ.<br />

By the <strong>de</strong>finition of the + norm, for all ϕ ∈ D(Λ), ϕ2 + =Bϕ2<br />

L2. Therefore the clo<strong>sur</strong>e<br />

¯B : H+(Λ) → L 2 Λ,R n<br />

of B is an isometry. Let ¯B ′ : L2(Λ, Rn ) → H−(Λ) be the <strong>du</strong>al operator and <strong>de</strong>fine L = ¯B ′ ¯B.<br />

This is a map from H+(Λ) to H−(Λ) and it satisfies<br />

Λ<br />

〈 ¯B ′ ¯Bϕ,ψ〉=(ϕ, ψ)+<br />

for all ϕ,ψ ∈ H+(Λ). Therefore it is the Riesz isomorphism that i<strong>de</strong>ntifies the Hilbert space<br />

H+(Λ) with the <strong>du</strong>al H−(Λ) and so G is re<strong>la</strong>ted to the inverse of L by<br />

G(f, g) = (f, g)−,Λ = L −1 f,g .<br />

On the domain ϕ ∈ D(Λ) we can omit the clo<strong>sur</strong>es so that L ϕ is our differential operator B ′ Bϕ.


688 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Now we give some results exp<strong>la</strong>ining in more <strong>de</strong>tail the construction of the <strong>de</strong>composition of<br />

Theorem 2.6.<br />

For U any convex boun<strong>de</strong>d open subset of Λ, <strong>le</strong>tpU be the orthogonal projection in H+(Λ)<br />

onto the subspace H+(U). In particu<strong>la</strong>r pU = 0forU =∅.Weset<br />

Ux = (x + U)∩ Λ.<br />

Lemma 2.7. For each ϕ ∈ H+(Λ) there exists a unique vector Tϕin H+(Λ) such that<br />

(T ϕ, ψ)+ = 1<br />

<br />

|U|<br />

The linear operator T : H+(Λ) ↦→ H+(Λ) is a contraction.<br />

dx(pUx ϕ,ψ)+ for all ψ ∈ H+(Λ). (2.6)<br />

The main ingredient of the <strong>de</strong>composition is the following theorem allowing to “cut out” from<br />

G a bilinear form which is positive-<strong>de</strong>finite and of finite range.<br />

Theorem 2.8. Let U be a convex, boun<strong>de</strong>d open subset of Λ.Forf ∈ H−(Λ), <strong>de</strong>fine<br />

Then the bilinear form G1 such that<br />

is:<br />

(1) positive-<strong>de</strong>finite;<br />

(2) finite range with range 2diamU.<br />

A ′ U f = f − T ′ f. (2.7)<br />

G(f, f ) = G1(f, f ) + G A ′ U f,A ′ <br />

U f<br />

We construct the finite range <strong>de</strong>composition by an iterated application of Theorem 2.8. Let<br />

U0 be a convex open boun<strong>de</strong>d set containing the origin and for j = 1, 2,...<strong>le</strong>t Uj be a sequence<br />

of domains obtained by scaling U,<br />

For each domain construct T ′<br />

j<br />

Uj = x: L −j x ∈ U .<br />

(2.8)<br />

, A ′<br />

j , ˜Gj by rep<strong>la</strong>cing U by Uj in the construction of G1 given<br />

above. Set f1 = f and for j 2, set fj = A ′<br />

j−1 fj−1. Define bilinear forms Gj for j 1by<br />

Gj (f, f ) = ˜Gj (fj ,fj ).<br />

Proposition 2.9.<br />

(1) G(f, f ) = n j=1 Gj (f, f ) +A ′<br />

n ...A ′<br />

1f 2 −,Λ .<br />

(2) Given L>1, <strong>le</strong>t the diameter of U0 be chosen <strong>le</strong>ss than 1<br />

2 (1 − L−1 ). Then for all j 1 the<br />

range of Gj is <strong>le</strong>ss than Lj .


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 689<br />

(3) Let U0 be as in part (2). Then there exists a constant c = c(U0)>1 such that for all n ∈ N,<br />

for all L>1, for all f ∈ B ′ BD(Λ),<br />

<br />

A ′<br />

n<br />

′<br />

...A 1f (n−p)∨0<br />

<br />

c<br />

<br />

f −,Λ,<br />

−,Λ L<br />

where f = B ′ Bϕ and p 0 is smal<strong>le</strong>st integer such that the diameter of the support of ϕ is<br />

<strong>le</strong>ss than L p .<br />

(4) For all f ∈ H−(Λ) and therefore, in particu<strong>la</strong>r, f ∈ D(Λ), G(f, f ) = <br />

j1 Gj (f, f ).<br />

Let us discuss a bit more our operator AU , which is the key ingredient of our <strong>de</strong>composition.<br />

Remark 2.10. Let U be a boun<strong>de</strong>d open subset of Λ. The linear operator <strong>de</strong>fined on H+(Λ) by<br />

PU = I − pU<br />

is cal<strong>le</strong>d the Poisson operator of the domain U for the following reason. Consi<strong>de</strong>r ψ = PU ϕ,<br />

where ϕ ∈ H+(Λ). Then (a) ψ satisfies B ′ Bψ(x) = 0forx∈U, where B ′ B is the partial differential<br />

operator applied in the distribution sense, and (b) ψ(x) − ϕ(x) ∈ H+(U). Part (b) is<br />

obvious by the <strong>de</strong>finition of PU . It imp<strong>le</strong>ments the boundary condition on ψ that ψ = ϕ outsi<strong>de</strong><br />

U. To check (a), <strong>le</strong>t φ ∈ D(U) be a test function and recall that B ′ B on the domain D(U)<br />

is the Riesz isomorphism. Therefore<br />

<br />

ψB ′ Bφdx = (PU ϕ,φ)+ = (ϕ, PU φ)+ = 0.<br />

Λ<br />

For examp<strong>le</strong>, when B ′ B =− and U has a smooth boundary, conditions (a), (b) constitute<br />

solving the Dirich<strong>le</strong>t prob<strong>le</strong>m insi<strong>de</strong> U with boundary conditions given by ϕ restricted to ∂U.<br />

Poisson kernels also provi<strong>de</strong> another useful representation for the operator AU of Theorem<br />

2.8.<br />

Lemma 2.11. Assume that U is a boun<strong>de</strong>d open set. Then<br />

AU ϕ = 1<br />

<br />

|U|<br />

R d<br />

(2.9)<br />

1x+U P(x+U)∩Λϕdx, (2.10)<br />

where P(x+U)∩Λ was <strong>de</strong>fined in (2.9). IfP(x+U)∩Λϕ(y) is jointly mea<strong>sur</strong>ab<strong>le</strong> with respect to<br />

(x, y) then (2.10) can be un<strong>de</strong>rstood as an integral <strong>de</strong>fined pointwise, and we can write<br />

AU ϕ(y) = 1<br />

<br />

P(x+U)∩Λϕ(y)dx. (2.11)<br />

|U|<br />

y−U<br />

Remark 2.12. In the case L = B ′ B =−, or more generally, L = λ 2 I − , we can use this<br />

to give the following interpretation to the construction of G1. The inner pro<strong>du</strong>ct (f1,f2)−,Λ<br />

is the interaction energy of two distributions fi of e<strong>le</strong>ctrostatic charge. Suppose that fi is an


690 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

approximate point mass distribution located at point Qi. The finite range of G1 can be un<strong>de</strong>rstood<br />

in terms of rep<strong>la</strong>cing fi by another distribution ρi chosen so that (1) ρi is supported in a compact<br />

region Ki centred on Qi; (2) the potential outsi<strong>de</strong> Ki is unchanged. Then, for Qi sufficiently far<br />

apart that Ki do not over<strong>la</strong>p (f1,f2)−,Λ − (ρ1,ρ2)−,Λ = 0.<br />

The factor P ′ Ux fi constructs a charge distribution, with these properties, on the boundary of<br />

Ux ∋ Qi. Then AU spreads the charge distribution fi out by (1) spreading out the charge at Qi<br />

onto the boundary of Ux and (2) averaging over all Ux containing point A. This very specific<br />

choice of ρi achieves the difficult simultaneous goals of making G1 positive-<strong>de</strong>finite and smal<strong>le</strong>r<br />

than G(f, f ) and finite range.<br />

The advantage compared with our construction in [4] is that this particu<strong>la</strong>r average over Ux<br />

allows us to avoid reliance on trans<strong>la</strong>tion invariance and to generalise so as to inclu<strong>de</strong> Green’s<br />

functions of operators that are of higher than second or<strong>de</strong>r.<br />

3. Proofs for Section 2<br />

In or<strong>de</strong>r to prove Lemma 2.3 we need the following result. Define Dλϕ(x) = ϕ(x/λ) where<br />

ϕ ∈ D(V ) and ϕ is consi<strong>de</strong>red to be a function on R d by <strong>de</strong>fining it to be zero outsi<strong>de</strong> V ⊂ Λ.<br />

Lemma 3.1. Let V be an open convex subset of Λ containing x∗ = 0. If there exist C1 > 0,δ>0<br />

such that |ci,α(λx)| C1|ci,α(x)|, for all i, α, λ ∈[1−δ,1] and x ∈ V , then Dλ uniquely extends<br />

to an operator Dλ : H+(V ) → H+(V ). The resulting family of operators is uniformly boun<strong>de</strong>d<br />

in operator norm and is <strong>le</strong>ft strongly continuous in λ at λ = 1.<br />

Proof. Let ϕ ∈ D(V ). Then Dλϕ ∈ D(V ). Also<br />

E (Dλϕ,Dλϕ) = <br />

<br />

ci,α∂ α Dλϕ 2 dx = <br />

λ d−2|α|<br />

<br />

ci,α(λx)∂ α ϕ 2 dx<br />

i,α<br />

C(δ) <br />

<br />

ci,α(x)∂ α ϕ 2 dx = C(δ)E (ϕ, ϕ).<br />

i,α<br />

D(V ) is <strong>de</strong>nse in H+(V ) so this proves that Dλ is uniformly boun<strong>de</strong>d in operator norm and<br />

extends uniquely. By dominated convergence,<br />

i,α<br />

lim Dλϕ − ϕ+ = 0<br />

λ↑1<br />

for ϕ ∈ D(V ). By the uniform boun<strong>de</strong>dness of Dλ strong continuity on a <strong>de</strong>nse subset implies<br />

strong continuity. ✷<br />

Proof of Lemma 2.3. We give the proof for case x∗ = 0. Let ϕ ∈ {H+(V ): V ⊃ U}. Since<br />

{H+(V ): V ⊃ U} ⊃H+(U), it suffices to prove that ϕ ∈ H+(U). LetUn be the open set<br />

of all points in Λ within distance 2 −n of U. Then there exists a sequence ϕn ∈ D(Un) that<br />

converges to ϕ in H+ norm. By Lemma 3.1, for 1 − δ λ


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 691<br />

Proof of Theorem 2.6. This is a direct consequence of Proposition 2.9 so it suffices to prove the<br />

remaining results of Section 1.<br />

In or<strong>de</strong>r to show Lemma 2.7 and Theorem 2.8 we need some properties of the operators pU<br />

and T . Lemmas 3.3 and 3.5 show that operator T of (2.6) is well <strong>de</strong>fined and is a contraction.<br />

Lemma 3.10 and Proposition 3.11 are essential for the proof that bilinear form G1 in (2.8) is<br />

positive <strong>de</strong>finite and of finite range.<br />

Lemma 3.2. pU pV = 0 if U ∩ V =∅.<br />

Proof. First, note that if U,V are disjoint open sets, if ϕ ∈ H+(U) and ψ ∈ H+(V ), then<br />

(ψ, ϕ)+ = 0, because it is enough to prove this when ϕ ∈ D(U) and ψ ∈ D(V ) and then<br />

(ψ, ϕ)+ = (Bψ, Bϕ) L 2 = 0 because B is a partial differential operator (PDO). By this remark,<br />

for any ψ ∈ H+(Λ), (ψ, pU pV ϕ)+ = (pU ψ,pV ϕ)+ = 0. ✷<br />

Lemma 3.3. If H+ is continuous at U and U is boun<strong>de</strong>d, then pUx<br />

at x = 0.<br />

is strongly continuous in x<br />

Proof. We will give the proof in three steps (open and closed are <strong>de</strong>fined as subsets of Λ with<br />

re<strong>la</strong>tive topology):<br />

(1) If Un, n = 1, 2,..., is a sequence of open sets such that for every compact set K ⊂ U, there<br />

exists n(K) such that Un ⊃ K for n n(K). Then pUnpU converges strongly to pU .<br />

(2) If H+ is continuous at U, ifUn,n= 1, 2,..., is a sequence of open sets such that for every<br />

open V ⊃ U, there exists n(V ) such that Un ⊂ V for n n(V ). Then pUn (I −pU ) converges<br />

strongly to 0.<br />

(3) If H+ is continuous at U and U is boun<strong>de</strong>d, then pUx is strongly continuous in x at x = 0.<br />

(1) Let ϕ ∈ H+(Λ). We have to prove that pUnpU ϕ → pU ϕ in norm. Equiva<strong>le</strong>ntly, we must<br />

prove pUnϕ → ϕ for any ϕ ∈ H+(U). It suffices to prove this for ϕ ∈ D(U), because pUn are<br />

uniformly boun<strong>de</strong>d in operator norm. By the hypothesis on the sequence Un, Un contains the<br />

support of ϕ for all sufficiently <strong>la</strong>rge n and therefore pUnϕ = ϕ for all sufficiently <strong>la</strong>rge n.<br />

