26.10.2012 Views

Potential Theory in Classical Probability - Developers

Potential Theory in Classical Probability - Developers

Potential Theory in Classical Probability - Developers

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Lecture Notes <strong>in</strong> Mathematics 1954<br />

Editors:<br />

J.-M. Morel, Cachan<br />

F. Takens, Gron<strong>in</strong>gen<br />

B. Teissier, Paris


Philippe Biane · Luc Bouten · Fabio Cipriani<br />

Norio Konno · Nicolas Privault · Quanhua Xu<br />

Quantum<br />

<strong>Potential</strong> <strong>Theory</strong><br />

Editors:<br />

Michael Schürmann<br />

Uwe Franz<br />

ABC


Editors<br />

Uwe Franz<br />

Département de mathématiques<br />

de Besancon<br />

Université de Franche-Comté<br />

16, route de Gray<br />

25030 Besancon cedex<br />

France<br />

uwe.franz@univ-fcomte.fr<br />

and<br />

Graduate School of Information Sciences<br />

Tohoku University<br />

6-3-09 Aramaki-Aza-Aoba, Aoba-ku<br />

Sendai 980-8579<br />

Japan<br />

ISBN: 978-3-540-69364-2 e-ISBN: 978-3-540-69365-9<br />

DOI: 10.1007/978-3-540-69365-9<br />

Lecture Notes <strong>in</strong> Mathematics ISSN pr<strong>in</strong>t edition: 0075-8434<br />

ISSN electronic edition: 1617-9692<br />

Library of Congress Control Number: 2008932185<br />

Mathematics Subject Classification (2000): 58B34, 81R60, 31C12, 53C21<br />

Michael Schürmann<br />

Institut für Mathematik und Informatik<br />

Ernst-Moritz-Arndt-Universität Greifswald<br />

Jahnstrasse 15a<br />

17487 Greifswald<br />

Germany<br />

schurman@uni-greifswald.de<br />

c○ 2008 Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg<br />

This work is subject to copyright. All rights are reserved, whether the whole or part of the material is<br />

concerned, specifically the rights of translation, repr<strong>in</strong>t<strong>in</strong>g, reuse of illustrations, recitation, broadcast<strong>in</strong>g,<br />

reproduction on microfilm or <strong>in</strong> any other way, and storage <strong>in</strong> data banks. Duplication of this publication<br />

or parts thereof is permitted only under the provisions of the German Copyright Law of September 9,<br />

1965, <strong>in</strong> its current version, and permission for use must always be obta<strong>in</strong>ed from Spr<strong>in</strong>ger. Violations<br />

are liable to prosecution under the German Copyright Law.<br />

The use of general descriptive names, registered names, trademarks, etc. <strong>in</strong> this publication does not<br />

imply, even <strong>in</strong> the absence of a specific statement, that such names are exempt from the relevant protective<br />

laws and regulations and therefore free for general use.<br />

Cover design: SPi Publish<strong>in</strong>g Services<br />

Pr<strong>in</strong>ted on acid-free paper<br />

987654321<br />

spr<strong>in</strong>ger.com


Preface<br />

This volume conta<strong>in</strong>s the notes of lectures given at the School “Quantum<br />

<strong>Potential</strong> <strong>Theory</strong>: Structure and Applications to Physics”. This school was<br />

held at the Alfried Krupp Wissenschaftskolleg <strong>in</strong> Greifswald from February<br />

26 to March 9, 2007. We thank the lecturers for the hard work they accomplished<br />

<strong>in</strong> prepar<strong>in</strong>g and giv<strong>in</strong>g these lectures and <strong>in</strong> writ<strong>in</strong>g these notes.<br />

Their lectures give an <strong>in</strong>troduction to current research <strong>in</strong> their doma<strong>in</strong>s, which<br />

is essentially self-conta<strong>in</strong>ed and should be accessible to Ph.D. students. We<br />

hope that this volume will help to br<strong>in</strong>g together researchers from the areas<br />

of classical and quantum probability, functional analysis and operator algebras,<br />

and theoretical and mathematical physics, and contribute <strong>in</strong> this way<br />

to develop<strong>in</strong>g further the subject of quantum potential theory.<br />

We are greatly <strong>in</strong>debted to the Alfried Krupp von Bohlen und Halbach-<br />

Stiftung for the f<strong>in</strong>ancial support, without which the school would not have<br />

been possible. We are also very thankful for the support by the University of<br />

Greifswald and the University of Franche-Comté. One of the organisers (UF)<br />

was supported by a Marie Curie Outgo<strong>in</strong>g International Fellowship of the EU<br />

(Contract Q-MALL MOIF-CT-2006-022137).<br />

Special thanks go to Melanie H<strong>in</strong>z who helped with the preparation and<br />

organisation of the school and who took care of all of the logistics.<br />

F<strong>in</strong>ally, we would like to thank all the students for com<strong>in</strong>g to Greifswald<br />

and help<strong>in</strong>g to make the school a success.<br />

Sendai and Greifswald, Uwe Franz<br />

June 2008 Michael Schürmann<br />

v


Contents<br />

Introduction .................................................. 1<br />

<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> ..................... 3<br />

Nicolas Privault<br />

1 Introduction.......................................... 3<br />

2 Analytic<strong>Potential</strong><strong>Theory</strong>.............................. 4<br />

3 MarkovProcesses ..................................... 24<br />

4 StochasticCalculus.................................... 31<br />

5 Probabilistic Interpretations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

References ................................................. 58<br />

Introduction to Random Walks<br />

on Noncommutative Spaces ................................... 61<br />

Philippe Biane<br />

1 Introduction.......................................... 61<br />

2 Noncommutative Spaces and Random Variables . . . . . . . . . . . 62<br />

3 QuantumBernoulliRandomWalks...................... 68<br />

4 BialgebrasandGroupAlgebras ......................... 73<br />

5 Random Walk on the Dual of SU(2)..................... 76<br />

6 RandomWalksonDualsofCompactGroups ............. 80<br />

7 The Case of SU(n).................................... 82<br />

8 Choquet-Deny Theorem for Duals of Compact Groups . . . . . 87<br />

9 The Mart<strong>in</strong> Compactification of the Dual of SU(2) ........ 90<br />

10 Central Limit Theorems for the Bernoulli Random Walk . . . 94<br />

11 The Heisenberg Group and the Noncommutative<br />

BrownianMotion ..................................... 100<br />

12 DilationsforNoncompactGroups ....................... 106<br />

13 Pitman’s Theorem and the Quantum Group SUq(2) ....... 110<br />

References ................................................. 114<br />

vii


viii Contents<br />

Interactions between Quantum <strong>Probability</strong> and Operator<br />

Space <strong>Theory</strong> ................................................. 117<br />

Quanhua Xu<br />

1 Introduction.......................................... 117<br />

2 CompletelyPositiveMaps.............................. 119<br />

3 Concrete Operator Spaces and Completely<br />

Bounded Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

4 Ruan’sTheorem:AbstractOperatorSpaces .............. 126<br />

5 Complex Interpolation and Operator Hilbert Spaces . . . . . . . 130<br />

6 Vector-valued Noncommutative Lp-spaces ................ 132<br />

7 NoncommutativeKh<strong>in</strong>tch<strong>in</strong>eTypeInequalities............ 137<br />

8 Embedd<strong>in</strong>g of OH <strong>in</strong>to Noncommutative L1 .............. 156<br />

References ................................................. 158<br />

Dirichlet Forms on Noncommutative Spaces .................. 161<br />

Fabio Cipriani<br />

1 Introduction.......................................... 161<br />

2 Dirichlet Forms on C ∗ -algebras and KMS-symmetric<br />

Semigroups........................................... 168<br />

3 Dirichlet Forms <strong>in</strong> Quantum Statistical Mechanics . . . . . . . . . 218<br />

4 Dirichlet Forms and Differential Calculus on C ∗ -algebras . . . 224<br />

5 Noncommutative <strong>Potential</strong> <strong>Theory</strong> and Riemannian<br />

Geometry ............................................ 245<br />

6 DirichletFormsandNoncommutativeGeometry .......... 259<br />

7 Appendix ............................................ 265<br />

8 ListofExamples ...................................... 270<br />

References ................................................. 272<br />

Applications of Quantum Stochastic Processes <strong>in</strong> Quantum<br />

Optics ........................................................ 277<br />

Luc Bouten<br />

1 Quantum <strong>Probability</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277<br />

2 ConditionalExpectations............................... 286<br />

3 QuantumStochasticCalculus........................... 290<br />

4 QuantumFilter<strong>in</strong>g .................................... 298<br />

References ................................................. 306<br />

Quantum Walks .............................................. 309<br />

Norio Konno<br />

Part I: Discrete-Time Quantum Walks<br />

1 LimitTheorems ...................................... 311<br />

2 DisorderedCase...................................... 350<br />

3 ReversibleCellularAutomata .......................... 357<br />

4 QuantumCellularAutomata........................... 368<br />

5 Cycle ............................................... 377<br />

6 AbsorptionProblems.................................. 387


Contents ix<br />

Part II: Cont<strong>in</strong>uous-Time Quantum Walks<br />

7 One-DimensionalLattice .............................. 401<br />

8 Tree ................................................ 410<br />

9 UltrametricSpace .................................... 420<br />

10 Cycle ............................................... 432<br />

References ................................................. 441<br />

Index ......................................................... 453


Contributors<br />

Philippe Biane<br />

Institut d’électronique et d’<strong>in</strong>formatique Gaspard-Monge, 77454 Marne-la-<br />