(2) It suffices to prove that pUnϕ+ → 0 for all ϕ ∈ H+(U) ⊥ . Every subsequence of pUnϕ has a weakly convergent subsequence, because the ball in H+(Λ) is weakly compact. Let φ be<br />

the limit of a subsequence. Then, by the hypothesis on Un, for any V ⊃ U, φ ∈ H+(V ), because<br />

H+(V ) is a subspace and therefore weakly closed. By the continuity hypothesis, Definition 2.1,<br />

φ ∈ H+(U). Therefore, along this subsequence<br />

pUn ϕ2 + = (pUn ϕ,ϕ)+ → (φ, ϕ)+ = 0.<br />

Since this is valid for every subsequence, pUn ϕ+ → 0.<br />

(3) Let ϕ ∈ H+(Λ) and <strong>le</strong>t x → 0. We have<br />

pUx∩U ϕ = pUx∩U pU ϕ → pU ϕ by (1),<br />

pUx∪U ϕ − pU ϕ = pUx∪U ϕ − pUx∪U pU ϕ → 0 by(2). (3.1)


692 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Notice that pUx∪U ϕ = pUx ϕ + ψx, where ψx ∈ H+(Ux) ⊥ and pUx ϕ = pUx∩U ϕ + ζx, where<br />

ζx ∈ H+(Ux). Therefore by orthogonality of ψx and ζx and then (3.1) we obtain<br />

ψx 2 + +ζx 2 + =ψx + ζx 2 + =pUx∪U ϕ − pUx∩U ϕ 2 + → 0.<br />

Hence ψx tends to zero and consequently, using (3.1) again, pUx ϕ → pU as x → 0. ✷<br />

The support of Bϕ is contained in supp ϕ for ϕ ∈ D(Λ) because B is a PDO. Since any<br />

ϕ ∈ H+(U) can be approximated by a sequence ϕn with supp ϕn ⊂ U,<br />

¯BH+(U) ⊂ L 2 (U). (3.2)<br />

From now on we assume that U is a boun<strong>de</strong>d open convex set and Λ is convex so that, by<br />

Hypothesis 2.2, H+ is continuous at Ux = (x + U) ∩ Λ for all x ∈ R d .Fixϕ ∈ H+(Λ) and<br />

<strong>de</strong>fine<br />

Fx = ¯BpUx ϕ.<br />

Since ¯B is an isometry, Lemma 3.3 implies that x ↦→ Fx is a norm-continuous map from R d<br />

to L 2 .AlsoFx is boun<strong>de</strong>d in L 2 norm uniformly in x. Therefore (Fx,Fy) L 2 is continuous in<br />

(x, y), so that the following integrals are well <strong>de</strong>fined or possibly positive infinite.<br />

Lemma 3.4.<br />

<br />

<br />

dx<br />

dy <br />

<br />

(Fx,Fy) L2 |U|<br />

where |U| <strong>de</strong>notes the Lebesgue mea<strong>sur</strong>e of U.<br />

dx(Fx,Fx) L 2,<br />

Proof. Let ɛ>0. Choose disjoint cubes of si<strong>de</strong> ɛ whose union equals R d and insert the partition<br />

of unity 1 = 1Δ,<br />

<br />

<br />

dx<br />

dy <br />

<br />

(Fx,Fy) L2 =<br />

<br />

dx<br />

<br />

<br />

dy<br />

<br />

Δ<br />

(Fx, 1ΔFy) L 2<br />

The sum over Δ was interchanged with the inner pro<strong>du</strong>ct using Fubini’s theorem and<br />

¯Fx(·)Fy(·) ∈ L1.<br />

By (3.2), we can insert 1Uy∩Δ=∅,Ux∩Δ=∅. Taking the sum over Δ outsi<strong>de</strong> the absolute values<br />

we bound by<br />

<br />

<br />

dx<br />

dy <br />

Insert the Cauchy–Schwartz inequality in the form<br />

Δ<br />

<br />

1Uy∩Δ=∅,Ux∩Δ=∅<br />

(Fx, 1ΔFy) L2 (Fx, 1ΔFy) L 2 1<br />

2 (Fx, 1ΔFx) L 2 + 1<br />

2 (Fy, 1ΔFy) L 2.<br />

<br />

.<br />

<br />

<br />

<br />

.


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 693<br />

The resulting integrand is jointly mea<strong>sur</strong>ab<strong>le</strong> and non-negative so integrals and sums can be put<br />

in any or<strong>de</strong>r. Therefore both terms give the same contribution and we bound by<br />

<br />

<br />

dx dy1Uy∩Δ=∅(Fx, 1ΔFx) L2. Δ<br />

Let |Uɛ|= dy1Uy∩Δ=∅. For the future, note that |Uɛ| tends as ɛ → 0tothevolume|U| of U.<br />

Then the preceding expression equals<br />

|Uɛ| <br />

<br />

<br />

dx(Fx, 1ΔFx) L2 =|Uɛ| dx(Fx,Fx) L2 and the <strong>le</strong>mma follows by taking ɛ → 0. ✷<br />

Lemma 3.5.<br />

Δ<br />

<br />

dx (pUx ϕ,ψ)+<br />

<br />

|U|ϕ+ψ+.<br />

Proof. It is sufficient to prove case ψ = ϕ because<br />

<br />

dx (pUx ϕ,ψ)+<br />

<br />

<br />

= dx <br />

<br />

(pUx ϕ,pUx ψ)+ dxpUx ϕ+pUx ψ+<br />

<br />

<br />

<br />

pUx ϕ2 + dx<br />

1/2 <br />

pUy ψ2 + dy<br />

1/2 .<br />

Now we consi<strong>de</strong>r case ϕ = ψ for which there are no absolute values because (pUx ϕ,ϕ)+ 0.<br />

By the Cauchy–Schwartz inequality,<br />

<br />

<br />

<br />

<br />

<br />

λi(pUx ϕ,ϕ)+ = λipUx ϕ,ϕ λipUx ϕ,<br />

i i i<br />

i<br />

i<br />

+ i<br />

<br />

1/2 λj pUx ϕ ϕ+<br />

j<br />

j<br />

+<br />

<br />

<br />

1/2 = ¯λiλj (pUx ϕ,pUx ϕ)+ ϕ+.<br />

i j<br />

By consi<strong>de</strong>ring Riemann sums, this implies<br />

<br />

dx(pUxϕ,ϕ)+ <br />

<br />

<br />

dx<br />

By Lemma 3.4,<br />

<br />

dx<br />

i,j<br />

dy(pUxϕ,pUy ϕ)+<br />

<br />

=<br />

<br />

dx<br />

<br />

=|U|<br />

1/2 dy(pUxϕ,pUy ϕ)+ ϕ+. (3.3)<br />

<br />

dy(Fx,Fy) L2 |U|<br />

dx(pUxϕ,pUx ϕ)+<br />

<br />

=|U|<br />

Putting this into (3.3), we obtain the <strong>le</strong>mma for the case ψ = ϕ. ✷<br />

dx(Fx,Fx) L 2<br />

dx(pUx ϕ,ϕ)+.


694 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Proof of Lemma 2.7. By the Riesz representation theorem and Lemmas 3.3 and 3.5, Tϕexists,<br />

the map ϕ ↦→ Tϕis linear and T 1. ✷<br />

Remark 3.6. The integral <strong>de</strong>fining T also exists in the stronger sense of Bochner integration [14,<br />

p. 132].<br />

To prove Theorem 2.8 we are going to need some more properties of the operator T .<br />

Lemma 3.7.<br />

(T ϕ, T ϕ)+ (ϕ, T ϕ)+.<br />

Proof. Using (2.6) twice, first with ψ = Tϕwe have<br />

Hence by the <strong>de</strong>finition of Fx,<br />

(T ϕ, T ϕ)+ = 1<br />

<br />

|U|<br />

(T ϕ, T ϕ)+ = 1<br />

<br />

|U|<br />

1<br />

<br />

|U|<br />

= 1<br />

<br />

|U|<br />

dx 1<br />

<br />

|U|<br />

dx 1<br />

<br />

|U|<br />

dy(pUx ϕ,pUy ϕ)+. (3.4)<br />

dy(Fx,Fy) L 2<br />

dx(Fx,Fx) L 2 by Lemma 3.4<br />

dx(pUx ϕ,ϕ)+. ✷<br />

Let T ′ : H−(Λ) → H−(Λ) be the operator <strong>du</strong>al to T on the <strong>du</strong>al space H−(Λ) of boun<strong>de</strong>d<br />

linear functionals on H+(Λ). This means<br />

〈T ′ f,ϕ〉=〈f,T ϕ〉.<br />

Lemma 3.8. The supports of T ′ f , A ′ f , Tf and A f are contained in the clo<strong>sur</strong>e of<br />

<br />

x∈supp f Ux.<br />

Proof. For A ,T this follows easily from the <strong>de</strong>finitions. Since A ′ = I − T ′ it suffices to consi<strong>de</strong>r<br />

T ′ .Letϕ be a test function with support outsi<strong>de</strong> the clo<strong>sur</strong>e of <br />

x∈supp f Ux. Then<br />

〈T ′ f,ϕ〉=〈f,T ϕ〉= 1<br />

<br />

|U|<br />

because the integrand is zero if U x ∩ supp f =∅. ✷<br />

dx〈f,pUx ϕ〉=0<br />

Likewise p ′ U : H−(Λ) → H−(Λ) is the <strong>du</strong>al of pU . Recalling the discussion following Theorem<br />

2.6, it is easy to verify that p ′ U = L pU L −1 and p ′ U is an orthogonal projection.


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 695<br />

Lemma 3.9. For the operator p ′ U <strong>de</strong>fined above we have:<br />

(1) p ′ U f = 0 if supp f ⊂ (U)c ;<br />

(2) p ′ U p′ V = 0 if U ∩ V =∅.<br />

Proof. (1) Let ϕ ∈ D(U). 〈p ′ U f,ϕ〉=〈f,pU ϕ〉=0 because supp pU ϕ ⊂ U.<br />

(2) Take the <strong>du</strong>al of Lemma 3.2. ✷<br />

Lemma 3.10. The bilinear form:<br />

(1) (f, T ′ g)−,Λ has range diam U and is positive-<strong>de</strong>finite;<br />

(2) (T ′ f,T ′ g)−,Λ has range 2diamU and is positive-<strong>de</strong>finite.<br />

Proof. (1) By using the isomorphism L and (2.6) we have<br />

(T ′ f,g)−,Λ = 1<br />

<br />

|U|<br />

dx p ′ Ux f,g<br />

<br />

1<br />

= −,Λ |U|<br />

dx p ′ Ux f,p′ Ux g<br />

−,Λ .<br />

By Lemma 3.9 the integrand vanishes at x except when supp f and supp g both intersect the<br />

clo<strong>sur</strong>e of Ux. Therefore, it is i<strong>de</strong>ntically zero if the distance between supp f and supp g is <strong>la</strong>rger<br />

than diam U.<br />

The above formu<strong>la</strong> proves that (T ′ f,f )−,Λ 0. To prove that (T ′ f,g)−,Λ is positive<strong>de</strong>finite,<br />

it suffices to prove that (T ϕ,ϕ)+ > 0 when ϕ = 0 because T ′ = L T L −1 . Since<br />

(Tϕ,ϕ)+ = 1<br />

<br />

|U|<br />

dx(pUxϕ,pUx ϕ)+<br />

it is c<strong>le</strong>ar that the right-hand si<strong>de</strong> is non-negative and vanishes iff pUx ϕ = 0 for all x. Suppose<br />

that pUx ϕ = 0 for all x.Letψ∈D(Ux) for some x. Then<br />

(ϕ, ψ)+ = (ϕ, pUx ψ)+ = (pUx ϕ,ψ)+ = 0.<br />

Therefore ϕ is orthogonal to the subspace generated by <br />

x D(Ux). By using a partition of unity<br />

any function in D(Λ) can be written as a sum of functions in <br />

x D(Ux) so ϕ is orthogonal to a<br />

<strong>de</strong>nse subset and therefore is zero.<br />

(2) (T ′ f,T ′ g)−,Λ is c<strong>le</strong>arly positive-<strong>de</strong>finite. From the proof of Lemma 3.4,<br />

(T ′ f,T ′ g)−,Λ = 1<br />

<br />

|U|<br />

dy 1<br />

<br />

|U|<br />

dx p ′ Uy f,p′ Ux g<br />

−,Λ .<br />

The rest of the argument is simi<strong>la</strong>r to (1): the clo<strong>sur</strong>es of Ux,Uy must intersect, otherwise the<br />

integrand vanishes by Lemma 3.9. Also by Lemma 3.9, Ux must intersect supp g and Uy must<br />

intersect supp f . ✷<br />

Proposition 3.11.<br />

(1) (T ′ f,T ′ f)−,Λ (f, T ′ f)−,Λ;<br />

(2) T ′ f −,Λ f −,Λ.