Vallée Cedex 2, France<br />

biane@univ-mlv.fr<br />

Luc Bouten<br />

Caltech Physical Measurement and Control, MC 266-33, 1200 E California<br />

Blvd, Pasadena, CA 91125, USA<br />

bouten@caltech.edu<br />

Fabio Cipriani<br />

Dipartimento di Matematica, Politecnico di Milano, Piazza Leonardo da<br />

V<strong>in</strong>ci, 32 I-20133 Milano, Italy<br />

fabio.cipriani@polimi.it<br />

Norio Konno<br />

Department of Applied Mathematics, Yokohama National University, 79-5<br />

Tokiwadai, Yokohama 240-8501, Japan<br />

konno@ynu.ac.jp<br />

Nicolas Privault<br />

Department of Mathematics, City University of Hong Kong, 83 Tat Chee<br />

Avenue, Kowloon Tong, Hong Kong<br />

nprivaul@cityu.edu.hk<br />

Quanhua Xu<br />

UFR Sciences et techniques, Université deFranche-Comté,16,routedeGray,<br />

F-25 030 Besançon Cedex, France<br />

quanhua.xu@univ-fcomte.fr<br />

xi


Introduction<br />

The term potential theory comes from 19th century physics, where the fundamental<br />

forces like gravity or electrostatic forces were described as the gradients<br />

of potentials, i.e. functions which satisfy the Laplace equation. Hence<br />

potential theory was the study of solutions of the Laplace equation. Nowadays<br />

the fundamental forces <strong>in</strong> physics are described by systems of non-l<strong>in</strong>ear partial<br />

differential equations such as the E<strong>in</strong>ste<strong>in</strong> equations and the Yang-Mills<br />

equations, and the Laplace equation arises only as a limit<strong>in</strong>g case. Nevertheless,<br />

the Laplace equation is still used <strong>in</strong> applications <strong>in</strong> many areas of physics<br />

and eng<strong>in</strong>eer<strong>in</strong>g like heat conduction and electrostatics. And the term “potential<br />

theory” has survived as a convenient label for the theory of functions<br />

satisfy<strong>in</strong>g the Laplace equation, i.e. so-called harmonic functions.<br />

In the 20th century, with the development of probability and stochastic<br />

processes, it was discovered that potential theory is <strong>in</strong>timately related to the<br />

theory of Markov processes, <strong>in</strong> particular diffusion processes and Brownian<br />

motion. The distributions of these processes evolve accord<strong>in</strong>g to a heat equation,<br />

and <strong>in</strong>variant distributions satisfy a Laplace-type equation. Conversely,<br />

these processes can be used to express solutions of, e.g., the Laplace equation.<br />

For more details see Nicolas Privault’s lecture “<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong><br />

<strong>Probability</strong>” <strong>in</strong> this volume.<br />

The notions of quantum stochastic processes and quantum Markov processes<br />

were <strong>in</strong>troduced <strong>in</strong> the 1970’s and allow to describe open quantum<br />

systems <strong>in</strong> close analogy to classical probability and classical Markov processes.<br />

Roughly speak<strong>in</strong>g, one can now recognize two different trends <strong>in</strong> the<br />

subsequent development of the theory of quantum Markov processes. The<br />

first is guided by physical applications, studies concrete physically motivated<br />

models, and develops tools for filter<strong>in</strong>g noisy quantum signals or control<strong>in</strong>g<br />

noisy quantum systems. The second aims to develop a mathematical theory,<br />

by generaliz<strong>in</strong>g or extend<strong>in</strong>g key results of the theory of Markov processes<br />

to the quantum (or noncommutative) case, and by look<strong>in</strong>g for analogues of<br />

important tools that greatly <strong>in</strong>fluenced the development of classical potential<br />

theory, like stochastic calculus, Dirichlet forms, or boundaries of random<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 1<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


2 Introduction<br />

walks. In our school the first direction was represented by Luc Bouten’s lecture<br />

“Applications of Controlled Quantum Processes <strong>in</strong> Quantum Optics”,<br />

the second by Philippe Biane’s lecture “Introduction to Random Walks on<br />

Noncommutative Spaces” and by Fabio Cipriani’s lecture “Noncommutative<br />

Dirichlet Forms”, see also the correspond<strong>in</strong>g chapters of this book.<br />

Besides provid<strong>in</strong>g important background material on operator algebras<br />

and noncommutative analogues of function spaces used <strong>in</strong> other lectures,<br />

Quanhua Xu’s lecture on “Interactions between Quantum <strong>Probability</strong> and<br />

Operator Space <strong>Theory</strong>” shows how quantum probability can be applied to<br />

modern functional analysis. For example, a clever choice of sequences of quantum<br />

random variables plays an essential role <strong>in</strong> establish<strong>in</strong>g key results like<br />

noncommutative Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities.<br />

Central questions from probabilistic potential theory like the computation<br />

of hitt<strong>in</strong>g times and the study of the asymptotic behaviour of a walk are also<br />

the ma<strong>in</strong> topic <strong>in</strong> Norio Konno’s lecture on “Quantum Walks”. These quantum<br />

walks are not quantum Markov processes <strong>in</strong> the sense of the lectures<br />

by Biane, Bouten, and Cipriani, but another type of quantum analogue of<br />

random walks and Markov cha<strong>in</strong>s, and many of the classical potential theoretical<br />

methods have <strong>in</strong>terest<strong>in</strong>g analogues adapted to this case. By giv<strong>in</strong>g<br />

an <strong>in</strong>troduction and survey of this quickly develop<strong>in</strong>g field this lecture was<br />

an enrichment of the school and nicely complements the other chapters.<br />

The goal of the School “Quantum <strong>Potential</strong> <strong>Theory</strong>: Structure and Applications<br />

to Physics” and these lecture notes is two-fold. First of all we want to<br />

provide an <strong>in</strong>troduction to the rapidly develop<strong>in</strong>g theory of quantum Markov<br />

semigroups and quantum Markov processes with its manifold aspects rang<strong>in</strong>g<br />

from functional analysis and probability theory to quantum physics. We<br />

hope that we have succeeded <strong>in</strong> prepar<strong>in</strong>g a monograph that is accessible<br />

to graduate students <strong>in</strong> mathematics and physics. But furthermore we also<br />

hope that this book will catch the <strong>in</strong>terest of experienced mathematicians<br />

and physicists work<strong>in</strong>g <strong>in</strong> this field or related fields, <strong>in</strong> order to stimulate<br />

more communication between researchers work<strong>in</strong>g on “pure” and “applied”<br />

aspects. We believe that a strong collaboration between these communities<br />

will be to everybody’s benefit. Keep<strong>in</strong>g <strong>in</strong> m<strong>in</strong>d the physical applications will<br />

help to sharpen the theoreticians’ eye for the relevant questions and properties,<br />

and new powerful mathematical tools will allow to get a better and<br />

deeper understand<strong>in</strong>g of concrete physical systems.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong><br />

<strong>Probability</strong><br />

Nicolas Privault<br />

Abstract These notes are an elementary <strong>in</strong>troduction to classical potential<br />

theory and to its connection with probabilistic tools such as stochastic<br />

calculus and the Markov property. In particular we review the probabilistic<br />

<strong>in</strong>terpretations of harmonicity, of the Dirichlet problem, and of the Poisson<br />

equation us<strong>in</strong>g Brownian motion and stochastic calculus.<br />

1 Introduction<br />

The orig<strong>in</strong>s of potential theory can be traced to the physical problem of<br />

reconstruct<strong>in</strong>g a repartition of electric charges <strong>in</strong>side a planar or a spatial<br />

doma<strong>in</strong>, given the measurement of the electrical field created on the boundary<br />

of this doma<strong>in</strong>.<br />

In mathematical analytic terms this amounts to represent<strong>in</strong>g the values of<br />

a function h <strong>in</strong>side a doma<strong>in</strong> given the data of the values of h on the boundary<br />

of the doma<strong>in</strong>. In the simplest case of a doma<strong>in</strong> empty of electric charges,<br />

the problem can be formulated as that of f<strong>in</strong>d<strong>in</strong>g a harmonic function h on<br />

E (roughly speak<strong>in</strong>g, a function with vanish<strong>in</strong>g Laplacian, see § 2.2 below),<br />

given its values prescribed by a function f on the boundary ∂E, i.e. as the<br />

Dirichlet problem: ⎧<br />

⎨ ∆h(y) =0, y ∈ E,<br />

⎩<br />

h(y) =f(y), y ∈ ∂E.<br />

N. Privault<br />

Department of Mathematics, City University of Hong Kong, 83 Tat Chee Avenue,<br />

Kowloon Tong, Hong Kong<br />

e-mail: nprivaul@cityu.edu.hk<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 3<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


4 N. Privault<br />

Close connections between the notion of potential and the Markov property<br />

have been observed at early stages of the development of the theory, see e.g.<br />

[Doo84] and references there<strong>in</strong>. Thus a number of potential theoretic problems<br />

have a probabilistic <strong>in</strong>terpretation or can be solved by probabilistic methods.<br />

These notes aim at gather<strong>in</strong>g both analytic and probabilistic aspects of<br />

potential theory <strong>in</strong>to a s<strong>in</strong>gle document. We partly follow the po<strong>in</strong>t of view<br />

of Chung [Chu95] with complements on analytic potential theory com<strong>in</strong>g<br />

from Helms [Hel69], some additions on stochastic calculus, and probabilistic<br />

applications found <strong>in</strong> Bass [Bas98].<br />

More precisely, we proceed as follow. In Section 2 we give a summary<br />

of classical analytic potential theory: Green kernels, Laplace and Poisson<br />

equations <strong>in</strong> particular, follow<strong>in</strong>g ma<strong>in</strong>ly Brelot [Bre65], Chung [Chu95] and<br />

Helms [Hel69]. Section 3 <strong>in</strong>troduces the Markovian sett<strong>in</strong>g of semigroups<br />

which will be the ma<strong>in</strong> framework for probabilistic <strong>in</strong>terpretations. A sample<br />

of references used <strong>in</strong> this doma<strong>in</strong> is Ethier and Kurtz [EK86], Kallenberg<br />

[Kal02], and also Chung [Chu95]. The probabilistic <strong>in</strong>terpretation of potential<br />

theory also makes significant use of Brownian motion and stochastic<br />

calculus. They are summarized <strong>in</strong> Section 4, see Protter [Pro05], Ikeda and<br />

Watanabe [IW89], however our presentation of stochastic calculus is given<br />

<strong>in</strong> the framework of normal mart<strong>in</strong>gales due to their l<strong>in</strong>ks with quantum<br />

stochastic calculus, cf. Biane [Bia93]. In Section 5 we present the probabilistic<br />

connection between potential theory and Markov processes, follow<strong>in</strong>g<br />

Bass [Bas98], Dynk<strong>in</strong> [Dyn65], Kallenberg [Kal02], and Port and<br />

Stone [PS78]. Our description of the Mart<strong>in</strong> boundary <strong>in</strong> discrete time follows<br />

that of Revuz [Rev75].<br />

2 Analytic <strong>Potential</strong> <strong>Theory</strong><br />

2.1 Electrostatic Interpretation<br />

Let E denote a closed region of Rn , more precisely a compact subset hav<strong>in</strong>g<br />

a smooth boundary ∂E with surface measure σ. Gauss’s law is the ma<strong>in</strong> tool<br />

for determ<strong>in</strong><strong>in</strong>g a repartition of electric charges <strong>in</strong>side E, giventhevaluesof<br />

the electrical field created on ∂E. It states that given a repartition of charges<br />

q(dx) the flux of the electric field U across the boundary ∂E is proportional<br />

to the sum of electric charges enclosed <strong>in</strong> E. Namelywehave<br />

�<br />

�<br />

q(dx) =ɛ0 〈n(x), U(x)〉σ(dx), (2.1)<br />

E<br />

∂E<br />

where q(dx) is a signed measure represent<strong>in</strong>g the distribution of electric<br />

charges, ɛ0 > 0 is the electrical permittivity constant, U(x) denotes the


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 5<br />

electric field at x ∈ ∂E, andn(x) represents the outer (i.e. oriented towards<br />

the exterior of E) unit vector orthogonal to the surface ∂E.<br />

On the other hand the divergence theorem, which can be viewed as a<br />

particular case of the Stokes theorem, states that if U : E → Rn is a C1 vector field we have<br />

�<br />

�<br />

div U(x)dx = 〈n(x), U(x)〉σ(dx), (2.2)<br />

E<br />

∂E<br />

where the divergence div U is def<strong>in</strong>ed as<br />

div U(x) =<br />

n� ∂Ui<br />

(x).<br />

∂xi<br />

The divergence theorem (2.2) can be <strong>in</strong>terpreted as a mathematical formulation<br />

of the Gauss law (2.1). Under this identification, div U(x) is proportional<br />

to the density of charges <strong>in</strong>side E, which leads to the Maxwell equation<br />

i=1<br />

ɛ0 div U(x)dx = q(dx), (2.3)<br />

where q(dx) is the distribution of electric charge at x and U is viewed as the<br />

<strong>in</strong>duced electric field on the surface ∂E.<br />

When q(dx) hasthedensityq(x) atx, i.e. q(dx) =q(x)dx, andthefield<br />

U(x) derives from a potential V : R n → R+, i.e. when<br />

U(x) =∇V (x), x ∈ E,<br />

Maxwell’s equation (2.3) takes the form of the Poisson equation:<br />

where the Laplacian ∆ =div∇ is given by<br />

ɛ0∆V (x) =q(x), x ∈ E, (2.4)<br />

∆V (x) =<br />

n�<br />

j=1<br />

∂2V ∂x2 (x), x ∈ E.<br />

i<br />

In particular, when the doma<strong>in</strong> E is empty of electric charges, the potential<br />

V satisfies the Laplace equation<br />

∆V (x) =0 x ∈ E.<br />

As mentioned <strong>in</strong> the <strong>in</strong>troduction, a typical problem <strong>in</strong> classical potential<br />

theory is to recover the values of the potential V (x)<strong>in</strong>sideE from its values on<br />

the boundary ∂E, giventhatV (x) satisfies the Poisson equation (2.4). This<br />

can be achieved <strong>in</strong> particular by represent<strong>in</strong>g V (x), x ∈ E, asan<strong>in</strong>tegralwith<br />

respect to the surface measure over the boundary ∂E, or by direct solution<br />

of the Poisson equation for V (x).


6 N. Privault<br />

Consider for example the Newton potential kernel<br />

V (x) = q<br />

ɛ0sn<br />

1<br />

�x − y� n−2 , x ∈ Rn \{y},<br />

created by a s<strong>in</strong>gle charge q at y ∈ Rn ,wheres2 =2π, s3 =4π, and<strong>in</strong><br />

general<br />

sn = 2πn/2<br />

, n ≥ 2,<br />

Γ (n/2)<br />

is the surface of the unit n − 1-dimensional sphere <strong>in</strong> Rn .<br />

The electrical field created by V is<br />

U(x) :=∇V (x) = q<br />

ɛ0sn<br />

x − y<br />

�x − y� n−1 , x ∈ Rn \{y},<br />

cf, Figure 1. Lett<strong>in</strong>g B(y, r), resp. S(y, r), denote the open ball, resp. the<br />

sphere, of center y ∈ Rn and radius r>0, we have<br />

�<br />

�<br />

∆V (x)dx = 〈n(x), ∇V (x)〉σ(dx)<br />

B(y,r)<br />

S(y,r)<br />

�<br />

= 〈n(x), U(x)〉σ(dx)<br />

S(y,r)<br />

= q<br />

,<br />

ɛ0<br />

where σ denotes the surface measure on S(y, r).<br />

From this and the Poisson equation (2.4) we deduce that the repartition of<br />

electric charge is<br />

q(dx) =qδy(dx)<br />

i.e. we recover the fact that the potential V is generated by a s<strong>in</strong>gle charge<br />

located at y. We also obta<strong>in</strong> a version of the Poisson equation (2.4) <strong>in</strong> distribution<br />

sense:<br />

1<br />

∆x<br />

= snδy(dx),<br />

�x − y�n−2 where the Laplacian ∆x is taken with respect to the variable x. On the other<br />

hand, tak<strong>in</strong>g E = B(0,r)\B(0,ρ)wehave∂E = S(0,r) ∪ S(0,ρ)and<br />

�<br />

�<br />

�<br />

∆V (x)dx = 〈n(x), ∇V (x)〉σ(dx)+ 〈n(x), ∇V (x)〉σ(dx)<br />

E<br />

hence<br />

S(0,r)<br />

= csn − csn =0,<br />

1<br />

S(0,ρ)<br />

∆x<br />

�x − y�n−2 =0, x ∈ Rn \{y}.<br />

The electrical permittivity ɛ0 will be set equal to 1 <strong>in</strong> the sequel.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 7<br />

-2 -1.5 -1 -0.5 0 0.5 1 1.5 2<br />

Fig. 1 Electrical field generated by a po<strong>in</strong>t mass at y =0.<br />

2.2 Harmonic Functions<br />

The notion of harmonic function will be first <strong>in</strong>troduced from the mean value<br />

property. Let<br />

σ x 1<br />

r (dy) = σ(dy)<br />

snrn−1 denote the normalized surface measure on S(x, r), and recall that<br />

�<br />

� ∞<br />

f(x)dx = sn r n−1<br />

�<br />

f(z)σ y r (dz)dr.<br />

0<br />

S(y,r)<br />

Def<strong>in</strong>ition 2.1. A cont<strong>in</strong>uous real-valued function on an open subset O of<br />

Rn is said to be harmonic, resp. superharmonic, <strong>in</strong> O if one has<br />

�<br />

f(x) = f(y)σ x r (dy),<br />

resp.<br />

�<br />

f(x) ≥<br />

S(x,r)<br />

f(y)σ<br />

S(x,r)<br />

x r (dy),<br />

for all x ∈ O and r>0 such that B(x, r) ⊂ O.


8 N. Privault<br />

Next we show the equivalence between the mean value property and the<br />

vanish<strong>in</strong>g of the Laplacian.<br />

Proposition 2.2. A C2 function f is harmonic, resp. superharmonic, on an<br />

open subset O of Rn if and only if it satisfies the Laplace equation<br />

resp. the partial differential <strong>in</strong>equality<br />

∆f(x) =0, x ∈ O,<br />

∆f(x) ≤ 0, x ∈ O.<br />

Proof. In spherical coord<strong>in</strong>ates, us<strong>in</strong>g the divergence formula and the identity<br />

�<br />

d<br />

f(y + rx)σ<br />

dr S(0,1)<br />

0 �<br />

1(dx) = 〈x, ∇f(y + rx)〉σ<br />

S(0,1)<br />

0 1(dx)<br />

= r<br />

�<br />

∆f(y + rx)dx,<br />

sn<br />

yields<br />

�<br />

B(y,r)<br />

∆f(x)dx = r n−1<br />

�<br />

= snr n−2<br />

�<br />

If f is harmonic, this shows that<br />

�<br />

B(0,1)<br />

S(0,1)<br />

�<br />

n−2 d<br />

= snr<br />

dr<br />

�<br />

n−2 d<br />

= snr<br />

dr<br />

B(y,r)<br />

B(0,1)<br />

∆f(y + rx)dx<br />

〈x, ∇f(y + rx)〉σ 0 1(dx)<br />

f(y + rx)σ<br />

S(0,1)<br />

0 1 (dx)<br />

f(x)σ<br />

S(y,r)<br />

y r (dx).<br />

∆f(x)dx =0,<br />

for all y ∈ E and r > 0 such that B(y, r) ⊂ O, hence ∆f = 0 on O.<br />

Conversely, if ∆f =0onO then<br />

�<br />

is constant <strong>in</strong> r, hence<br />

�<br />

f(y) = lim<br />

ρ→0<br />

S(y,ρ)<br />

f(x)σ<br />

S(y,r)<br />

y r (dx)<br />

f(x)σ y ρ (dx) =<br />

�<br />

f(x)σ<br />

S(y,r)<br />

y r (dx), r > 0.<br />

The proof is similar <strong>in</strong> the case of superharmonic functions. ⊓⊔


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 9<br />

The fundamental harmonic functions based at y ∈ Rn are the functions which<br />

are harmonic on Rn \{y} and depend only on r = �x−y�, y ∈ Rn .Theysatisfy<br />

the Laplace equation<br />

∆h(x) =0, x ∈ R n ,<br />

<strong>in</strong> spherical coord<strong>in</strong>ates, with<br />

∆h(r) = d2h − 1) dh<br />

(r)+(n<br />

dr2 r dr (r).<br />

In case n = 2, the fundamental harmonic functions are given by the logarithmic<br />

potential<br />

⎧<br />

⎪⎨ −<br />

hy(x) =<br />

⎪⎩<br />

1<br />

log �x − y�, x �= y,<br />

s2<br />

+∞, x = y,<br />

and by the Newton potential kernel <strong>in</strong> case n ≥ 3:<br />

⎧<br />

⎪⎨<br />

1 1<br />

, x �= y,<br />

(n − 2)sn �x − y�n−2 hy(x) =<br />

⎪⎩<br />

+∞, x = y.<br />

More generally, for a ∈ R and y ∈ R n , the function<br />

x ↦→ �x − y� a ,<br />

(2.5)<br />

(2.6)<br />

is superharmonic on R n , n ≥ 3, if and only if a ∈ [2 − n, 0], and harmonic<br />