696 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Proof. These follow from the same properties already established for T in Lemma 3.7 and below<br />

the <strong>de</strong>finition (2.6). ✷<br />

Proof of Theorem 2.8. Since A ′ U = I − T ′ ,<br />

G1(f, f ) = G(f, f ) − G A ′ U f,A ′ <br />

U f<br />

= (f, f )−,Λ − A ′ U f,A ′ <br />

U f −,Λ<br />

= 2(f, T ′ f)−,Λ − (T ′ f,T ′ f)−,Λ<br />

(f, T ′ f)−,Λ,<br />

where the <strong>la</strong>st bound is from Proposition 3.11. Thus G1 is positive-<strong>de</strong>finite. It is finite range by<br />

Lemma 3.10. ✷<br />

Lemma 3.12.<br />

<br />

A ′ U f −,Λ f −,Λ. (3.5)<br />

Proof. By Proposition 3.11,<br />

′<br />

A U f,A ′ <br />

U f −,Λ = (f, f )−,Λ − 2(f, T ′ f)−,Λ + (T ′ f,T ′ f)−,Λ<br />

(f, f )−,Λ − (f, T ′ f)−,Λ<br />

= f,A ′ <br />

U f −,Λ .<br />

Therefore A ′ U f 2 −,Λ A ′ U f −,Λf −,Λ. ✷<br />

In the next <strong>le</strong>mma we assume that U is an open boun<strong>de</strong>d convex set that contains the origin<br />

and, for R>1 we <strong>de</strong>fine the di<strong>la</strong>ted set UR ={y: 1 R y ∈ U}. Since di<strong>la</strong>tion and trans<strong>la</strong>tion<br />

preserve convexity, H+ is continuous at x + UR. We <strong>de</strong>fine T with U rep<strong>la</strong>ced by UR.<br />

Lemma 3.13. Let A = I − T and <strong>le</strong>t Dϕ be the diameter of the support of ϕ, then there exists a<br />

constant C such that for all ϕ ∈ H+(Λ),<br />

A ϕ+ C Dϕ<br />

R ϕ+.<br />

Proof. We assume that R>Dϕ since, otherwise, the result follows from Lemma 3.12. Let V =<br />

UR and Vx = V + x. Then<br />

A ϕ = ϕ −|V | −1<br />

<br />

dxpVx∩Λϕ.<br />

By (3.2), ¯BpVx∩Λ = 1Vx ¯BpVx∩Λ. Also, since dx1Vx (y) is in<strong>de</strong>pen<strong>de</strong>nt of y and equals |V |,<br />

¯BA ϕ =|V | −1<br />

<br />

dx1Vx ¯B(I − pVx∩Λ)ϕ.


Furthermore,<br />

D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 697<br />

1Vx ¯B(I<br />

<br />

0 if supp ϕ ⊂ Vx,<br />

− pVx )ϕ =<br />

0 if supp ϕ ⊂ Rd \ V x<br />

where we used (3.2) in the second case, so that<br />

A ϕ+ =¯BA ϕL2 |V | −1<br />

<br />

dx 1Vx ¯B(I − pVx )ϕ L2 =|V | −1<br />

<br />

dx 1Vx ¯B(I − pVx )ϕ L2 X(ϕ)<br />

|V | −1 X(ϕ) ϕ+,<br />

where X(ϕ) is the set of x such that 1Vx ¯B(I − pVx∩Λ)ϕ = 0. By the previous equation X(ϕ) is<br />

contained in the set of x such that Vx intersects both supp ϕ and the comp<strong>le</strong>ment of supp ϕ. If<br />

x ∈ X(ϕ), then the smal<strong>le</strong>st ball that covers Vx neither contains nor is disjoint from the smal<strong>le</strong>st<br />

ball that covers supp ϕ. Therefore the distance between the centres of these balls lies in the<br />

interval [R − Dϕ,R+ Dϕ] and there are constants C1,C <strong>de</strong>pending on U such that<br />

|V | −1 <br />

X(ϕ) C1R −d (R + Dϕ) d − (R − Dϕ) d C Dϕ<br />

. ✷<br />

R<br />

Proof of Proposition 2.9. (1) The first item follows immediately by in<strong>du</strong>ction using Theorem<br />

2.8 with U rep<strong>la</strong>ced by Uj ,j = 1, 2,...,n.<br />

(2) Given L, <strong>le</strong>tU0 have diameter D 1 2 (1 − L−1 ). Recall from just before Proposition 2.9<br />

that f1 = f and, for j 2, fj = A ′<br />

j−1 fj−1. By in<strong>du</strong>ction using Lemma 3.8, for j 2,<br />

<br />

j−1<br />

supp fj ⊂ y: dist(y, supp f) D<br />

We find, following the proof of Theorem 2.8, that<br />

Gj (f, f ) = 2 fj ,T ′<br />

j fj<br />

<br />

−,Λ − T ′<br />

j fj ,T ′<br />

j fj<br />

<br />

−,Λ .<br />

k=1<br />

L k<br />

<br />

. (3.6)<br />

By Lemma 3.10, Gj (f, g) = 0 when the supports of fj and the analogously <strong>de</strong>fined gj are<br />

separated by at <strong>le</strong>ast 2Lj D. Therefore, Gj has range 2D j k=1 Lk Lj .<br />

(3) Recall that L ϕ = B ′ Bϕ for ϕ ∈ D(Λ) and that L is the Riesz isometry. The operators<br />

Aj are self-adjoint operators on H+(Λ). It easily follows that A ′<br />

j = LAjL −1 . Therefore<br />

<br />

A ′ ′<br />

n ...A 1f =An ...A1ϕ+ =ϕn+1+, (3.7)<br />

−,Λ<br />

where we have <strong>de</strong>fined ϕj in<strong>du</strong>ctively by setting ϕ1 = ϕ and, for j>1, ϕj = Aj−1ϕj−1. Since<br />

(3.6) says that the diameter of the support of fj is <strong>le</strong>ss than diam supp f + L j−1 and since<br />

Lemma 3.8 allows us to make exactly the same in<strong>du</strong>ction for ϕj , the diameter of the support of<br />

ϕj is boun<strong>de</strong>d by diam supp ϕ + L j−1 which, for j − 1 p, is <strong>le</strong>ss than 2L j−1 because p 0is


698 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

<strong>de</strong>fined so that diam supp ϕ L p . By Lemma 3.13, with U rep<strong>la</strong>ced by U0 and R = L j we have<br />

ϕj+1+ =Aj ϕj + CL −1 ϕj + for j − 1 p. By in<strong>du</strong>ction,<br />

ϕn+1+ C n−p L −(n−p) ϕp+<br />

for n p. Combine this with (3.7) and Lemma 3.12 to obtain<br />

<br />

′<br />

A n ...A ′<br />

1f C − n−p L −(n−p) ϕ+.<br />

Since ϕ+ =L−1f −,wehaveprovedpart(3).<br />

(4) Since L = ¯B ′ ¯B is the isomorphism from H+(Λ) to H−(Λ) and since D(Λ) is <strong>de</strong>nse in<br />

H+(Λ), thesetB ′ BD(Λ) is <strong>de</strong>nse in H−(Λ). From (3) we know that the form R <strong>de</strong>fined by<br />

R(f,f ) = G(f, f ) − <br />

Gj (f, f )<br />

vanishes for f in the <strong>de</strong>nse set B ′ BD(Λ). By(1),G(f, f ) R(f,f ) 0soR is boun<strong>de</strong>d and<br />

therefore continuous on H−(Λ). Therefore R(f,f ) = 0 for all f . ✷<br />

Proof of Lemma 2.11. Since p(x+U)∩Λϕ vanishes outsi<strong>de</strong> x + U, wehave<br />

AU ϕ = ϕ −|U| −1<br />

<br />

j1<br />

dx1x+U p(x+U)∩Λϕ =|U| −1<br />

<br />

dx1x+U (ϕ − p(x+U)∩Λϕ)<br />

which finishes the proof of (2.11). The second part of the <strong>le</strong>mma is obvious. ✷<br />

4. Examp<strong>le</strong>s<br />

In this section we apply the general theory of the previous sections to the case B ′ B =− and<br />

obtain explicit formu<strong>la</strong>s for the operator AU which is the key ingredient of our <strong>de</strong>composition.<br />

In one dimension we also obtain a formu<strong>la</strong> for AU when B ′ B is a diffusion operator.<br />

The calcu<strong>la</strong>tions in this and the next section are motivated by a <strong>de</strong>sire for insight into the<br />

smoothing properties of Ak. The bilinear form G <strong>de</strong>fines a linear operator, f ↦→ G(f, ·) from<br />

H−(Λ) to H+(Λ). Using the same <strong>le</strong>tters for forms and operators, the construction for Gj given<br />

above Proposition 2.9 is<br />

Gj = A1 ...Aj−1<br />

G − Aj GA ′<br />

j<br />

′<br />

A j−1 ...A ′<br />

1<br />

for j 2. Thinking of A ′<br />

j as A acting to the <strong>le</strong>ft we see that the integral operator kernel of G<br />

should be smoothed by the operators Ak, ifAk is smoothing. G1 = G − A1GA ′<br />

j will not be<br />

smoother than G because it inherits a singu<strong>la</strong>rity on the diagonal from G.<br />

Throughout this section we set U = BL—the ball centered at 0 and of radius L. B is the<br />

standard unit ball.<br />

We first give the formu<strong>la</strong> for the kernel of the operator A away from the boundary of the<br />

set Λ.<br />

(4.1)


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 699<br />

Proposition 4.1. Assume that L>0, d 1, B =∇, and Λ is some boun<strong>de</strong>d open set in R d ,<br />

such that the set Λ−2L ={x ∈ Λ: dist(x, ∂Λ) > 2L} is nonempty. If y ∈ Λ−2L then ALϕ(y) =<br />

d<br />

R hL(y − x)ϕ(x)dx, where<br />

hL(x) =<br />

1<br />

|B|L 2 |x| d<br />

<br />

∂BL<br />

2x · z −|x| 2 1|x−z|LσL(dz), (4.2)<br />

where σL <strong>de</strong>notes the <strong>sur</strong>face mea<strong>sur</strong>e (normalized to 1) on a sphere of radius L. In the case<br />

d = 1, σL(dx) = 1<br />

2 (δ−L(dx) + δL(dx)) where δt(dx) <strong>de</strong>notes a unit point mass at x = t.<br />

Notice that since σL is a uniform mea<strong>sur</strong>e on the sphere ∂BL the value of hL(x) <strong>de</strong>pends only<br />

on |x|.<br />

Proof. From Theorem 2.8 and Lemma 2.11, for ϕ ∈ H+(Λ) and y ∈ Λ−2L we have<br />

By trans<strong>la</strong>tion invariance,<br />

ALϕ(y) = 1<br />

|BL|<br />

ALϕ(y) = 1<br />

|BL|<br />

<br />

y+BL<br />

<br />

y+BL<br />

where ϕx(z) = ϕ(x + z).<br />

Recall the well-known formu<strong>la</strong> for the Poisson kernel of the ball BL<br />

<br />

PBLϕ(y) =<br />

∂BL<br />

Px+BL ϕ(y)dx. (4.3)<br />

PBL ϕx(y − x)dx, (4.4)<br />

Ld−2 (L2 −|y| 2 )<br />

|y − z| d<br />

ϕ(z)σL(dz).<br />

Note that this is valid also for d = 1. Applying this to (4.4), then changing the or<strong>de</strong>r of integration<br />

and substituting x = x ′ − z we obtain<br />

ALϕ(y) = 1<br />

<br />

|BL|<br />

∂BL<br />

= 1<br />

<br />

|BL|<br />

= 1<br />

|BL|<br />

<br />

=<br />

R d<br />

∂BL<br />

<br />

<br />

<br />

∂BL<br />

1x∈y+BL<br />

1x∈z+y+BL<br />

1z∈x−y+BL<br />

hL(y − x)ϕ(x)dx<br />

Ld−2 (L2 −|y − x| 2 )<br />

|y − x − z| d<br />

ϕ(x + z) dx σL(dz)<br />

Ld−2 (L2 −|y − x + z| 2 )<br />

|y − x| d<br />

ϕ(x)dx σL(dz)<br />

Ld−2 (L2 −|y − x + z| 2 )<br />

|y − x| d<br />

σL(dz)ϕ(x) dx


700 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

with<br />

hL(x) = Ld−2<br />

<br />

|BL|<br />

=<br />

∂BL<br />

1<br />

|B|L 2 |x| d<br />

because |x − z| 2 =|x| 2 − 2x · z + L 2 . ✷<br />

L 2 −|x − z| 2<br />

<br />

∂BL<br />

|x| d<br />

1|x−z|LσL(dz)<br />

2x · z −|x| 2 1|x−z|LσL(dz)<br />

The function hL of this proposition can be written in a more explicit form, namely:<br />

Proposition 4.2. In the notation of Proposition 4.1, ifd = 1 then<br />

hL(x) = 1<br />

<br />

1 −<br />

2L<br />

|x|<br />

<br />

1|x|2L. (4.5)<br />

2L<br />

If d = 2 then<br />

If d = 3 then<br />

hL(x) = 1<br />

π 2 L 2<br />

= 1<br />

π 2 L 2<br />

hL(x) =<br />

1<br />

|x|/2L<br />

2L<br />

√ 1 − r 2<br />

|x|<br />

r 2<br />

dr1|x|2L<br />

2 − 1 − arccos |x|<br />

<br />

1|x|2L. (4.6)<br />

2L<br />

3<br />

8πL|x| 2<br />

<br />

1 − |x|<br />

2 1|x|2L. (4.7)<br />

2L<br />

These expressions show that the kernel hL of AL is singu<strong>la</strong>r at x = 0, but the singu<strong>la</strong>rity is<br />

integrab<strong>le</strong>. In Section 5 we will see that AL increases the smoothness (away from the boundary)<br />

by at <strong>le</strong>ast one <strong>de</strong>rivative.<br />

Proof. Assume first that d = 1. Then the right-hand si<strong>de</strong> of (4.2) equals<br />