when a =2− n.<br />

We now focus on the Dirichlet problem on the ball E = B(y, r). We consider<br />

<strong>in</strong> case n =2,and<br />

if n ≥ 3, and let<br />

h0(r) =− 1<br />

h0(r) =<br />

s2<br />

log(r), r > 0,<br />

1<br />

, r > 0,<br />

(n − 2)snrn−2 x ∗ := y +<br />

r2 (x − y)<br />

�y − x�2


10 N. Privault<br />

denote the <strong>in</strong>verse of x ∈ B(y, r) with respect to the sphere S(y, r). Note the<br />

relation<br />

�z − x ∗ �<br />

�<br />

� = �<br />

r<br />

�z − y −<br />

2<br />

�<br />

�<br />

(x − y) �<br />

�y − x�2 �<br />

�<br />

�<br />

r �<br />

= �<br />

�x − y�<br />

r �<br />

�x − y� � (z − y) − (x − y) �<br />

r<br />

�y − x� � ,<br />

hence for all z ∈ S(y, r) andx ∈ B(y, r),<br />

�z − x ∗ �x − z�<br />

� = r , (2.7)<br />

�x − y�<br />

s<strong>in</strong>ce � �<br />

��� z − y x − y �<br />

�x − y� − r �<br />

�z − y� �y − x� � = �x − z�.<br />

The function z ↦→ hx(z) isnotC2 , hence not harmonic, on ¯ B(y, r). Instead of<br />

hx we will use z ↦→ hx∗(z), which is harmonic on a neighborhood of ¯ B(y, r),<br />

to construct a solution of the Dirichlet problem on B(y, r).<br />

Lemma 2.3. The solution of the Dirichlet problem (2.10) for E = B(y, r)<br />

with boundary condition hx, x ∈ B(y, r), is given by<br />

� � ∗ �y − x��z − x �<br />

x ↦→ h0<br />

, z ∈ B(y, r).<br />

r<br />

Proof. We have if n ≥ 3:<br />

� � ∗ �y − x��z − x �<br />

h0<br />

=<br />

r<br />

=<br />

rn−2 (n − 2)sn�x − y�n−2 1<br />

�z − x∗�n−2 rn−2 hx∗(z), �x − y�n−2 and if n =2:<br />

� � ∗ �y − x��z − x �<br />

h0<br />

= −<br />

r<br />

1<br />

� � ∗ �y − x��z − x �<br />

log<br />

s2<br />

r<br />

� �<br />

1 �y − x�<br />

= hx∗(y) − log<br />

.<br />

s2 r<br />

This function is harmonic <strong>in</strong> z ∈ B(y, r) and is equal to hx on S(y, r)<br />

from (2.7). ⊓⊔<br />

Figure 3 represents the solution of the Dirichlet problem with boundary condition<br />

hx on S(y, r) forn = 2 as obta<strong>in</strong>ed after truncation of Figure 2.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 11<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−3<br />

−2 −1<br />

x 0 1 2 3 4 −3 −2 −1 0 1 2 3 4<br />

y<br />

Fig. 2 Graph of hx and of h0(�y − x��z − x ∗ �/r) whenn =2.<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−1.5<br />

−1−0.5<br />

0<br />

x 0.5<br />

1<br />

1.5<br />

2<br />

Fig. 3 Solution of the Dirichlet problem.<br />

−2<br />

2<br />

1<br />

1.5<br />

0.5<br />

−0.5<br />

0 y<br />

−1<br />

−1.5<br />

2.3 Representation of a Function on E from its Values<br />

on ∂E<br />

As mentioned <strong>in</strong> the <strong>in</strong>troduction, it can be of <strong>in</strong>terest to compute a repartition<br />

of charges <strong>in</strong>side a doma<strong>in</strong> E from the values of the field generated on<br />

the boundary ∂E.<br />

In this section we present the representation formula for the values of<br />

an harmonic function <strong>in</strong>side an arbitrary doma<strong>in</strong> E as an <strong>in</strong>tegral over its<br />

boundary ∂E, that follows from the Green identity. This formula uses a kernel<br />

which can be explicitly computed <strong>in</strong> some special cases, e.g. when E = B(y, r)<br />

is an open ball, <strong>in</strong> which case it is called the Poisson formula, cf. Section 2.4<br />

below.<br />

Assume that E is an open doma<strong>in</strong> <strong>in</strong> R n with smooth boundary ∂E and let<br />

∂nf(x) =〈n(x), ∇f(x)〉<br />

denote the normal derivative of f on ∂E.


12 N. Privault<br />

Apply<strong>in</strong>g the divergence theorem (2.2) to the products u(x)∇v(x) and<br />

v(x)∇u(x), where u, v are C2 functions on E, yields Green’s identity:<br />

�<br />

�<br />

(u(x)∆v(x)−v(x)∆u(x))dx = (u(x)∂nv(x)−v(x)∂nu(x))dσ(x). (2.8)<br />

E<br />

∂E<br />

On the other hand, tak<strong>in</strong>g u = 1 <strong>in</strong> the divergence theorem yields Gauss’s<br />

<strong>in</strong>tegral theorem �<br />

∂nv(x)dσ(x) =0, (2.9)<br />

∂E<br />

provided v is harmonic on E.<br />

In the next def<strong>in</strong>ition, hx denotes the fundamental harmonic function def<strong>in</strong>ed<br />

<strong>in</strong> (2.5) and (2.6).<br />

Def<strong>in</strong>ition 2.4. The Green kernel G E (·, ·) onE is def<strong>in</strong>ed as<br />

G E (x, y) :=hx(y) − wx(y), x,y ∈ E,<br />

where for all x ∈ Rn , wx is a smooth solution to the Dirichlet problem<br />

⎧<br />

⎨ ∆wx(y) =0, y ∈ E,<br />

⎩<br />

wx(y) =hx(y), y ∈ ∂E.<br />

(2.10)<br />

In the case of a general boundary ∂E, the Dirichlet problem may have no<br />

solution, even when the boundary value function f is cont<strong>in</strong>uous. Note that<br />

s<strong>in</strong>ce (x, y) ↦→ hx(y) is symmetric, the Green kernel is also symmetric <strong>in</strong> two<br />

variables, i.e.<br />

G E (x, y) =G E (y, x), x,y ∈ E,<br />

and GE (·, ·) vanishes on the boundary ∂E. The next proposition provides an<br />

<strong>in</strong>tegral representation formula for C2 functions on E us<strong>in</strong>g the Green kernel.<br />

In the case of harmonic functions, it reduces to a representation from the<br />

values on the boundary ∂E, cf. Corollary 2.6 below.<br />

Proposition 2.5. For all C2 functions u on E we have<br />

�<br />

u(x) = u(z)∂nG E �<br />

(x, z)σ(dz)+ G E (x, z)∆u(z)dz, x ∈ E. (2.11)<br />

∂E<br />

E<br />

Proof. We do the proof <strong>in</strong> case n ≥ 3, the case n = 2 be<strong>in</strong>g similar. Given<br />

x ∈ E, apply Green’s identity (2.8) to the functions u and hx, wherehxis harmonic on E\B(x, r) forr>0 small enough, to obta<strong>in</strong><br />

�<br />

�<br />

hx(y)∆u(y)dy − (hx(y)∂nu(y) − u(y)∂nhx(y))dσ(y)<br />

E\B(x,r)<br />

= 1<br />

n − 2<br />

�<br />

S(x,r)<br />

∂E<br />

� �<br />

1<br />

n − 2<br />

∂nu(y)+ u(y) dσ(y),<br />

snrn−2 snrn−1


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 13<br />

s<strong>in</strong>ce<br />

1 ∂<br />

y ↦→ ∂n<br />

=<br />

�y − x�n−2 ∂ρ ρ2−n<br />

n − 2<br />

|ρ=r = − .<br />

rn−1 In case u is harmonic, from the Gauss <strong>in</strong>tegral theorem (2.9) and the mean<br />

value property of u we get<br />

�<br />

�<br />

hx(y)∆u(y)dy − (hx(y)∂nu(y) − u(y)∂nhx(y))dσ(y)<br />

E\B(x,r)<br />

�<br />

∂E<br />

= u(y)σ x r (dy)<br />

S(x,r)<br />

= u(x).<br />

In the general case we need to pass to the limit as r tends to 0, which gives<br />

the same result:<br />

�<br />

�<br />

u(x) = hx(y)∆u(y)dy + (u(y)∂nhx(y) − hx(y)∂nu(y))dσ(y).<br />

E\B(x,r)<br />

∂E<br />

(2.12)<br />

Our goal is now to avoid us<strong>in</strong>g the values of the derivative term ∂nu(y)on∂E<br />

<strong>in</strong> the above formula. To this end we note that from Green’s identity (2.8)<br />

we have<br />

�<br />

�<br />

wx(y)∆u(y)dy = (wx(y)∂nu(y) − u(y)∂nwx(y))dσ(y)<br />

E<br />

∂E �<br />

= (hx(y)∂nu(y) − u(y)∂nwx(y))dσ(y) (2.13)<br />

∂E<br />

for any smooth solution wx to the Dirichlet problem (2.10). Tak<strong>in</strong>g the difference<br />

between (2.12) and (2.13) yields<br />

�<br />

�<br />

u(x) = (hx(y) − wx(y))∆u(y)dy + u(y)∂n(hx(y) − wx(y))dσ(y)<br />

E �<br />

∂E<br />

= G E �<br />

(x, y)∆u(y)dy + u(y)∂nG E (x, y)dσ(y), x ∈ E.<br />

E<br />

∂E<br />

Us<strong>in</strong>g (2.5) and (2.6), Relation (2.11) can be reformulated as<br />