1<br />

|B|L 2 |x| d<br />

<br />

∂BL<br />

2<br />

2xz − x 1|x−z|LσL(dz) = 1<br />

2L2 1 2<br />

2xz − x<br />

|x| 2<br />

1|x−z|L<br />

z=±L<br />

= 1<br />

4L2 2<br />

2L|x|−|x|<br />

|x|<br />

<br />

1|x−z|L =<br />

z=±L<br />

1<br />

<br />

1 −<br />

2L<br />

|x|<br />

<br />

2L<br />

1|x|2L.<br />

Case d 2. Referring to (4.2) <strong>le</strong>t α be the ang<strong>le</strong> between x and z so that x · z = L|x| cos α<br />

and the constraint |x − z| L is equiva<strong>le</strong>nt to 0 α arccos(|x|/2L). Then


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 701<br />

hL(x) =<br />

1<br />

|B|L2 |x| d−1<br />

<br />

2L <br />

cos α −|x| 10αarccos(|x|/2L)σL(dz).<br />

For d 3 we can integrate over all points z ∈ ∂BL with α fixed. This set of points constitutes a<br />

(d − 2)-dimensional ball of radius L sin α. Keeping in mind that σL(dz) is normalized, we obtain<br />

hL(x) =<br />

2L<br />

|B|L 2 |x| d−1<br />

1<br />

cd−2<br />

<br />

arccos(|x|/2L)<br />

0<br />

<br />

cos α − |x|<br />

<br />

sin<br />

2L<br />

d−2 αdα (4.8)<br />

with cd−2 = π<br />

0 sind−2 αdα. This equation also holds for d = 2, because the normalised mea<strong>sur</strong>e<br />

on the (d − 1) = 1-dimensional sphere is σL(dz) = 1 π dα with 0 α π and cd−2 = π. Putting<br />

d = 2 into (4.8) we obtain<br />

hL(x) =<br />

=<br />

2<br />

π 2 L|x|<br />

<br />

arccos(|x|/2L)<br />

0<br />

<br />

cos α − |x|<br />

<br />

dα<br />

2L<br />

2<br />

π 2 <br />

sin<br />

L|x|<br />

arccos |x|/2L − arccos |x|/2L <br />

|x|<br />

2L<br />

= 1<br />

π 2L2 2 2L<br />

− 1 − arccos<br />

|x|<br />

|x|/2L <br />

.<br />

This finishes the proof for d = 2 since for r


702 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Proposition 4.4. Let d = 1, B = d/dx, Λ = (a, b) and L>0, then the operator AL of Theorem<br />

2.8 is of the form ALϕ(y) = <br />

(a,b) aL(y, x)ϕ(x) dx, where<br />

⎧<br />

0 for (y, x) /∈ (a, b) × (a, b),<br />

⎪⎨ y−a<br />

2L(x−a) for a


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 703<br />

ALϕ(y) = 1<br />

<br />

2L<br />

a+L<br />

a−L<br />

+ 1<br />

<br />

2L<br />

y − a<br />

ϕ(x + L)<br />

x + L − a 1|x−y|L dx<br />

b+L<br />

b−L<br />

+ 1<br />

<br />

2L<br />

b−L<br />

a+L<br />

+ 1<br />

<br />

2L<br />

= 1<br />

2L<br />

<br />

b−L<br />

a+L<br />

a+2L<br />

a<br />

+ 1<br />

2L<br />

+ 1<br />

2L<br />

b − y<br />

ϕ(x − L)<br />

b − (x − L) 1|x−y|L dx<br />

ϕ(x − L)<br />

ϕ(x + L)<br />

x + L − y<br />

1|x−y|L dx<br />

2L<br />

y − (x − L)<br />

1|x−y|L dx<br />

2L<br />

y − a<br />

ϕ(x)<br />

x − a 1yx dx + 1<br />

2L<br />

<br />

b−2L<br />

a<br />

b<br />

a+2L<br />

ϕ(x)<br />

ϕ(x)<br />

which gives (4.10).<br />

If b − a


704 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

We now proceed to the question of regu<strong>la</strong>rity of ALϕ.Letϕ ∈ H+(Λ) then, in particu<strong>la</strong>r, ϕ is<br />

a continuous function on [a,b] which vanishes at a and b. Assume first that b − a>4L. By<br />

(4.14) we obtain that for y ∈ (a + 2L,b − 2L)<br />

ALϕ(y) =<br />

y<br />

y−2L<br />

From this we obtain for y ∈ (a + 2L,b − 2L)<br />

and<br />

2L − y + x<br />

ϕ(x)<br />

(2L) 2<br />

y+2L <br />

dx +<br />

d<br />

dy ALϕ(y) = 1<br />

(2L) 2<br />

<br />

−<br />

y<br />

y−2L<br />

y<br />

ϕ(x)dx +<br />

ϕ(x)<br />

<br />

y+2L<br />

y<br />

2L + y − x<br />

(2L) 2<br />

ϕ(x)dx<br />

<br />

dx. (4.15)<br />

(4.16)<br />

d2 dy2 ALϕ(y) = 1<br />

(2L) 2<br />

<br />

ϕ(y − 2L) + ϕ(y + 2L) − 2ϕ(y) . (4.17)<br />

If y ∈ (a, a + 2L) then, again by (4.14) we have<br />

and<br />

ALϕ(y) =<br />

y<br />

a<br />

+<br />

2L − y + x<br />

ϕ(x)<br />

(2L) 2<br />

a+2L <br />

dx +<br />

<br />

y+2L<br />

a+2L<br />

ϕ(x)<br />

2L + y − x<br />

(2L) 2<br />

d<br />

dy ALϕ(y) = 1<br />

(2L) 2<br />

y<br />

− ϕ(x)dx +<br />

a<br />

<br />

a+2L<br />

y<br />

y<br />

y − a<br />

ϕ(x)<br />

2L(x − a) dx<br />

dx, (4.18)<br />

ϕ(x) 2L<br />

dx +<br />

x − a<br />

<br />

y+2L<br />

a+2L<br />

ϕ(x)dx<br />

<br />

(4.19)<br />

d2 dy2 ALϕ(y) = 1<br />

(2L) 2<br />

<br />

ϕ(y + 2L) − ϕ(y) − ϕ(y) 2L<br />

<br />

. (4.20)<br />

y − a<br />

Now it is easy to see, taking into account that ϕ(a) = 0, that if we <strong>le</strong>t y → a + 2L all the<br />

corresponding limits in (4.15)–(4.20) coinci<strong>de</strong>. Hence ALϕ is also twice continuously differentiab<strong>le</strong><br />

at y = a + 2L. The same reasoning shows that ALϕ is twice continuously differentiab<strong>le</strong><br />

in (b − 2L,b) and at b − 2L, which proves the statement of the proposition in case b − a>4L.<br />

The cases 2L b − a 4L and b − a 2L can be investigated in a simi<strong>la</strong>r way.


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 705<br />

Now assume that f ∈ H−(Λ) ∩ Cb(Λ) and b − a>4L. We will show that A ′ Lf is twice<br />

continuously differentiab<strong>le</strong> in (a + 2L,b − 2L) but does not have to be differentiab<strong>le</strong> at a + 2L.<br />

(Simi<strong>la</strong>r argument can be applied to show that the same happens at b − 2L.) Note that<br />

A ′ Lf(x)= <br />

and by (4.10) we have that for x ∈ (a + 2L,b − 2L)<br />

A ′ L f(x)=<br />

x<br />

x−2L<br />

Simi<strong>la</strong>rly as in (4.16) we obtain<br />

R<br />

2L + y − x<br />

(2L) 2<br />

f(y)dy+<br />

d<br />

dx A ′ 1<br />

Lf(x)= (2L) 2<br />

<br />

−<br />

x<br />

x−2L<br />

aL(x, y)f (y) dy<br />

<br />

x+2L<br />

x<br />

f(y)dy+<br />

2L − y + x<br />

(2L) 2<br />

f(y)dy.<br />

<br />

x+2L<br />

x<br />

<br />

f(y)dy . (4.21)<br />

C<strong>le</strong>arly, the second <strong>de</strong>rivative also exists in (a +2L,b−2L). On the other hand, if x ∈ (a, a +2L)<br />

then<br />

A ′ L f(x)=<br />

Hence for x ∈ (a, a + 2L) we have<br />

<br />

a<br />

x<br />

y − a<br />

2L(x − a) f(y)dy+<br />

a<br />

<br />

x+2L<br />

x<br />

2L − y + x<br />

(2L) 2<br />

f(y)dy.<br />

d<br />

dx A ′ 1<br />

Lf(x)= (2L) 2<br />

x<br />

x+2L <br />

2L(y − a)<br />

−<br />

f(y)dy+ f(y)dy . (4.22)<br />

(x − a) 2<br />

Letting x → a + 2L in (4.21) and (4.22) we see that <strong>le</strong>ft and right <strong>de</strong>rivatives of A ′ Lf at a + 2L<br />

are equal if and only if<br />

which is not true in general. ✷<br />

<br />

a+2L<br />

a<br />

f(y)dy=<br />

<br />

a+2L<br />

a<br />

y − a<br />

2L f(y)dy,<br />

Proposition 4.4 can be exten<strong>de</strong>d to the case when the operator −B ′ B generates a onedimensional<br />

non-<strong>de</strong>generate diffusion process. Operator AL can be written in terms of the sca<strong>le</strong><br />

function of the diffusion process.<br />

x


706 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Proposition 4.6. Assume that d = 1, Λ = (a, b), B = c(x) d<br />

, where c is twice continuously<br />

dx<br />

differentiab<strong>le</strong> and c2 (x) const > 0. Let s <strong>de</strong>note the sca<strong>le</strong> function of the diffusion process<br />

generated by −B ′ B. Then the kernel of the operator AL is of the form:<br />

⎧<br />

s(y)−s(a)<br />

2L(s(x)−s(a))<br />

s(b)−s(y)<br />

for a


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 707<br />

which coinci<strong>de</strong>s with (4.23). If b − a


708 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

Proposition 5.2. For each ϕ ∈ H+(R d ) the limit<br />

exists in H+(Λ) and<br />

Aϕ = lim<br />

n→∞ A L −n ...A L −2A L −1A1ϕ (5.4)<br />

A ϕ = h ∗ ϕ, (5.5)<br />

where h ∈ C∞ 0 (Rd ) and the support of h is contained in |x| DL<br />

L−1 . Moreover,<br />

G(f, g) = Γ0(f, g) + G(A ′ f,A ′ g),<br />

d<br />

f,g ∈ H− R , (5.6)<br />

where Γ0 is a positive-<strong>de</strong>finite operator of finite range smal<strong>le</strong>r than 2D + 2DL/(L − 1).<br />

By rep<strong>la</strong>cing A by A in the construction <strong>de</strong>scribed above Proposition 2.9, we have<br />

Corol<strong>la</strong>ry 5.3. For D>0 there exists c(D) such that for each L c(D) the iteration of (5.6)<br />

generates C ∞ finite range <strong>de</strong>compositions of G.<br />

Proof of Proposition 5.2. Using the results of the previous section we have ALϕ = hL ∗ ϕ with<br />

hL given by (4.2). Hence<br />

A L −n ...A L −2A L −1A1ϕ = h L −n ∗···∗h L −1 ∗ h1 ∗ ϕ.<br />

Notice that by the <strong>de</strong>finition of AL as an average of Poisson kernels we have that each hL<br />

integrates to 1.<br />

Let X0,X1,... be in<strong>de</strong>pen<strong>de</strong>nt random variab<strong>le</strong>s with the same <strong>de</strong>nsity h1. Then, by (4.9)<br />

L −k Xk has <strong>de</strong>nsity h L −k . It is c<strong>le</strong>ar that ∞ k=0 L −k Xk converges almost <strong>sur</strong>ely to some random<br />

variab<strong>le</strong> Y which has a <strong>de</strong>nsity. Let us <strong>de</strong>note it by h. It is also c<strong>le</strong>ar that Y is boun<strong>de</strong>d by<br />

DL/(L − 1), i.e. support of h is the ball centered at 0 and of radius DL/(L − 1). Notice that<br />

mk=0 L −k Xk has <strong>de</strong>nsity ˜hm = h1 ∗···∗h L −m. Choose m such that<br />

m <br />

ˆhk(x) CL,n<br />

, (5.7)<br />

|x| d+2<br />

k=1<br />

we can do it by (4.9) and Lemma 5.1. (5.7) implies that ˜hm is continuously differentiab<strong>le</strong>.<br />

From the <strong>de</strong>finition of h it follows that<br />

h(x) = ˜hm ∗ L −(m+1)d h L −(m+1) · (x) (5.8)<br />

and since ˜hm is continuously differentiab<strong>le</strong>, (5.8) implies that h is smooth.<br />

For n m we can write<br />

<br />

h1 ∗···∗hL−n(x) = E ˜hm x −<br />

n<br />

k=m+1<br />

L −k Xk<br />

<br />

.