� �<br />

�<br />

1<br />

1<br />

1<br />

u(x) =<br />

u(z)∂n<br />

−<br />

∂nu(z) σ(dz)<br />

(n − 2)sn ∂E �x − z�n−2 �x − z�n−2 �<br />

1<br />

1<br />

+<br />

∆u(z)dz, x ∈ B(y, r),<br />

(n − 2)sn �x − z�n−2 E<br />

⊓⊔


14 N. Privault<br />

if n ≥ 3, and if n =2:<br />

u(x) =− 1<br />

�<br />

(u(z)∂n log �x − z�−(log �x − z�)∂nu(z))σ(dz)<br />

s2 ∂E �<br />

− (log �x − z�)∆u(z)dz, x ∈ B(y, r).<br />

E<br />

Corollary 2.6. When u is harmonic on E we get<br />

�<br />

u(x) = u(y)∂nG E (x, y)dσ(y), x ∈ E. (2.14)<br />

∂E<br />

As a consequence of Lemma 2.3, the Green kernel GB(y,r) (·,y)relativetothe<br />

ball B(y, r) isgivenforx∈B(y, r) by<br />

G B(y,r) ⎧<br />

−<br />

⎪⎨<br />

(x, z) =<br />

⎪⎩<br />

1<br />

�<br />

r �z − x�<br />

log<br />

s2 �y − x� �z − x∗ �<br />

, z ∈ B(y, r)\{x}, x �= y,<br />

�<br />

− 1<br />

� �<br />

�z − y�<br />

log<br />

, z ∈ B(y, r)\{x}, x = y,<br />

s2 r<br />

+∞, z = x,<br />

if n =2,cf.Figure4and<br />

⎧ �<br />

1<br />

1<br />

−<br />

(n − 2)sn �z − x�n−2 G B(y,r) z ∈ B(y, r)\{x}, x �= y,<br />

⎪⎨ �<br />

(x, z) = 1<br />

(n − 2)sn<br />

⎪⎩<br />

+∞, z = x,<br />

1 1<br />

−<br />

�z − y�n−2 rn−2 z ∈ B(y, r)\{x}, x = y,<br />

r n−2<br />

�x − y� n−2<br />

�<br />

,<br />

1<br />

�z − x∗�n−2 �<br />

,<br />

if n ≥ 3.<br />

The Green kernel on E = R n , n ≥ 3, is obta<strong>in</strong>ed by lett<strong>in</strong>g r go to <strong>in</strong>f<strong>in</strong>ity:<br />

G Rn<br />

(x, z) =<br />

1<br />

(n − 2)sn�z − x� n−2 = hx(z) =hz(x), x,z ∈ R n . (2.15)


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 15<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

−2<br />

−1.5 −1−0.5<br />

0<br />

0.5<br />

1<br />

1.5<br />

2 −2 −1.5 −1 0<br />

−0.5<br />

Fig. 4 Graph of z ↦→ G(x, z) withx ∈ E = B(y, r) andn =2.<br />

2.4 Poisson Formula<br />

0 0.5 1 1.5<br />

Given u a sufficiently <strong>in</strong>tegrable function on S(y, r) weletI B(y,r)<br />

u (x) denote<br />

the Poisson <strong>in</strong>tegral of u over S(y, r), def<strong>in</strong>ed as:<br />

I B(y,r)<br />

u (x) = 1<br />

�<br />

r<br />

snr S(y,r)<br />

2 −�y − x�2 �z − x�n u(z)σ(dz), x ∈ B(y, r).<br />

Next is the Poisson formula obta<strong>in</strong>ed as a consequence of Proposition 2.5.<br />

Theorem 2.7. Let n ≥ 2. Ifuhas cont<strong>in</strong>uous second partial derivatives on<br />

E = ¯ B(y, r) then for all x ∈ B(y, r) we have<br />

�<br />

G B(y,r) (x, z)∆u(z)dz. (2.16)<br />

u(x) =I B(y,r)<br />

u (x)+<br />

B(y,r)<br />

Proof. We use the relation<br />

�<br />

u(x) = u(z)∂nG B(y,r) �<br />

(x, z)σ(dz)+<br />

S(y,r)<br />

B(y,r)<br />

x ∈ B(y, r), and the fact that<br />

z ↦→ ∂nG B(y,r) (x, z) =− 1<br />

�<br />

�y − x�<br />

∂n log<br />

r<br />

=<br />

s2<br />

1 r<br />

(n − 2)<br />

2 −�x− y�2 snr�z − x�2 G B(y,r) (x, z)∆u(z)dz,<br />

�z − x∗ �<br />

�<br />

�z − x�<br />

if n = 2, and similarly for n ≥ 3. ⊓⊔<br />

2


16 N. Privault<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

2<br />

x 1<br />

0 −1 −2<br />

2 1.5 1 0.5 0 y<br />

−0.5<br />

−1<br />

−1.5<br />

Fig. 5 Poisson kernel graphs on S(y, r) forn =2fortwovaluesofx ∈ B(y, r).<br />

When u is harmonic on a neighborhood of ¯ B(y, r) we obta<strong>in</strong> the Poisson<br />

representation formula of u on E = ¯ B(y, r) us<strong>in</strong>g the values of u on the<br />

sphere ∂E = S(y, r), as a consequence of Theorem 2.7.<br />

Corollary 2.8. Assume that u is harmonic on a neighborhood of ¯ B(y, r). We<br />

have<br />

u(x) = 1<br />

�<br />

r<br />

snr S(y,r)<br />

2 −�y − x�2 �z − x�n u(z)σ(dz),<br />

for all n ≥ 2.<br />

x ∈ B(y, r), (2.17)<br />

Similarly, Theorem 2.7 also shows that<br />

u(x) ≤I B(y,r)<br />

u (x), x ∈ B(y, r),<br />

when u is superharmonic on B(y, r). Note also that when x = y, Relation<br />

(2.17) recovers the mean value property of harmonic functions:<br />

1<br />

u(y) =<br />

snrn−1 �<br />

u(z)σ(dz),<br />

S(y,r)<br />

and the correspond<strong>in</strong>g <strong>in</strong>equality for superharmonic functions.<br />

The function<br />

z ↦→ r2 −�x − y� 2<br />

�x − z� n<br />

is called the Poisson kernel on S(y, r) atx ∈ B(y, r) cf. Figure 5.<br />

A direct calculation shows that the Poisson kernel is harmonic on R n \<br />

(S(y, r) ∪{z}):<br />

∆x<br />

r2 −�y − x�2 =0, x /∈S(y, r) ∪{z}, (2.18)<br />

�z − x�n −2


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 17<br />

hence all Poisson <strong>in</strong>tegrals are harmonic functions on B(y, r). Moreover<br />

Theorem 2.17 of du Plessis [dP70] asserts that for any z ∈ S(y, r), lett<strong>in</strong>g<br />

x tend to z without enter<strong>in</strong>g S(y, r) wehave<br />

lim<br />

x→z IB(y,r) u<br />

(x) =u(z).<br />

Hence the Poisson <strong>in</strong>tegral solves the Dirichlet problem (2.10) on B(y, r).<br />

Proposition 2.9. Given f a cont<strong>in</strong>uous function on S(y, r), the Poisson<br />

<strong>in</strong>tegral<br />

I B(y,r)<br />

f (x) = 1<br />

�<br />

r<br />

snr S(y,r)<br />

2 −�x− y�2 �x − z�n f(z)σ(dz), x ∈ B(y, r),<br />

provides a solution of the Dirichlet problem<br />

⎧<br />

⎨ ∆w(x) =0, x ∈ B(y, r),<br />

⎩<br />

w(x) =f(x), x ∈ S(y, r),<br />

with boundary condition f.<br />

Recall that the Dirichlet problem on B(y, r) may not have a solution when<br />

the boundary condition f is not cont<strong>in</strong>uous.<br />

In particular, from Lemma 2.3 we have<br />

for n ≥ 3, and<br />

I B(y,r)<br />

(z) =<br />

hx<br />

I B(y,r)<br />

1<br />

(z) =hx∗(z) − log<br />

hx<br />

s2<br />

rn−2 hx∗(z), x ∈ B(y, r),<br />

�x − y�n−2 � �x − y�<br />

r<br />

�<br />

, x ∈ B(y, r),<br />

for n =2,wherex ∗ denotes the <strong>in</strong>verse of x with respect to S(y, r). This<br />

function solves the Dirichlet problem with boundary condition hx, andthe<br />

correspond<strong>in</strong>g Green kernel satisfies<br />

G B(y,r) (x, z) =hx(z) −I B(y,r)<br />

hx (z), x,z ∈ B(y, r).<br />

When x = y the solution I B(y,r)<br />

(z) oftheDirichletproblemonB(y, r) with<br />

hy<br />

boundary condition hy is constant and equal to (n − 2) −1 s −1<br />

n r 2−n on B(y, r),<br />

andwehavetheidentity<br />

1<br />

snrn−1 �<br />

r<br />

S(y,r)<br />

2 −�y − z�2 �x − z�n σ(dx) =r 2−n , z ∈ B(y, r),<br />

for n ≥ 3, cf. Figure 6.


18 N. Privault<br />

−0.6<br />

−0.8<br />

−1<br />

−1.2<br />

−1.4<br />

−1.6<br />

−1.8<br />

−4<br />

−3 −2 −1<br />

x<br />

0<br />

1<br />

Fig. 6 Reduced of hx relative to D = R 2 \B(y, r) withx = y.<br />

2.5 <strong>Potential</strong>s and Balayage<br />

2<br />

3<br />

In electrostatics, the function x ↦→ hy(x) represents the Newton potential<br />

created at x ∈ Rn by a charge at y ∈ E, and the function<br />

�<br />

x ↦→ hy(x)q(dy)<br />

E<br />

represents the sum of all potentials generated at x ∈ Rn by a distribution<br />

q(dy) of charges <strong>in</strong>side E. In particular, when E = Rn , n ≥ 3, this sum equals<br />

�<br />

x ↦→ G Rn<br />

(x, y)q(dy)<br />

R n<br />

from (2.15), and the general def<strong>in</strong>ition of potentials orig<strong>in</strong>ates from this<br />

<strong>in</strong>terpretation.<br />

Def<strong>in</strong>ition 2.10. Given a measure µ on E,theGreenpotentialofµ is def<strong>in</strong>ed<br />

as the function<br />

x ↦→ G E �<br />

µ(x) := G E (x, y)µ(dy),<br />

where G E (x, y) is the Green kernel on E.<br />

<strong>Potential</strong>s will be used <strong>in</strong> the construction of superharmonic functions, cf.<br />