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 709<br />

By continuity of ˜hm we have<br />

lim<br />

n→∞ h1<br />

<br />

∗···∗hL−n(x) = E ˜hm<br />

−(m+1)<br />

x − L Y = h(x),<br />

where the <strong>la</strong>st equality follows by (5.8).<br />

In a very simi<strong>la</strong>r way, possibly choosing <strong>la</strong>rger m, we can show that also the <strong>de</strong>rivatives of<br />

h1 ∗···∗hn converge to the <strong>de</strong>rivatives of h and are boun<strong>de</strong>d. From this we obtain that for any<br />

ϕ ∈ D(Rd ) we have (5.4) with Aϕ = h ∗ ϕ. Moreover, since for each n we have AL−n 1,<br />

(5.4) holds for any ϕ ∈ H+(Rd ) and A 1.<br />

To prove the other part of the theorem, <strong>le</strong>t us <strong>de</strong>note<br />

G−k(f, g) = G(f, g) − G A ′<br />

L−k f,A ′<br />

L−kf d<br />

, f,g∈H−R , (5.9)<br />

and<br />

˜G−n(f, g) = G(f, g) − G A ′<br />

1 ...A ′ ′<br />

L−nf,A 1 ...A ′ L−ng d<br />

, f,g∈H−R . (5.10)<br />

Using (5.9) we can write<br />

˜G−n(f, g) = G−n(f, g) +<br />

n<br />

k=1<br />

′<br />

G−(n−k) A<br />

L−(n−k+1) ...A ′ ′<br />

L−nf,A L−(n−k+1) ...A ′ L−ng . (5.11)<br />

From Theorem 2.8 it follows that G−k is positive-<strong>de</strong>finite for each k, hence so is ˜G−n.Fromthe<br />

same theorem we have that range G−k is 2DL −k and we also know that<br />

diam supp A ′<br />

L −kϕ = diam supp ϕ + 2DL −k .<br />

Combining these two facts with (5.11) we obtain that ˜G−n has range smal<strong>le</strong>r than 2D +<br />

2DL/(L − 1). By (5.4)<br />

G(f, g) − G(A ′ f,A ′ G) = lim ˜G−n(f, g).<br />

n→∞<br />

Hence G(·,·) − G(A ′ ·,A ′ ·) is positive-<strong>de</strong>finite and of range 2D + 2DL/(L − 1). ✷<br />

Proof of Lemma 5.1. By (4.2) we have<br />

ˆh(y) =<br />

<br />

1<br />

<br />

e<br />

|x|2<br />

ix·y<br />

|B||x| d<br />

∂B<br />

σ(dz) 2x · z −|x| 2 1 2x·z|x| 2.<br />

Function h is symmetric, hence ˆh is real, so we can rep<strong>la</strong>ce e ix·y by cos(x · y). We change the<br />

or<strong>de</strong>r of integration and substitute x = rw where r =|x| and w = x/|x| to obtain


710 D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711<br />

ˆh(y) = |∂B|<br />

<br />

|B|<br />

∂B<br />

∂B<br />

<br />

σ(dz)<br />

∂B<br />

∂B<br />

σ(dw)1w·z>0<br />

<br />

2(w·z)<br />

0<br />

dr cos(rw · y) (2w · z) − r <br />

= |∂B|<br />

<br />

1 − cos(2(w · y)(w · z))<br />

σ(dz) σ(dw)1w·z>0<br />

|B|<br />

(w · y) 2<br />

.<br />

By symmetry with respect to z we obtain<br />

ˆh(y) = |∂B|<br />

<br />

|B|<br />

<br />

σ(dz)<br />

1 − cos(2(w · y)(w · z))<br />

σ(dw)<br />

2(w · y) 2<br />

,<br />

which is equal to (5.1).<br />

The estimate (5.2) is now obvious since by (5.1)<br />

∂B<br />

∂B<br />

ˆh(y) = sin2 y<br />

y 2<br />

( ˆh(y) can be also easily computed directly using (4.5)).<br />

The proof of (5.3) for d 2 is a litt<strong>le</strong> more involved. C<strong>le</strong>arly ˆh is boun<strong>de</strong>d since h is integrab<strong>le</strong>.<br />

Now assume that |y| > 1 and consi<strong>de</strong>r first the case d = 2. Let α be the ang<strong>le</strong> between w and z<br />

and β the ang<strong>le</strong> between y and w, then, using symmetries in (5.1), we have<br />

<br />

ˆh(y) = C<br />

π/2<br />

0<br />

<br />

dβ<br />

π/2<br />

Next we substitute u = cos α, v = cos β to obtain<br />

1<br />

<br />

ˆh(y) = C dv<br />

C1<br />

0<br />

<br />

1/2<br />

0<br />

0<br />

1<br />

0<br />

0<br />

dα sin2 (|y| cos β cos α)<br />

|y| 2 cos2 .<br />

β<br />

<strong>du</strong> sin2 (|y|u)<br />

|y| 2 v 2<br />

1<br />

√<br />

1 − v2 √ 1 − u2 1<br />

1<br />

dv <strong>du</strong><br />

1 + (|y|uv) 2<br />

u2 √<br />

1 − u2 + C1<br />

|y| 2<br />

1<br />

1/2<br />

0<br />

1<br />

1<br />

dv <strong>du</strong> √<br />

1 − v2 √ 1 − u2 0<br />

0<br />

C1<br />

∞<br />

1<br />

1<br />

dv <strong>du</strong><br />

|y|<br />

1 + v2 u C2 C3<br />

√ + <br />

1 − u2 |y| 2 |y| ,<br />

which proves (5.3) for d = 2.<br />

The proof for d 3 is easier. We start in the same way obtaining


D. Brydges, A. Ta<strong>la</strong>rczyk / Journal of Functional Analysis 236 (2006) 682–711 711<br />

ˆh(y) = Cd,1<br />

Cd,1<br />

Cd,2<br />

Cd,2<br />

|y|<br />

<br />

π/2<br />

0<br />

1<br />

0<br />

1<br />

0<br />

<br />

0<br />

<br />

dβ<br />

π/2<br />

0<br />

<br />

dv<br />

0<br />

0<br />

1<br />

dα sin2 (|y| cos β cos α)<br />

|y| 2 cos 2 β<br />

<strong>du</strong> sin2 (|y|uv)<br />

|y| 2 v 2<br />

1<br />

1<br />

dv <strong>du</strong><br />

u2<br />

1 + (|y|uv) 2<br />

∞<br />

1<br />

dv<br />

1 + v2 which finishes the proof of (5.3). ✷<br />

Acknow<strong>le</strong>dgments<br />

1<br />

0<br />

u<strong>du</strong>,<br />

sin d−2 α sin d−2 β<br />

The work of DB was supported in part by NSERC of Canada. AT is grateful to the University<br />

of British Columbia, Vancouver, Canada, for hospitality.<br />

References<br />

[1] G. Benfatto, N. Cassandro, G. Gal<strong>la</strong>votti, F. Nicolo, E. Olivieri, E. Presutti, E. Scacciatelli, Some probabilistic<br />

techniques in field theory, Comm. Math. Phys. 59 (1978) 143–166.<br />

[2] G. Benfatto, N. Cassandro, G. Gal<strong>la</strong>votti, F. Nicolo, E. Olivieri, E. Presutti, E. Scacciatelli, On the ultravio<strong>le</strong>tstability<br />

in the Eucli<strong>de</strong>an sca<strong>la</strong>r field theories, Comm. Math. Phys. 71 (1980) 95–130.<br />

[3] D.C. Brydges, P.K. Mitter, B. Scoppo<strong>la</strong>, Critical (Φ4 )3,ɛ, Comm. Math. Phys. 240 (1–2) (2003) 281–327.<br />

[4] D. Brydges, G. Guadagni, P.K. Mitter, Finite range <strong>de</strong>composition of Gaussian processes, J. Statist. Phys. 115 (1–2)<br />

(2004) 415–449, http://arxiv.org/abs/math-ph/0303013.<br />

[5] P. Fe<strong>de</strong>rbush, Quantum field theory in 90 minutes, Bull. Amer. Math. Soc. 17 (1987) 93–103.<br />

[6] K. Gawedzki, A. Kupiainen, A rigorous block spin approach to mass<strong>le</strong>ss <strong>la</strong>ttice theories, Comm. Math. Phys. 77 (1)<br />

(1980) 31–64.<br />

[7] K. Gawedzki, A. Kupiainen, Mass<strong>le</strong>ss <strong>la</strong>ttice ϕ4 4 theory: Rigorous control of a renormalizab<strong>le</strong> asymptotically free<br />

mo<strong>de</strong>l, Comm. Math. Phys. 99 (2) (1985) 197–252.<br />

[8] C. Hainzl, R. Seiringer, General <strong>de</strong>composition of radial functions on Rn and applications to N-body quantum<br />

systems, Lett. Math. Phys. 61 (1) (2002) 75–84.<br />

[9] K. Itô, Foundations of Stochastic Differential Equations in Infinite-Dimensional Spaces, CBMS–NSF Regional<br />

Conf. Ser. in Appl. Math., vol. 47, SIAM, Phi<strong>la</strong><strong>de</strong>lphia, PA, 1984.<br />

[10] M. Reed, B. Simon, Methods of Mo<strong>de</strong>rn Mathematical Physics. I. Functional Analysis, Aca<strong>de</strong>mic Press, New York,<br />

1972.<br />

[11] M. Reed, B. Simon, Methods of Mo<strong>de</strong>rn Mathematical Physics. II. Fourier Analysis, Self-adjointness, Aca<strong>de</strong>mic<br />

Press, New York, 1975.<br />

[12] K.G. Wilson, J. Kogut, The renormalization group and the ɛ expansion, Phys. Rep. (Sect. C of Phys. Lett.) 12 (1974)<br />

75–200.<br />

[13] K.G. Wilson, The renormalization group and critical phenomena, Rev. Mo<strong>de</strong>rn Phys. 55 (3) (1983) 583–600.<br />

[14] K. Yosida, Functional Analysis, sixth ed., Springer-Ver<strong>la</strong>g, Berlin, 1980.


Journal of Functional Analysis 236 (2006) 712–725<br />

www.elsevier.com/locate/jfa<br />

Weak <strong>semi</strong>-continuity of the <strong>du</strong>ality pro<strong>du</strong>ct<br />

in Sobo<strong>le</strong>v spaces<br />

Dorin Bucur<br />

Département <strong>de</strong> Mathématiques, UMR-CNRS 7122, Université <strong>de</strong> Metz, I<strong>le</strong> <strong>du</strong> Saulcy, 57045 Metz Ce<strong>de</strong>x 01, France<br />

Received 20 February 2006; accepted 13 March 2006<br />

Avai<strong>la</strong>b<strong>le</strong> online 2 May 2006<br />

Communicated by Paul Malliavin<br />

Abstract<br />

Given a weakly convergent sequence of positive functions in W 1,p<br />

0 (Ω), we prove the equiva<strong>le</strong>nce between<br />

its convergence in the sense of obstac<strong>le</strong>s and the lower <strong>semi</strong>-continuity of the term by term <strong>du</strong>ality<br />

pro<strong>du</strong>ct associated to (the p-Lap<strong>la</strong>cian of) weakly convergent sequences of p-superharmonic functions of<br />

W 1,p<br />

0 (Ω). This result implicitly gives new characterizations for both the convergence in the sense of obstac<strong>le</strong>s<br />

of a weakly convergent sequence of positive functions and for the weak l.s.c. of the <strong>du</strong>ality pro<strong>du</strong>ct.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Duality pro<strong>du</strong>ct; γ -Convergence; Obstac<strong>le</strong> convergence<br />

1. Intro<strong>du</strong>ction<br />

The limit of the sca<strong>la</strong>r pro<strong>du</strong>ct of two weakly convergent sequences in a Hilbert space is,<br />

a priori, uncontrol<strong>la</strong>b<strong>le</strong>. In some particu<strong>la</strong>r situations, as for examp<strong>le</strong> in Sobo<strong>le</strong>v spaces, when the<br />

sequences of functions are solutions (or supersolutions) of partial differential equations, extra information<br />

can be obtained on the sca<strong>la</strong>r pro<strong>du</strong>ct of the limits by using qualitative properties of the<br />

solutions of the PDEs. In this paper, we are interested in <strong>du</strong>ality pro<strong>du</strong>cts in the Sobo<strong>le</strong>v spaces<br />

W −1,q × W 1,p<br />

0 involving p-superharmonic and positive functions. We characterize all sequences<br />

of positive functions, such that the <strong>du</strong>ality pro<strong>du</strong>ct with the p-Lap<strong>la</strong>cian of p-superharmonic<br />

functions is lower <strong>semi</strong>-continuous.<br />

E-mail address: bucur@math.univ-metz.fr.<br />

0022-1236/$ – see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.jfa.2006.03.014


D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 713<br />

More precisely, <strong>le</strong>t (un), (vn) ⊆ W 1,p<br />

0 (Ω) be two sequences of non-negative functions and<br />

assume they weakly converge to u and v, respectively. Assuming moreover that −pun 0in<br />

the sense of distributions on Ω, we won<strong>de</strong>r whether<br />

<br />

lim inf<br />

n→∞<br />

Ω<br />

|∇un| p−2 <br />

∇un∇vn dx <br />

Ω<br />

|∇u| p−2 ∇u∇vdx. (1)<br />

Un<strong>de</strong>r no further assumption, this assertion is in general false. For an extensive study of this<br />

question we refer the rea<strong>de</strong>r to [3] (see also [4] for further results), where the authors give sufficient<br />

conditions on the sequence (vn) in or<strong>de</strong>r that (1) holds true. Precisely, the main hypothesis<br />

is that the functions vn are also p-superharmonic −pvn 0. An examp<strong>le</strong> showing that the<br />

positivity condition vn 0 is not (in general) sufficient for the lower <strong>semi</strong>-continuity of (1) is<br />

also given for p = 2. The examp<strong>le</strong> relies on the emergence of the “strange term” appearing in<br />

the re<strong>la</strong>xation process of domains through γ -convergence (see [10]) and gives an intuitive hint<br />

on the fact that, in or<strong>de</strong>r that (1) is true, (vn) should vary such that their <strong>le</strong>vel sets do not pro<strong>du</strong>ce<br />

re<strong>la</strong>xation mea<strong>sur</strong>es via γ -convergence (see [5] for a <strong>de</strong>tai<strong>le</strong>d intro<strong>du</strong>ction to γ -convergence and<br />

Section 2 for a short review).<br />

The purpose of this paper is to give a characterization of all W 1,p<br />

0 (Ω)-weakly convergent<br />

sequences (vn) of non-negative functions for which (1) holds true for every W 1,p<br />

0 (Ω)-weakly<br />

convergent sequence (un) such that −pun 0.<br />

We prove that the necessary and sufficient condition that (vn) has to satisfy is to converge<br />

in the sense of obstac<strong>le</strong>s (see Section 2 for the precise <strong>de</strong>finition and [1,12] for <strong>de</strong>tails). This<br />

convergence is, in a certain sense, weaker than the strong convergence of W 1,p<br />

0 (Ω) and stronger<br />

than the weak convergence of W 1,p<br />

0 (Ω). Since (vn) is assumed by hypothesis weakly convergent<br />

in W 1,p<br />

0 (Ω), proving that it also converges in the sense of obstac<strong>le</strong>s is equiva<strong>le</strong>nt to the possibility<br />

of finding a sequence θn strongly convergent to v in W 1,p<br />

0 (Ω), such that θn vn a.e. This is a<br />

consequence of the characterization of the obstac<strong>le</strong> convergence via the Mosco convergence of<br />

the convex sets<br />

Kvn = u ∈ W 1,p<br />

0 (Ω): u vn a.e. .<br />

In or<strong>de</strong>r to <strong>de</strong>scribe the obstac<strong>le</strong>s, we make use on fine quasi-continuity properties of Sobo<strong>le</strong>v<br />

functions. The proof of the main result of the paper relies on the characterization of the obstac<strong>le</strong><br />

obst<br />

γ<br />

convergence vn −→ v in terms of the γ -convergence of the <strong>le</strong>vel sets {vn >t} −→ {v>t}, and<br />

on the know<strong>le</strong>dge of the re<strong>la</strong>xed mea<strong>sur</strong>es associated to a γ -convergent sequence of quasi-open<br />

sets.<br />

Of course, the difficult part is the necessity. In [3], the hypothesis on the p-superharmonicity<br />

of vn in<strong>sur</strong>es the fact that min{vn,v} converges strongly to v in W 1,p<br />

0 (Ω)! So, taking θn =<br />

min{vn,v} we recover the obstac<strong>le</strong> convergence and fall into the sufficient part of the characterization<br />

result.<br />

All results in this paper hold for A-superharmonic functions, where −div A is a non-linear<br />

operator of p-Lap<strong>la</strong>cian type. Precisely, assuming A : W 1,p<br />

0 (Ω) ↦→ W −1,q (Ω) is simi<strong>la</strong>r to the<br />

p-Lap<strong>la</strong>cian (see the exact <strong>de</strong>finition in Section 2) one can prove that if (vn) ⊆ W 1,p<br />

0 (Ω) is a<br />

weakly convergent sequence of non-negative functions, then vn converges in the sense of obsta-<br />

c<strong>le</strong>s to the same limit v if and only if for every sequence of functions (un) ⊆ W 1,p<br />

0 (Ω), such that<br />

−div(a(x, ∇un)) 0 and un ⇀uweakly in W 1,p<br />

0 (Ω) we have


714 D. Bucur / Journal of Functional Analysis 236 (2006) 712–725<br />

<br />

lim inf<br />

n→∞<br />

Ω<br />

<br />

a(x,∇un)∇vn dx <br />

Ω<br />

a(x,∇u)∇vdx. (2)<br />

For the simplicity of the exposition, our results are presented for the p-Lap<strong>la</strong>ce operator. We<br />

point out the fact that the convergence of (vn) into the sense of obstac<strong>le</strong>s is in<strong>de</strong>pen<strong>de</strong>nt on the<br />

choice of the operator −div(a(x, ·)).<br />

Section 2 contains a review of the main tools used in the paper, Section 3 contains the proof of<br />

the characterization result and the <strong>la</strong>st section is <strong>de</strong>voted to some examp<strong>le</strong>s. A particu<strong>la</strong>r attention<br />

is given to uniformly oscil<strong>la</strong>ting obstac<strong>le</strong>s.<br />

2. Obstac<strong>le</strong>s, capacity and γ -convergence<br />

2.1. Capacity and re<strong>la</strong>xation mea<strong>sur</strong>es<br />

is<br />

Let Ω ⊆ R N be a boun<strong>de</strong>d open set and <strong>le</strong>t 1 0 there exists a continuous<br />

(respectively lower <strong>semi</strong>-continuous) function fɛ : Ω → R such that capp ({f = fɛ},Ω)


D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 715<br />

(i) ∞|E(B) = 0 for every Borel set B ⊆ Ω with cap p (B ∩ E,Ω) = 0;<br />

(ii) ∞|E(B) =+∞for every Borel set B ⊆ Ω with cap p (B ∩ E,Ω) > 0.<br />

Definition 2.1. We say that a sequence (μn) of mea<strong>sur</strong>es in M p<br />

0 (Ω) γp-converges to a mea<strong>sur</strong>e<br />

μ ∈ M p<br />

1,p<br />

0 (Ω) if and only if Fμn : W0 (Ω) → R,<br />

Fμn (u) =<br />

Γ -converges in Lp (Ω) to Fμ, where<br />

<br />

Fμ(u) =<br />

<br />

Ω<br />

Ω<br />

|∇u| p <br />

dx +<br />

Ω<br />

|∇u| p <br />

dx +<br />

Ω<br />

|u| p dμn<br />

|u| p dμ.<br />

In or<strong>de</strong>r to simplify notations and since p is fixed, we drop the in<strong>de</strong>x p and instead of γp we<br />

note γ .<br />

We recall that Fn : W 1,p<br />

0 (Ω) → R Γ -converges to F in Lp (Ω) if for every u ∈ Lp (Ω) there<br />

exists a sequence un ∈ Lp (Ω) such that un → u in Lp (Ω) and<br />

Fμ(u) lim sup Fμn<br />

n→∞<br />

(un),<br />

and for every convergent sequence un → u in L p (Ω)<br />

Fμ(u) lim inf Fμn (un).<br />

n→∞<br />

In Definition 2.1, by the i<strong>de</strong>ntification of a quasi-open set A with the mea<strong>sur</strong>e ∞Ω\A, we<br />

implicitly have the <strong>de</strong>finition of the γ -convergence of a sequence of quasi-open sets. In general,<br />

the γ -limit of a sequence of quasi-open sets is a mea<strong>sur</strong>e of M p<br />

0 (Ω). In particu<strong>la</strong>r, this mea<strong>sur</strong>e<br />

can be itself of the form ∞Ω\A.<br />

Note that the γ -convergence is metrizab<strong>le</strong> by the following distance:<br />

<br />

dp(μ1,μ2) = |wμ1 − wμ2 | dx,<br />

where wμ is the variational solution of<br />

Ω<br />

−pwμ + μ|wμ| p−2 wμ = 1 (3)<br />

in W 1,p<br />

0 (Ω) ∩ Lp (Ω, μ) (see [5,15]). The precise sense of the equation is the following: wμ ∈<br />

W 1,p<br />

0 (Ω) ∩ Lp (Ω, μ) and for every φ ∈ W 1,p<br />

0 (Ω) ∩ Lp (Ω, μ)<br />

<br />

Ω<br />

|∇wμ| p−2 <br />

∇wμ∇φdx+<br />

Ω<br />

|wμ| p−2 <br />

wμφdμ=<br />

Ω<br />

φdx.


716 D. Bucur / Journal of Functional Analysis 236 (2006) 712–725<br />

In view of the result of Hedberg [16], if A is an open subset of Ω, the solution of this equation<br />

associated to the mea<strong>sur</strong>e ∞Ω\A is nothing else but the solution in the sense of distributions of<br />

−pw = 1 inA, w ∈ W 1,p<br />

0 (A).<br />

Throughout the paper, by wμ we <strong>de</strong>note the solution of (3) associated to the mea<strong>sur</strong>e μ, and<br />

by wA the solution of the same equation associated to the mea<strong>sur</strong>e ∞Ω\A.<br />

We refer to [14] for the following result.<br />

Proposition 2.2. The space M p<br />

0 (Ω), endowed with the distance dp, is a compact metric space.<br />

Moreover, the c<strong>la</strong>ss of mea<strong>sur</strong>es of the form ∞Ω\A, with A open (and smooth) subset of Ω, is<br />

<strong>de</strong>nse in M p<br />

0 (Ω).<br />

Given a mea<strong>sur</strong>e μ ∈ M p<br />

0 (Ω), we call the regu<strong>la</strong>r set of the mea<strong>sur</strong>e μ the quasi-open set<br />

{wμ > 0}. We also notice that this set, which is <strong>de</strong>noted Aμ, coinci<strong>de</strong>s up to a set of zero capacity<br />

with the union of all finely open sets of finite μ mea<strong>sur</strong>e.<br />

Lemma 2.3. Assume (An), (Bn) are two sequences of quasi-open sets which γ -converge to μA,<br />

μB, respectively. If capp (An ∩ Bn) = 0 for every n ∈ N, then cap(AμA ∩ AμB ) = 0.<br />

Proof. We notice that wAn · wBn = 0 q.e. Passing to the limit a.e. and using the quasi-continuity<br />

of wμA and wμB , we get wμA · wμB = 0 q.e., hence the conclusion. ✷<br />

On M p<br />

0 (Ω), the following monotonicity is consi<strong>de</strong>red on the family of mea<strong>sur</strong>es:<br />

μ1 μ2 if ∀A ⊆ Ω, A quasi-open, μ1(A) μ2(A).<br />

We notice (see [5,13]) that every monotone (increasing or <strong>de</strong>creasing) sequence of mea<strong>sur</strong>es is<br />

γ -convergent.<br />

2.2. The obstac<strong>le</strong> prob<strong>le</strong>m<br />

Although the obstac<strong>le</strong> prob<strong>le</strong>m can be <strong>de</strong>fined properly in the frame of mea<strong>sur</strong>ab<strong>le</strong> or quasilower<br />

<strong>semi</strong>-continuous functions, in the most part of the paper we restrict ourselves to obstac<strong>le</strong>s<br />

which are e<strong>le</strong>ments of W 1,p<br />

0 (Ω). Roughly speaking, for q = p/(p − 1), given a function f ∈<br />

W −1,q (Ω) and a mea<strong>sur</strong>ab<strong>le</strong> function v, the solution of the obstac<strong>le</strong> prob<strong>le</strong>m associated to v and<br />

f is the unique solution of the minimization prob<strong>le</strong>m<br />

where<br />

<br />

1<br />

min<br />

p<br />

Ω<br />

|∇φ| p <br />

−〈f,φ〉<br />

W −1,q 1,p : φ ∈ Kv ,<br />

(Ω)×W0 (Ω)<br />

Kv = φ ∈ W 1,p<br />

0 (Ω): φ v a.e. Ω .