Proposition 5.1 below. Conversely, it is natural to ask whether a superharmonic<br />

function can be represented as the potential of a measure. In general,<br />

recall (cf. Proposition 2.5) the relation<br />

�<br />

u(x) = u(z)∂nG E �<br />

(x, z)σ(dz)+ G E (x, z)∆u(z)dz, x ∈ E, (2.19)<br />

∂E<br />

E<br />

E<br />

−3<br />

−2<br />

−1<br />

0<br />

1<br />

y<br />

2<br />

3<br />

4


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 19<br />

which yields, if E = B(y, r):<br />

u(x) =I E �<br />

u (x)+ G E (x, z)∆u(z)dz, x ∈ B(y, r), (2.20)<br />

E<br />

where the Poisson <strong>in</strong>tegral IE u is harmonic from (2.18). Relation (2.20) can be<br />

seen as a decomposition of u <strong>in</strong>to the sum of a harmonic function on B(y, r)<br />

and a potential. The general question of represent<strong>in</strong>g superharmonic functions<br />

as potentials is exam<strong>in</strong>ed next <strong>in</strong> Theorem 2.12 below, with application to<br />

the construction of Mart<strong>in</strong> boundaries <strong>in</strong> Section 2.6.<br />

Def<strong>in</strong>ition 2.11. Consider<br />

i) an open subset E of Rn with Green kernel GE ,<br />

ii) a subset D of E, and<br />

iii) a non-negative superharmonic function u on E.<br />

The <strong>in</strong>fimum on E over all non-negative superharmonic function on E which<br />

are (po<strong>in</strong>twise) greater than u on D is called the reduced function of u relative<br />

to D, and denoted by RD u .<br />

In other terms, lett<strong>in</strong>g Φu denote the set of non-negative superharmonic functions<br />

v on E such that v ≥ u on D, wehave<br />

The lower regularization<br />

R D u<br />

:= <strong>in</strong>f{v ∈ Φu}.<br />

ˆR D u (x) := lim <strong>in</strong>f<br />

y→x RD u (y), x ∈ E,<br />

is called the balayage of u, and is also a superharmonic function.<br />

In case D = R n \B(y, r), the reduced function of hy relative to D is given by<br />

R D hy (z) =hy(z)1 {z/∈B(y,r)} + h0(r)1 {z∈B(y,r)}, z ∈ R n , (2.21)<br />

cf. Figure 7. More generally if v is superharmonic then R D v<br />

= IB(y,r)<br />

v<br />

R 2 \D = B(y, r), cf. p. 62 and p. 100 of Brelot [Bre65]. Hence if D =<br />

B c (y, r) =R n \B(y, r) andx ∈ B(y, r), we have<br />

R Bc (y,r)<br />

hx<br />

(z) =I B(y,r)<br />

(z)<br />

hx<br />

� � ∗ �x − y��z − x �<br />

= h0<br />

r<br />

=<br />

rn−2 (n − 2)sn�x − y�n−2 1<br />

�z − x∗ , z ∈ B(y, r).<br />

�n−2 on


20 N. Privault<br />

Us<strong>in</strong>g Proposition 2.5 and the identity<br />

∆R Bc (y,r)<br />

hy<br />

= σ y r ,<br />

<strong>in</strong> distribution sense, we can represent R Bc (y,r)<br />

hy<br />

R Bc (y,r)<br />

hy<br />

(x)<br />

�<br />

= R<br />

S(y,R)<br />

Bc (y,r)<br />

(z)∂nG hy<br />

B(y,R) �<br />

(x, z)σ(dz)+<br />

�<br />

= h0(R)+<br />

G<br />

S(y,r)<br />

B(y,R) (x, z)σ y r (dz), x ∈ B(y, R).<br />

Lett<strong>in</strong>g R go to <strong>in</strong>f<strong>in</strong>ity yields<br />

R Bc �<br />

(y,r)<br />

(x) =<br />

hy<br />

�<br />

=<br />

S(y,r)<br />

on E = B(y, R), R>r,as<br />

G<br />

B(y,R)<br />

B(y,R) (x, z)σ y r (dz)<br />

G Rn<br />

(x, z)σ y r (dz)<br />

hx(z)σ<br />

S(y,r)<br />

y r (dz)<br />

= hx(y), x /∈ B(y, r),<br />

s<strong>in</strong>ce hx is harmonic on Rn \B(y, r), and<br />

R Bc � � � ∗<br />

(y,r)<br />

�y − x��z − x �<br />

(x) = h0<br />

σ hy<br />

S(y,r)<br />

r<br />

y r (dz)<br />

� � ∗ �y − x��y − x �<br />

= h0<br />

r<br />

= h0(r), x ∈ B(y, r),<br />

which recovers the decomposition (2.21).<br />

More generally we have the follow<strong>in</strong>g result, for which we refer to Theorem 7.12<br />

of Helms [Hel69].<br />

Theorem 2.12. If D is a compact subset of E and u is a non-negative superharmonic<br />

function on E then ˆ RD u is a potential, i.e. there exists a measure<br />

µ on E such that<br />

ˆR D �<br />

u (x) = G<br />

E<br />

E (x, y)µ(dy), x ∈ E. (2.22)<br />

If moreover RD v is harmonic on D then µ is supported by ∂D,cf.Theorem6.9<br />

<strong>in</strong> Helms [Hel69].


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 21<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−4<br />

−3<br />

−2<br />

−1<br />

x<br />

0<br />

1<br />

2<br />

Fig. 7 Reduced of hx relative to D = R 2 \B(y, r) withx ∈ B(y, r)\{y}.<br />

2.6 Mart<strong>in</strong> Boundary<br />

3<br />

The Mart<strong>in</strong> boundary theory extends the Poisson <strong>in</strong>tegral representation described<br />

<strong>in</strong> Section 2.4 to arbitrary open doma<strong>in</strong>s with a non smooth boundary,<br />

or hav<strong>in</strong>g no boundary. Given E a connected open subset of Rn ,theMart<strong>in</strong><br />

boundary ∆E of E is def<strong>in</strong>ed <strong>in</strong> an abstract sense as ∆E := Ê\E, whereÊ<br />

is a suitable compactification of E. Our aim is now to show that every nonnegative<br />

harmonic function on E admits an <strong>in</strong>tegral representation us<strong>in</strong>g its<br />

values on the boundary ∆E.<br />

For this, let u be non-negative and harmonic on E, and consider an <strong>in</strong>creas<strong>in</strong>g<br />

sequence (En)n∈N of open sets with smooth boundaries (∂En)n∈N<br />

and compact closures, such that<br />

E =<br />

−3<br />

∞�<br />

En.<br />

n=0<br />

Then the balayage REn u of u relative to En co<strong>in</strong>cides with u on En, andfrom<br />

Theorem 2.12 it can be represented as<br />

R En<br />

�<br />

u (x) = G E (x, y)dµn(y)<br />

where µn is a measure supported by ∂En. Notethats<strong>in</strong>ce<br />

and<br />

En<br />

−2<br />

−1<br />

R En<br />

u (x) =u(x), x ∈ Ek, k ≥ n,<br />

lim<br />

n→∞ GE (x, y) |y∈∂En =0, x ∈ E,<br />

0<br />

1<br />

y<br />

2<br />

3<br />

4


22 N. Privault<br />

the total mass of µn has to <strong>in</strong>crease to <strong>in</strong>f<strong>in</strong>ity:<br />

lim<br />

n→∞ µn(En) =∞.<br />

For this reason one decides to renormalize µn by fix<strong>in</strong>g x0 ∈ E1 and lett<strong>in</strong>g<br />

so that ˜µn has total mass<br />

˜µn(dy) :=G E (x0,y)µn(dy), n ∈ N,<br />

˜µn(En) =u(x0) < ∞,<br />

<strong>in</strong>dependently of n ∈ N. Next, let the kernel Kx0 be def<strong>in</strong>ed as<br />

Kx0(x, y) := GE (x, y)<br />

GE , x,y ∈ E,<br />

(x0,y)<br />

with the relation<br />

ˆR En<br />

�<br />

u (x) = Kx0(x, y)d˜µn(y).<br />

En<br />

The construction of the Mart<strong>in</strong> boundary ∆E of E relies on the follow<strong>in</strong>g<br />

theorem by Constant<strong>in</strong>escu and Cornea, cf. Helms [Hel69], Chapter 12.<br />

Theorem 2.13. Let E denote a non-compact, locally compact space and consider<br />

a family Φ of cont<strong>in</strong>uous mapp<strong>in</strong>gs<br />

f : E → [−∞, ∞].<br />

Then there exists a unique compact space Ê <strong>in</strong> which E is everywhere dense,<br />

such that:<br />

a) every f ∈ Φ can be extended to a function f ∗ on Ê by cont<strong>in</strong>uity,<br />

b) the extended functions separate the po<strong>in</strong>ts of the boundary ∆E = Ê\E<br />

<strong>in</strong> the sense that if x, y ∈ ∆E with x �= y, thereexistsf ∈ Φ such that<br />

f ∗ (x) �= f ∗ (y).<br />

The Mart<strong>in</strong> space is then def<strong>in</strong>ed as the unique compact space Ê <strong>in</strong> which E<br />

is everywhere dense, and such that the functions<br />

{y ↦→ Kx0(x, y) : x ∈ E}<br />

admit cont<strong>in</strong>uous extensions which separate the boundary Ê\E. Such a<br />

compactification Ê of E exists and is unique as a consequence of the<br />

Constant<strong>in</strong>escu-Cornea Theorem 2.13, applied to the family<br />

Φ := {x ↦→ Kx0(x, z) : z ∈ E}.<br />

In this way the Mart<strong>in</strong> boundary of E is def<strong>in</strong>ed as<br />

∆E := Ê\E.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 23<br />

The Mart<strong>in</strong> boundary is unique up to an homeomorphism, namely if x0,<br />

x ′ 0 ∈ E, then<br />

Kx ′ 0 (x, z) = Kx0(x, z)<br />

Kx0(x ′ 0 ,z)<br />

and<br />

Kx0(x, z) = Kx ′ (x, z)<br />

0<br />

Kx ′ 0 (x0,z)<br />

may have different limits <strong>in</strong> Ê′ which still separate the boundary ∆E. For<br />

this reason, <strong>in</strong> the sequel we will drop the <strong>in</strong>dex x0 <strong>in</strong> Kx0(x, z). In the next<br />

theorem, the Poisson representation formula is extended to Mart<strong>in</strong> boundaries.<br />

Theorem 2.14. Any non-negative harmonic function h on E can be represented<br />

as<br />

�<br />

h(x) = K(z,x)ν(dz), x ∈ E,<br />

∆E<br />

where ν is a non-negative Radon measure on ∆E.<br />

Proof. S<strong>in</strong>ce ˜µn(En) is bounded (actually it is constant) <strong>in</strong> n ∈ N we can<br />

extract a subsequence (˜µnk )k∈N converg<strong>in</strong>g vaguely to a measure µ on Ê, i.e.<br />

�<br />

�<br />

lim<br />

k→∞<br />

f(x)˜µnk (dx) = f(x)µ(dx), f ∈Cc(E),<br />

Ê<br />

Ê<br />

see e.g. Ex. 10, p. 81 of Hirsch and Lacombe [HL99]. The cont<strong>in</strong>uity of z ↦→<br />

K(x, z) <strong>in</strong>z ∈ Ê then implies<br />

E<br />

h(x) = lim ˆR<br />

n→∞<br />

En<br />

u (x)<br />

= lim ˆR<br />

k→∞<br />

En k<br />

u (x)<br />

�<br />

= lim<br />

k→∞<br />

K(z,x)νnk<br />

Ê<br />

(dz)<br />

�<br />

= K(z,x)µ(dz), x ∈ E.<br />

Ê<br />

F<strong>in</strong>ally, ν is supported by ∆E s<strong>in</strong>ce for all f ∈Cc(En),<br />

�<br />

�<br />

f(x)µ(dx) = lim f(x)˜µnk (dx) =0.<br />

k→∞<br />

When E = B(y, r) one can check by explicit calculation that<br />

lim<br />

ζ→z Ky(x, ζ) = r2 −�x − y� 2<br />

�z − x� n , z ∈ S(y, r),<br />

Ek<br />

⊓⊔


24 N. Privault<br />

is the Poisson kernel on S(y, r). In this case we have<br />

µ = σ y r ,<br />

which is the normalized surface measure on S(y, r), and the Mart<strong>in</strong> boundary<br />

∆B(y, r) ofB(y, r) equals its usual boundary S(y, r).<br />

3 Markov Processes<br />

3.1 Markov Property<br />

Let C0(R n ) denote the class of cont<strong>in</strong>uous functions tend<strong>in</strong>g to 0 at <strong>in</strong>f<strong>in</strong>ity.<br />

Recall that f is said to tend to 0 at <strong>in</strong>f<strong>in</strong>ity if for all ε>0thereexistsa<br />

compact subset K of R n such that |f(x)| ≤ε for all x ∈ R n \K.<br />

Def<strong>in</strong>ition 3.1. An Rn-valued stochastic process, i.e. a family (Xt)t∈R+ of<br />

random variables on a probability space (Ω,F,P), is a Markov process if for<br />

all t ∈ R+ the σ-fields<br />

F + t := σ(Xs : s ≥ t)<br />

and<br />

Ft := σ(Xs : 0 ≤ s ≤ t).<br />

are conditionally <strong>in</strong>dependent given Xt.<br />

The above condition can be restated by say<strong>in</strong>g that for all A ∈ F + t and<br />

B ∈Ft we have<br />

P (A ∩ B | Xt) =P (A | Xt)P (B | Xt),<br />

cf. Chung [Chu95]. This def<strong>in</strong>ition naturally entails that:<br />

i) (Xt)t∈R+ is adapted with respect to (Ft)t∈R+, i.e. Xt is Ft-measurable,<br />

t ∈ R+, and<br />

ii) Xu is conditionally <strong>in</strong>dependent of Ft given Xt, for all u ≥ t, i.e.<br />

IE[f(Xu) |Ft] =IE[f(Xu) | Xt], 0 ≤ t ≤ u,<br />

for any bounded measurable function f on Rn .<br />

In particular,<br />

P (Xu ∈ A |Ft) =IE[1A(Xu) |Ft] =IE[1A(Xu) | Xt] =P (Xu ∈ A | Xt),<br />

A ∈B(R n ). Processes with <strong>in</strong>dependent <strong>in</strong>crements provide simple examples<br />

of Markov processes. Indeed, if (Xt)t∈R n has <strong>in</strong>dependent <strong>in</strong>crements, then<br />

for all bounded measurable functions f, g we have


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 25<br />

IE[f(Xt1,...,Xtn)g(Xs1,...,Xsn) | Xt]<br />

=IE[f(Xt1−Xt + x,...,Xtn − Xt + x)<br />

×g(Xs1 − Xt + x,...,Xsn − Xt + x)]x=Xt<br />

=IE[f(Xt1 − Xt + x,...,Xtn − Xt + x)]x=Xt<br />

×IE[g(Xs1 − Xt + x,...,Xsn − Xt + x)]x=Xt<br />

=IE[f(Xt1,...,Xtn) | Xt]IE[g(Xs1,...,Xsn) | Xt],<br />

0 ≤ s1 < ···


26 N. Privault<br />

and<br />

�<br />

Ts,tf(x) =<br />

R n<br />

f(y)ps,t(y − x)dy.<br />

In discrete time, a sequence (Xn)n∈N of random variables is called a homogeneous<br />

Markov cha<strong>in</strong> with transition kernel P if<br />

IE[f(Xn) |Fm] =P n−m f(Xm), 0 ≤ m ≤ n. (3.1)<br />

In the sequel we will assume that (Xt)t∈R+ is time homogeneous, i.e. µs,t<br />

depends only on the difference t−s, and we will denote it by µt−s. Inthiscase<br />

the family (T0,t)t∈R+ is denoted by (Tt)t∈R+. It def<strong>in</strong>es a transition semigroup<br />

associated to (Xt)t∈R+, with<br />

�<br />

Ttf(x) =IE[f(Xt) | X0 = x] =<br />

and satisfies the semigroup property<br />

R n<br />

f(y)µt(x, dy), x ∈ R n ,<br />

TtTsf(x) =IE[Tsf(Xt) | X0 = x]<br />

=IE[IE[f(Xt+s) | Xs] | X0 = x]<br />

=IE[IE[f(Xt+s) |Fs] | X0 = x]<br />

=IE[f(Xt+s) | X0 = x]<br />

= Tt+sf(x),<br />

which can be formulated as the Chapman-Kolmogorov equation<br />

�<br />

µs+t(x, A) =µs ∗ µt(x, A) = µs(x, dy)µt(y, A). (3.2)<br />

By <strong>in</strong>duction, us<strong>in</strong>g (3.2) we obta<strong>in</strong><br />

Px((Xt1,...,Xtn) ∈ B1 ×···×Bn)<br />

� �<br />

= ··· µ0,t1(x, dx1) ···µtn−1,tn(xn−1,dxn)<br />

B1<br />

Bn<br />

for 0


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 27<br />

pt(x) =<br />

1<br />

exp<br />

(2πt) n/2<br />

�<br />

− 1<br />

2t �x�2 R n<br />

�<br />

, x ∈ R n .<br />

From now on we assume that (Tt)t∈R+ is a strongly cont<strong>in</strong>uous Feller semigroup,<br />

i.e. a family of positive contraction operators on C0(Rn ) such that<br />

i) Ts+t = TsTt, s, t ≥ 0,<br />

ii) Tt(C0(R n )) ⊂C0(R n ),<br />

iii) Ttf(x) → f(x) ast → 0, f ∈C0(R n ), x ∈ R n .<br />

The resolvent of (Tt)t∈R+ is def<strong>in</strong>ed as<br />

i.e.<br />

Rλf(x) :=<br />

Rλf(x) =IEx<br />

� ∞<br />

0<br />

�� ∞<br />

0<br />

e −λt Ttf(x)dt, x ∈ R n ,<br />

e −λt �<br />

f(Xt)dt , x ∈ R n ,<br />

for sufficiently <strong>in</strong>tegrable f on R n ,whereIEx denotes the conditional expectation<br />

given that {X0 = x}. It satisfies the resolvent equation<br />

Rλ − Rµ =(µ − λ)RλRµ, λ,µ > 0.<br />

We refer to [Kal02] for the follow<strong>in</strong>g result.<br />

Theorem 3.2. Let (Tt)t∈R+ be a Feller semigroup on C0(R n ) with resolvent<br />

Rλ, λ>0. Then there exists an operator A with doma<strong>in</strong> D⊂C0(R n ) such<br />

that<br />

R −1<br />

λ<br />

= λI − A, λ > 0. (3.3)<br />

The operator A is called the generator of (Tt)t∈R+ and it characterizes<br />

(Tt)t∈R+. Furthermore, the semigroup (Tt)t∈R+ is differentiable <strong>in</strong> t for all<br />

f ∈Dand it satisfies the forward and backward Kolmogorov equations<br />

dTtf<br />

dt = TtAf, and<br />

Note that the <strong>in</strong>tegral<br />

� ∞<br />

0<br />

pt(x, y)dt =<br />

dTtf<br />

dt<br />

Γ (n/2 − 1)<br />

= ATtf.<br />

2π n/2 �x − y� n−2<br />

n − 2<br />

=<br />

2sn�x − y�n−2 =(n/2 − 1)hy(x)


28 N. Privault<br />

of the Gaussian transition density function is proportional to the Newton<br />

potential kernel x ↦→ 1/�x − y�n−2 for n ≥ 3. Hence the resolvent R0f associated<br />

to the Gaussian semigroup Ttf(x) = �<br />

Rn f(y)pt(x, y)dy is a potential<br />

<strong>in</strong> the sense of Def<strong>in</strong>ition 2.10, s<strong>in</strong>ce:<br />

R0f(x) =<br />

� ∞<br />

0<br />

=IEx<br />

�<br />

=<br />

R n<br />

Ttf(x)dt<br />

�<br />

f(Bt)dt<br />

�� ∞<br />

0<br />

f(y)<br />

= n − 2<br />

�<br />

2sn<br />

More generally, for λ>0wehave<br />

R n<br />

� ∞<br />

Rλf(x) =(λI − A) −1 f(x)<br />

=<br />

=<br />

=<br />

=<br />

�<br />

=<br />

� ∞<br />

0<br />

� ∞<br />

e<br />

0<br />

−λt e tA f(x)dt<br />

� ∞<br />

0<br />

� ∞<br />

0<br />

R n<br />

0<br />

e −λt Ttf(x)dt<br />

pt(x, y)dtdy<br />

f(y)<br />

dy.<br />

�x − y�n−2 e −λt Ttf(x)dt<br />

�<br />

f(y)e −λt pt(x, y)dydt<br />

R n<br />

f(y)g λ (x, y)dy, x ∈ R n ,<br />

where g λ (x, y) istheλ-potentialkerneldef<strong>in</strong>edas<br />

g λ (x, y) :=<br />

� ∞<br />

0<br />

e −λt pt(x, y)dt, (3.4)<br />

and Rλf is also called a λ-potential.<br />

Recall that the Hille-Yosida theorem also allows one to construct a strongly<br />

cont<strong>in</strong>uous semigroup from a generator A.<br />

In another direction it is possible to associate a Markov process to any<br />

time homogeneous transition function satisfy<strong>in</strong>g µ0(x, dy) =δx(dy) andthe<br />

Chapman-Kolmogorov equation (3.2), cf. e.g. Theorem 4.1.1 of Ethier and<br />

Kurtz [EK86].