D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 717<br />

If the obstac<strong>le</strong> v is an e<strong>le</strong>ment of W 1,p<br />

0 (Ω), the inequality φ v in the previous set can be<br />

equiva<strong>le</strong>ntly taken in the sense a.e. or p-q.e. (for quasi-continuous representatives). In the sequel<br />

we concentrate our attention only on obstac<strong>le</strong>s belonging to W 1,p<br />

0 (Ω).<br />

Definition 2.4. Let vn,v∈ W 1,p<br />

0 (Ω). We say that vn converges in the sense of obstac<strong>le</strong>s to v if<br />

Γ -converges in L p (Ω) to<br />

W 1,p<br />

<br />

0 (Ω) ∋ u ↦→<br />

Ω<br />

W 1,p<br />

<br />

0 (Ω) ∋ u ↦→<br />

where ∞{v}(u) = 0ifu vp-q.e. and +∞ if not. We write vn<br />

Ω<br />

|∇u| p dx +∞{vn}(u) (4)<br />

|∇u| p dx +∞{v}(u), (5)<br />

obst<br />

−→ v.<br />

obst<br />

The main consequence of the convergence of obstac<strong>le</strong>s vn −→ v is that for every f ∈<br />

W −1,q (Ω) the sequence of solutions un of the obstac<strong>le</strong> prob<strong>le</strong>m associated to vn and f converges<br />

strongly in W 1,p<br />

0 (Ω) to the solution associated to v.<br />

The convergence in the sense of obstac<strong>le</strong>s is equiva<strong>le</strong>nt to the convergence in the sense of<br />

Mosco of Kvn to Kv (see [1]), i.e. to the following two re<strong>la</strong>tions:<br />

1. ∀u ∈ Kv, ∃un ∈ Kvn such that un → u strongly in W 1,p<br />

0 (Ω).<br />

2. If unk ∈ Kvn<br />

1,p<br />

and unk ⇀uweakly in W<br />

k 0 (Ω), then u ∈ Kv.<br />

Notice that if vn converges weakly in W 1,p<br />

0 (Ω) to v, then the second Mosco condition is satisfied.<br />

We recall from [8, Corol<strong>la</strong>ry 4.9] and [12] the following characterization of the obstac<strong>le</strong> convergence.<br />

Theorem 2.5. Let vn,v ∈ W 1,p<br />

0 (Ω), vn 0. Then vn converges in the sense of obstac<strong>le</strong>s to v if<br />

and only if there exists a <strong>de</strong>nse set T ⊆ R such that<br />

Operators simi<strong>la</strong>r to the p-Lap<strong>la</strong>cian<br />

γ<br />

{vn >t} −→ {v>t} ∀t∈T. Assume that a : Ω × R N → R N is a Carathéodory function which is homogeneous of <strong>de</strong>gree<br />

p − 1 in the second variab<strong>le</strong><br />

a(x,tζ) =|t| p−2 ta(x, ζ ), t ∈ R, t= 0. (6)<br />

We notice that in or<strong>de</strong>r to have a precise <strong>de</strong>scription of the γ -limits of sequences of quasi-open<br />

sets, the homogeneity property is crucial (see [9,15]).


718 D. Bucur / Journal of Functional Analysis 236 (2006) 712–725<br />

We suppose as usual the monotonicity assumptions on a(x,ζ) (see for instance [15]): there exist<br />

two constants c0,c1 with 0


D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 719<br />

<br />

= lim inf<br />

n→∞<br />

Ω<br />

|∇un| p−2 <br />

∇un∇θn dx =<br />

Ω<br />

|∇u| p−2 ∇u∇vdx.<br />

For the <strong>la</strong>st equality, we notice, on the one hand, that |∇un| p−2∇un converges weakly to<br />

|∇u| p−2∇u in Lq (Ω, RN ) since −pun are positive and uniformly boun<strong>de</strong>d Radon mea<strong>sur</strong>es.<br />

This result is <strong>du</strong>e to Boccardo and Murat [2] (see also [15, Theorem 2.10]) and is re<strong>la</strong>ted to the<br />

pointwise convergence of the <strong>gradient</strong>s. On the second hand, we have that θn converges strongly<br />

in W 1,p<br />

0 (Ω).<br />

Sufficiency: (ii) ⇒ (i). Assume (ii) holds. We shall prove that vn converges in the sense of<br />

obstac<strong>le</strong>s to v. We rely both on the characterization of the obstac<strong>le</strong> convergence via the Mosco<br />

convergence, and since vn 0, on the characterization through the γ -convergence of the <strong>le</strong>velsets<br />

{vn >t}, Section 2, Theorem 2.5.<br />

We notice that from the weak convergence in W 1,p<br />

0 (Ω), vn ⇀v, the second Mosco condition<br />

M<br />

of the <strong>de</strong>sired Mosco convergence Kvn −→ Kv is automatically satisfied. Moreover, the first<br />

Mosco condition has to be proved only for the function v. In<strong>de</strong>ed, if there exists θn vn such<br />

that θn → v in W 1,p<br />

0 (Ω)-strong, then for every ψ v, the first Mosco condition holds with the<br />

sequence min{θn,ψ} vn.<br />

Assume for contradiction that ∃δ>0 and a subsequence (vnk ) such that<br />

lim inf<br />

k→∞ min u − v 1,p δ>0. (13)<br />

u∈Kvn<br />

W0 (Ω)<br />

k<br />

In or<strong>de</strong>r to achieve the contradiction in assumption (13), we construct a subsequence of (vnk ),<br />

which converges in the sense of obstac<strong>le</strong>s to v.<br />

Step 1. We construct a mapping<br />

[0, +∞) ↦→ μt ∈ M p<br />

0 (Ω),<br />

which is increasing in the sense of mea<strong>sur</strong>es of M p<br />

0 (Ω) and such that for a subsequence of (vnk )<br />

(<strong>de</strong>noted with the in<strong>de</strong>x r) and for a <strong>de</strong>nse set T ⊆[0, +∞) we have<br />

∀t ∈ T {vr >t} −→ μt.<br />

The construction of the mapping t ↦→ μt is done by a diagonal proce<strong>du</strong>re using the compactness<br />

and the metrizability of the γ -convergence in M p<br />

0 (Ω). We follow the approach used in [7] for<br />

constructing the re<strong>la</strong>xed space for obstac<strong>le</strong>s. Let T = Q ∩ R + ={t1,t2,...,tk,...}. Forr = 1,<br />

..., we successively extract γ -convergent subsequences for the sequences of the quasi-open sets<br />

{vnk >tr}, and <strong>de</strong>fine μtr being the γ -limit. By a diagonal proce<strong>du</strong>re, the metrizability of the<br />

γ -convergence gives the existence of a subsequence of (vn) and of a family of mea<strong>sur</strong>es (μtr )<br />

such that the <strong>le</strong>vel sets {vnk >tr} γ -converges to μtr .<br />

Using the <strong>de</strong>nsity of the set T in [0, +∞) and relying both on the monotonicity of the<br />

mea<strong>sur</strong>es and the γ -convergence of monotone sequences, we <strong>de</strong>fine the mapping t ↦→ μt by<br />

γ<br />

γ -continuity on the right. Finally, it can be conclu<strong>de</strong>d as in [8] that the convergence {vr >t} −→<br />

μt holds on R \ N, where N is at most countab<strong>le</strong>. Precisely, N is the set of discontinuity points<br />

of the monotonous real function [0, +∞) ∋ t → <br />

wμt Ω ∈ R.<br />

γ


720 D. Bucur / Journal of Functional Analysis 236 (2006) 712–725<br />

By abuse of notation and for the simplicity of the exposition, we renote the constructed subsequence<br />

by (vn).<br />

We use the hypothesis (ii) and choose in re<strong>la</strong>tion (12) the sequence (un) <strong>de</strong>fined in the following<br />

way:<br />

<br />

−pun = 0 in{vn >t},<br />

(14)<br />

in Ω \{vn >t}.<br />

un = wΩ<br />

Following [17], we have −pun 0inΩ, so the sequence (un) is admissib<strong>le</strong> in (ii). This<br />

sequence is boun<strong>de</strong>d in W 1,p<br />

0 (Ω), hence for a subsequence, still <strong>de</strong>noted using the same in<strong>de</strong>x,<br />

we have that un ⇀uweakly in W 1,p<br />

0 (Ω). Following [2], we also have that ∇un →∇u a.e. in Ω,<br />

since −pun are uniformly boun<strong>de</strong>d positive Radon mea<strong>sur</strong>es.<br />

The information on the γ -convergence of the <strong>le</strong>vel sets {vn >t} gives:<br />

• u = wΩ on Ω \ Aμ. This is a consequence of the fact that u − wΩ ∈ L p (Ω, μ).<br />

• On Aμ, u satisfies the equation<br />

−pu + μ|u − wΩ| p−2 (u − wΩ) = 0 (15)<br />

in the variational sense of W 1,p<br />

0 (Ω) ∩ Lp (Ω, μ), i.e. for every φ ∈ W 1,p<br />

0 (Ω) ∩ Lp (Ω, μ)<br />

we have<br />

<br />

|∇u| p−2 <br />

∇u∇φdx+ |u − wΩ| p−2 (u − wΩ)φ dμ = 0. (16)<br />

Ω<br />

Ω<br />

In<strong>de</strong>ed, for proving that u satisfies Eq. (15) on Aμ with the mea<strong>sur</strong>e μ issued from the<br />

γ -convergence of the <strong>le</strong>vel sets {vn >t}, we can use a simi<strong>la</strong>r argument as in [9]. Denoting<br />

θn = un − wΩ and θ = u − wΩ, it is enough the prove that<br />

gn =: p(θn + wΩ) − pθn<br />

H<br />

−→ g =: p(θ + wΩ) − pθ,<br />

the convergence H being un<strong>de</strong>rstood in the following sense: for every sequence<br />

φn ∈ W 1,p<br />

0 {vn >t} <br />

which converges weakly in W 1,p<br />

0 (Ω) to φ we have<br />

〈gn,φn〉 W −1,q (Ω)×W 1,p<br />

0 (Ω) →〈g,φ〉 W −1,q (Ω)×W 1,p<br />

0 (Ω) .<br />

Since we know that ∇un converges a.e. to ∇u, for every δ>0, there exists a set E such that<br />

|E|


lim <br />

<br />

n→∞<br />

Ω<br />

lim sup<br />

n→∞<br />

<br />

+<br />

D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 721<br />

<br />

|∇un| p−2 ∇un −|∇θn| p−2 <br />

<br />

∇θn ∇φn dx −<br />

E<br />

<br />

<br />

|∇un| p−2 ∇un −|∇θn| p−2 <br />

∇θn ∇φn<br />

dx<br />

E<br />

<br />

|∇u| p−2 ∇u −|∇θ| p−2 ∇θ ∇φ dx.<br />

For every ε>0, there exists M such that<br />

Ω<br />

<br />

<br />

p−2 p−2<br />

|∇u| ∇u −|∇θ| ∇θ ∇φdx<br />

<br />

|∇ρ| M ⇒ ∇(ρ + wΩ) p−2 ∇(ρ + wΩ) −|∇ρ| p−2 ∇ρ ε|∇ρ| p−1 .<br />

Thus, for a given ε>0wehave<br />

<br />

<br />

|∇un| p−2 ∇un −|∇θn| p−2 <br />

∇θn ∇φn<br />

dx<br />

E<br />

ε<br />

<br />

E∩{|∇θn|M}<br />

|∇θn| p−1 |∇φn| dx + 3M p−1<br />

C ε + 3M p−1 E ∩ |∇θn| t} ,<br />

H<br />

that {vn >t} γ -converges to μ and gn −→ g.<br />

Applying inequality (12) to the sequence (un) constructed above, we get<br />

<br />

lim inf<br />

n→∞<br />

|∇un|<br />

Ω<br />

p−2 <br />

∇un∇vn dx |∇u|<br />

Ω<br />

p−2 ∇u∇vdx,<br />

or, <strong>de</strong>composing the integrals by using vn − t = (vn − t) + − (vn − t) − ,<br />

<br />

lim inf<br />

n→∞<br />

Ω<br />

Ω<br />

|∇un| p−2 ∇un∇(vn − t) + <br />

dx −<br />

Ω<br />

|∇un| p−2 ∇un∇(vn − t) − dx<br />

<br />

|∇u| p−2 ∇u∇(v − t) + <br />

dx − |∇u| p−2 ∇u∇(v − t) − dx.<br />

Ω


722 D. Bucur / Journal of Functional Analysis 236 (2006) 712–725<br />

We have that un = wΩ on {vn >t}, hence we get<br />

<br />

|∇un| p−2 ∇un∇(vn − t) − <br />

dx → |∇u| p−2 ∇u∇(v − t) − dx.<br />

Ω<br />

Consequently<br />

<br />

lim inf<br />

n→∞<br />

Ω<br />

|∇un| p−2 ∇un∇(vn − t) + <br />

dx <br />

Ω<br />

Ω<br />

|∇u| p−2 ∇u∇(v − t) + dx.<br />

But, un is p-harmonic on {vn >t}, hence the integrals on the <strong>le</strong>ft-hand si<strong>de</strong> are equal to zero!<br />

On the right-hand si<strong>de</strong>, we use Eq. (16) satisfied by u on Aμ and get<br />

<br />

0 <br />

Ω<br />

|u − wΩ| p−2 (wΩ − u)(v − t) + dμ. (17)<br />

Since all terms un<strong>de</strong>r the sum are positive on the right-hand si<strong>de</strong>, we get<br />

<br />

|u − wΩ| p−2 (wΩ − u)(v − t) + dμ = 0.<br />

Ω<br />

By the comparison princip<strong>le</strong> of p-superharmonic functions we know wΩ(x) > u(x) p-q.e.<br />

on Aμ. Consequently we get that μ({v>t}) = 0.<br />

The main i<strong>de</strong>a is to prove that μ(Aμ) = 0 in which case μ =∞Ω\Aμ and thus Aμ ={v>t}.<br />

As a consequence we would get that the sequence of <strong>le</strong>vel sets {vn >t} γ -converges to {v >t}<br />

and Theorem 2.5 could be applied. It may be possib<strong>le</strong> that {v >t} is a strict subset of Aμ (in<br />

the sense of capacity), so this argument is not enough to conclu<strong>de</strong> that μ = 0onAμ, and needs<br />

further investigation.<br />

Step 2. From the weak-W 1,p<br />

0 (Ω) convergence vn ⇀v, we get<br />

(vn − t) + ⇀(v− t) +<br />

weakly in W 1,p<br />

0 (Ω), and from the γ -convergence of the sets {vn >t} to μ we get (v − t) + ∈<br />

W 1,p<br />

0 (Ω) ∩ Lp (Ω, μ).<br />

The same holds also for (v − (t + ε)) + for ε>0. Let us <strong>de</strong>note<br />

At = γ − lim<br />

ε→0 {v>t+ ε}.<br />

This γ -limit exists from the monotonicity of the sets. Obviously we get At ⊆ Aμ.<br />

On the other hand, following Lemma 2.3, Aμ ∩{vt} we get that Aμ ={v>t}<br />

up to a set of zero capacity.


D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 723<br />

Step 3. We conclu<strong>de</strong> the proof by noticing that the mapping t ↦→{v>t} is γ -continuous on R<br />

with the exception of an at most countab<strong>le</strong> family of points, hence we get that μt =∞{vt} on R<br />

with the exception of an at most countab<strong>le</strong> family of points, so the limit in the sense of obstac<strong>le</strong>s<br />

of the sequence (vn) is the function v. ✷<br />

Remark 3.2. In [3] the authors prove that if un,vn are weakly convergent sequences of nonnegative<br />

functions of W 1,p<br />

0 (Ω) such that −pun 0 and −pvn 0, then re<strong>la</strong>tion (1) holds<br />

true. The main technical argument is that min{vn,v} converges strongly in W 1,p<br />

0 (Ω) to v, which<br />

is obtained as a consequence of the p-superharmonicity of vn. We notice that the strong convergence<br />

min{vn,v}→v together with the weak convergence vn ⇀vin W 1,p<br />

0 (Ω) imply that<br />

obst<br />

−→ v, hence assertion (i) of Theorem 3.1 is satisfied.<br />

vn<br />

4. Further remarks<br />

The extension of the results of the paper to higher or<strong>de</strong>r operators is not an obvious matter.<br />

On the one hand, the positivity preserving property for higher or<strong>de</strong>r operators is <strong>de</strong>pending on<br />

the geometric set where the operator is <strong>de</strong>fined. For the bi-Lap<strong>la</strong>ce operator, positivity preserving<br />

holds, for examp<strong>le</strong>, on balls and the entire space but fails in general, even on smooth sets<br />

(see [11]). If the operators are positivity preserving, the sufficiency part of Theorem 3.1 still<br />

holds true. Neverthe<strong>le</strong>ss, <strong>de</strong>aling with the necessity part is more difficult, since the convergence<br />

of obstac<strong>le</strong>s and the re<strong>la</strong>xation of sets through γ -convergence is not known. Moreover, the <strong>la</strong>ck<br />

of reticu<strong>la</strong>rity of the Sobo<strong>le</strong>v spaces of or<strong>de</strong>r greater than 1 may be a supp<strong>le</strong>mentary difficulty<br />

for the necessity part.<br />

Remark 4.1. There is a significant non-symmetry between the two terms in the <strong>du</strong>ality pro<strong>du</strong>ct.<br />

The convergence in the sense of the obstac<strong>le</strong>s of vn is re<strong>la</strong>ted to the Mosco convergence of<br />

the sets Kvn and is in<strong>de</strong>pen<strong>de</strong>nt on the choice of the operator itself which is associated to the<br />

terms un. This means that if Theorem 3.1 holds for a sequence (vn)n and the p-Lap<strong>la</strong>ce operator,<br />

then Theorem 3.1 holds for the same sequence (vn)n and an operator simi<strong>la</strong>r to the p-Lap<strong>la</strong>cian.<br />

In the sufficient condition given in [3], the superharmonicity of the first sequence un serves<br />

for monotonicity of the <strong>du</strong>ality pro<strong>du</strong>ct of −pun against vn and the metric projection of v on<br />

the cone {ϕ vn}, respectively, and on the other hand, the p-superharmonicity of the second<br />

sequence vn is a sufficient condition to obtain the obstac<strong>le</strong> convergence as a consequence of the<br />

weak convergence.<br />

Remark 4.2 (Uniformly oscil<strong>la</strong>ting obstac<strong>le</strong>s). Assume that N p>N− 1 and vn ∈ W 1,p<br />

0 (Ω),<br />

vn 0 are continuous and have uniform oscil<strong>la</strong>tions, i.e. there exists a sequence of numbers (ln)n<br />

and a <strong>de</strong>nse set {t1,t2,...}⊆R+ such that<br />

∀r, n ∈ N, ♯{vn tr} lr, (18)<br />

where ♯A <strong>de</strong>notes the number of the connected components of the set A. Ifvn⇀vweakly in<br />

W 1,p<br />

0 (Ω), then vn converges in the sense of obstac<strong>le</strong>s to v. In<strong>de</strong>ed, the second Mosco condition<br />

is satisfied as a consequence of the weak convergence vn ⇀v. In or<strong>de</strong>r to prove the first Mosco<br />

condition, we follow the i<strong>de</strong>a of the proof of Theorem 3.1, and assume for contradiction that<br />

re<strong>la</strong>tion (13) holds for some δ>0. From (18) and the shape compactness/stability result of [6],


724 D. Bucur / Journal of Functional Analysis 236 (2006) 712–725<br />

there exists a subsequence of the sets {vn >tr} and a set Ωtr , such that {vnk >tr} γ -converges<br />

to Ωtr . Consequently, by a diagonal extraction proce<strong>du</strong>re as in Theorem 3.1, we construct an<br />

obstac<strong>le</strong> h such that {h >tr} =Ωtr and the sequence vnk converges in the sense of obstac<strong>le</strong>s<br />

to h.Fromtheγ-convergence of the <strong>le</strong>vel sets, we get (v − tr) + ∈ W 1,p<br />

(Ωtr 0 ), hence h v.This<br />

contradicts (13), from the Mosco convergence.<br />

It is need<strong>le</strong>ss to say that vn are not, in general, p-superharmonic.<br />

Remark 4.3. In [3], the authors give an examp<strong>le</strong> of a sequence of positive functions vn which<br />

are not superharmonic such that inequality (1) fails to be true for a suitab<strong>le</strong> sequence un of superharmonic<br />

functions. This construction is done around the pioneering result of Cioranescu and<br />

Murat [10] on the “strange term” appearing in the re<strong>la</strong>xation process through the γ -convergence.<br />

The presence of the strange term is an argument of non-γ -convergence of obstac<strong>le</strong>s! So, from<br />

this point of view it is not <strong>sur</strong>prising that inequality (1) is vio<strong>la</strong>ted, although the choice of un has<br />

to be done carefully.<br />

Remark 4.4. We give in the sequel an examp<strong>le</strong> of non-superharmonic functions vn for which<br />

the second assertion of Theorem 3.1 holds. Let η ∈ C1 (R2 , R) be a periodic function of period<br />

(l1,l2) and ϕ ∈ C∞ 0 (Ω). We consi<strong>de</strong>r the sequence of functions<br />

<br />

x<br />

vn = wΩ + ϕεη .<br />

ε<br />

It is c<strong>le</strong>ar that this sequence converges weakly but not strongly in H 1 0 (Ω) to v = wΩ, provi<strong>de</strong>d<br />

that ϕ or η are not the zero functions and ε ↓ 0. Inequality (1) is satisfied for all admissib<strong>le</strong><br />

sequences (un). This is a consequence of the obstac<strong>le</strong> convergence of vn towards v which can be<br />

easily proved. Neverthe<strong>le</strong>ss for small ε, the functions vn are not superharmonic!<br />

References<br />

[1] H. Attouch, Variational Convergence for Functions and Operators, Pitman, London, 1984.<br />

[2] L. Boccardo, F. Murat, Almost everywhere convergence of the <strong>gradient</strong>s of solutions to elliptic and parabolic equations,<br />

Nonlinear Anal. 19 (6) (1992) 581–597.<br />

[3] M. Briane, G. Mokobodzki, F. Murat, Variations on a strange <strong>semi</strong>-continuity result, J. Funct. Anal. 227 (1) (2005)<br />

78–112.<br />

[4] M. Briane, G. Mokobodzki, F. Murat, Semi-strong convergence of sequence satisfying a variational inequality,<br />

preprint, http://www.insa-rennes.fr/~mbriane.<br />

[5] D. Bucur, G. Buttazzo, Variational Methods in Shape Optimization Prob<strong>le</strong>ms, Progr. in Nonlinear Differential Equations,<br />

vol. 65, Birkhäuser, Basel, 2005.<br />

[6] D. Bucur, P. Trebeschi, Shape optimisation prob<strong>le</strong>ms governed by nonlinear state equations, Proc. Roy. Soc. Edinburgh<br />

Sect. A 128 (5) (1998) 945–963.<br />

[7] D. Bucur, P. Trebeschi, A new re<strong>la</strong>xation space for obstac<strong>le</strong>s, Acta Appl. Math. 79 (3) (2003) 177–194.<br />

[8] D. Bucur, G. Buttazzo, P. Trebeschi, An existence result for optimal obstac<strong>le</strong>s, J. Funct. Anal. 162 (1) (1999) 96–<br />

119.<br />

[9] J. Casado Diaz, A. Garroni, Asymptotic behavior of nonlinear elliptic systems on varying domains, SIAM J. Math.<br />

Anal. 31 (3) (2000) 581–624.<br />

[10] D. Cioranescu, F. Murat, Un terme étrange venu d’ail<strong>le</strong>urs, in: Nonlinear Partial Differential Equations and Their<br />

Applications, Collège <strong>de</strong> France Seminar, vol. II, Paris, 1979/1980, in: Res. Notes in Math., vol. 60, Pitman, Boston,<br />

MA, 1982, pp. 98–138, 389–390.<br />

[11] A. Dall’Acqua, G. Sweers, On domains for which the c<strong>la</strong>mped p<strong>la</strong>te system is positivity preserving, in: Partial<br />

Differential Equations and Inverse Prob<strong>le</strong>ms, in: Contemp. Math., vol. 362, Amer. Math. Soc., Provi<strong>de</strong>nce, RI,<br />

2004, pp. 133–144.


D. Bucur / Journal of Functional Analysis 236 (2006) 712–725 725<br />

[12] G. Dal Maso, Some necessary and sufficient conditions for convergence of sequences of uni<strong>la</strong>teral convex sets,<br />

J. Funct. Anal. 62 (1985) 119–159.<br />

[13] G. Dal Maso, Γ -convergence and μ-capacities, Ann. Sc. Norm. Super. Pisa Cl. Sci. 14 (1988) 423–464.<br />

[14] G. Dal Maso, U. Mosco, Wiener’s criterion and Γ -convergence, Appl. Math. Optim. 15 (1987) 15–63.<br />

[15] G. Dal Maso, F. Murat, Asymptotic behavior and correctors for Dirich<strong>le</strong>t prob<strong>le</strong>ms in perforated domains with<br />

homogeneous monotone operators, Ann. Sc. Norm. Super. Pisa Cl. Sci. 24 (1997) 239–290.<br />

[16] L.I. Hedberg, Spectral synthesis in Sobo<strong>le</strong>v spaces and uniqueness of solutions of Dirich<strong>le</strong>t prob<strong>le</strong>ms, Acta<br />

Math. 147 (3–4) (1981) 237–263.<br />

[17] J. Heinonen, T. Kilpe<strong>la</strong>inen, O. Martio, Nonlinear Potential Theory of Degenerate Elliptic Equations, C<strong>la</strong>rendon<br />

Press, Oxford, 1993.<br />

[18] W.P. Ziemer, Weakly Differentiab<strong>le</strong> Functions, Springer-Ver<strong>la</strong>g, Berlin, 1989.


Albeverio, Sergio, 634<br />

Barceló, Juan A., 1<br />

Bharali, Gautam, 351<br />

Bo˙zejko, Marek, 59<br />

Bryc, Włodzimierz, 59<br />

Brydges, David, 682<br />

Bucur, Dorin, 712<br />

Burq, Nico<strong>la</strong>s, 265<br />

Char<strong>le</strong>s, L., 299<br />

Crabb, M.J., 630<br />

Cranston, M., 78<br />

Gamblin, Didier, 201<br />

Kissin, Edward, 609<br />

Kosyak, A<strong>le</strong>xandre, 634<br />

Koufany, Khalid, 546<br />

Kucerovsky, Dan, 395<br />

Li, Hong-Quan, 369<br />

Mil<strong>le</strong>r, Luc, 592<br />

Mountford, T.S., 78<br />

0022-1236/2006 Published by Elsevier Inc.<br />

doi:10.1016/S0022-1236(06)00209-6<br />

Journal of Functional Analysis 236 (2006) 726<br />

AUTHOR INDEX FOR VOLUME 236<br />

Olofsson, An<strong>de</strong>rs, 517<br />

P<strong>la</strong>nchon, Fabrice, 265<br />

Porretta, A<strong>le</strong>ssio, 581<br />

Prio<strong>la</strong>, Enrico, 244<br />

Ruiz, Alberto, 1<br />

Schultz, Hanne, 457<br />

Schulze, Michael, 120<br />

Serre, Denis, 409<br />

Ta<strong>la</strong>rczyk, Anna, 682<br />

Van Schaftingen, Jean, 490<br />

Vassout, Stéphane, 161<br />

Vega, Luis, 1<br />

Véron, Laurent, 581<br />

Wang,Feng-Yu,244<br />

Wi´snicki, Andrzej, 447<br />

Yakubovich, Dmitry V., 25<br />

Zhang, Genkai, 546<br />

www.elsevier.com/locate/jfa

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!