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 29<br />

3.3 Hitt<strong>in</strong>g Times<br />

Def<strong>in</strong>ition 3.3. An a.s. non-negative random variable τ is called a stopp<strong>in</strong>g<br />

time with respect to a filtration Ft if<br />

{τ ≤ t} ∈Ft, t > 0.<br />

The σ-algebra Fτ is def<strong>in</strong>ed as the collection of measurable sets A such that<br />

A ∩{τ0. Note that for all s>0wehave<br />

{τ s}, {τ ≥ s} ∈Fτ.<br />

Def<strong>in</strong>ition 3.4. One says that the process (Xt)t∈R+ has the strong Markov<br />

property if for any P -a.s. f<strong>in</strong>ite Ft-stopp<strong>in</strong>g time τ we have<br />

IE[f(X(τ + t)) |Fτ] =IE[f(X(t)) | X0 = x]x=Xτ = Ttf(Xτ), (3.5)<br />

for all bounded measurable f.<br />

The hitt<strong>in</strong>g time τB of a Borel set B ⊂ R n is def<strong>in</strong>ed as<br />

τB =<strong>in</strong>f{t >0 : Xt ∈ B},<br />

with the convention <strong>in</strong>f ∅ =+∞. AsetB such that Px(τB < ∞) = 0 for all<br />

x ∈ R n is said to be polar.<br />

In discrete time it can be easily shown that hitt<strong>in</strong>g times are stopp<strong>in</strong>g<br />

times, from the relation<br />

{τB ≤ n} c = {τB >n} =<br />

n�<br />

{Xk /∈ B} ∈Fn,<br />

however <strong>in</strong> cont<strong>in</strong>uous time the situation is more complicated. From e.g.<br />

Lemma 7.6 of Kallenberg [Kal02], we have that τB is a stopp<strong>in</strong>g time provided<br />

B is closed and (Xt)t∈R+ is cont<strong>in</strong>uous, or B is open and (Xt)t∈R+ is rightcont<strong>in</strong>uous.<br />

Def<strong>in</strong>ition 3.5. ThelastexittimefromB is denoted by lB and def<strong>in</strong>ed as<br />

with lB =0ifτB =+∞.<br />

k=0<br />

lB =sup{t >0 : Xt ∈ B},<br />

We say that B is recurrent if Px(lB =+∞) =1,x ∈ R n ,andthatB is<br />

transient if lB < ∞, P -a.s., i.e. Px(lB =+∞) =0,x ∈ R n .


30 N. Privault<br />

3.4 Dirichlet Forms<br />

Before turn<strong>in</strong>g to a presentation of stochastic calculus we briefly describe the<br />

notion of Dirichlet form, which provides a functional analytic <strong>in</strong>terpretation<br />

of the Dirichlet problem. A Dirichlet form is a positive bil<strong>in</strong>ear form E def<strong>in</strong>ed<br />

on a doma<strong>in</strong> ID dense <strong>in</strong> a Hilbert space H of real-valued functions, such that<br />

i) the space ID, equipped with the norm<br />

�<br />

�f�ID := �f�2 H + E(f,f),<br />

is a Hilbert space, and<br />

ii) for any f ∈ ID we have m<strong>in</strong>(f,1) ∈ ID and<br />

E(m<strong>in</strong>(f,1), m<strong>in</strong>(f,1)) ≤E(f,f).<br />

The classical example of Dirichlet form is given by H = L 2 (R n )and<br />

�<br />

E(f,g) :=<br />

Rn i=1<br />

n� ∂f<br />

∂xi<br />

(x) ∂g<br />

(x)dx.<br />

∂xi<br />

The generator L of E is def<strong>in</strong>ed by Lf = g, f ∈ ID, if for all h ∈ ID we have<br />

E(f,h) =−〈g, h〉H.<br />

It is known, cf. e.g. Bouleau and Hirsch [BH91], that a self-ajo<strong>in</strong>t operator L<br />

on H = L 2 (R n ) with doma<strong>in</strong> Dom(L) is the generator of a Dirichlet form if<br />

and only if<br />

〈Lf,(f − 1) + 〉 L 2 (R n ) ≤ 0, f ∈ Dom(L).<br />

On the other hand, L is the generator of a strongly cont<strong>in</strong>uous semigroup<br />

(Pt)t∈R+ on H = L2 (Rn ) if and only if (Pt)t∈R+ is sub-Markovian, i.e. for all<br />

f ∈ L2 (Rn ),<br />

0 ≤ f ≤ 1=⇒ 0 ≤ Ptf ≤ 1, t ∈ R+.<br />

WerefertoMaandRöckner [MR92] for more details on the connection between<br />

stochastic processes and Dirichlet forms. Com<strong>in</strong>g back to the Dirichlet<br />

problem ⎧<br />

⎨ ∆u =0, x ∈ D,<br />

⎩<br />

u(x) =0, x ∈ ∂D.<br />

If f and g are C1 with compact support <strong>in</strong> D we have<br />

�<br />

n�<br />

�<br />

g(x)∆f(x)dx = −<br />

∂f<br />

(x)<br />

∂xi<br />

∂g<br />

(x)dx = −E(f,g).<br />

∂xi<br />

D<br />

i=1<br />

D


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 31<br />

Here, ID is the subspace of functions <strong>in</strong> L 2 (R n ) whose derivative <strong>in</strong> distribution<br />

sense belongs to L 2 (R n ), with norm<br />

�f� 2<br />

ID =<br />

�<br />

Rn |f(x)| 2 �<br />

dx +<br />

Rn n� �<br />

�<br />

� ∂f<br />

�<br />

�<br />

(x) �<br />

∂xi<br />

2<br />

dx.<br />

i=1<br />

Hence the Dirichlet problem for f canbeformulatedas<br />

E(f,g) =0,<br />

for all g <strong>in</strong> the completion of C∞ c (D) with respect to the �·�ID -norm. F<strong>in</strong>ally<br />

we mention the notion of capacity of an open set A, def<strong>in</strong>ed as<br />

C(A) :=<strong>in</strong>f{�u� ID : u ∈ ID and u ≥ 1onA}.<br />

The notion of zero-capacity set is f<strong>in</strong>er than that of zero-measure sets and<br />

gives rise to the notion of properties that hold <strong>in</strong> the quasi-everywhere sense,<br />

cf. Bouleau and Hirsch [BH91].<br />

4 Stochastic Calculus<br />

4.1 Brownian Motion and the Poisson Process<br />

Let (Ω,F,P) be a probability space and (Ft)t∈R+ a filtration, i.e. an <strong>in</strong>creas<strong>in</strong>g<br />

family of sub σ-algebras of F. We assume that (Ft)t∈R+ is cont<strong>in</strong>uous on<br />

the right, i.e.<br />

Ft = �<br />

Fs, t ∈ R+.<br />

s>t<br />

Recall that a process (Mt)t∈R+ <strong>in</strong> L1 (Ω) is called an Ft-mart<strong>in</strong>gale if<br />

IE[Mt|Fs] =Ms, 0≤ s ≤ t.<br />

For example, if (Xt) t∈[0,T ] is a (non homogeneous) Markov process with<br />

semigroup (Ps,t)0≤s≤t≤T satisfy<strong>in</strong>g<br />

Ps,tf(Xs) =IE[f(Xt) | Xs] =IE[f(Xt) |Fs], 0 ≤ s ≤ t ≤ T,<br />

on C 2 b (Rn ) functions, with<br />

Ps,t ◦ Pt,u = Ps,u, 0 ≤ s ≤ t ≤ u ≤ T,<br />

then (Pt,T f(Xt)) t∈[0,T ] is an Ft-mart<strong>in</strong>gale:<br />

IE[Pt,T f(Xt) |Fs] =IE[IE[f(XT ) |Ft] |Fs]<br />

=IE[f(XT ) |Fs]<br />

= Ps,T f(Xs), 0 ≤ s ≤ t ≤ T.


32 N. Privault<br />

Def<strong>in</strong>ition 4.1. A mart<strong>in</strong>gale (Mt)t∈R+ <strong>in</strong> L 2 (Ω) (i.e. IE[|Mt| 2 ] < ∞, t ∈<br />

R+) andsuchthat<br />

is called a normal mart<strong>in</strong>gale.<br />

IE[(Mt − Ms) 2 |Fs] =t − s, 0 ≤ s


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 33<br />

�<br />

α ↦→ IE e i〈ξ,α〉 ℓ2 � �<br />

(N) =IE e i � ∞<br />

n=0 ξnαn<br />

�<br />

=<br />

∞�<br />

e −α2n /2<br />

n=0<br />

1 − 2 =e �α�2<br />

ℓ2 (N), α ∈ ℓ 2 (N),<br />

i.e. 〈ξ,α〉 ℓ 2 (N) is a centered Gaussian random variable with variance �α� 2 ℓ 2 (N) .<br />

Let (en)n∈N denote an orthonormal basis of L 2 (R+).<br />

Def<strong>in</strong>ition 4.4. Given u ∈ L 2 (R+) with decomposition<br />

u =<br />

∞�<br />

〈u, en〉en,<br />

n=0<br />

we let J1 : L 2 (R+) −→ L 2 (R N ,γ N ) be def<strong>in</strong>ed as<br />

We have the isometry property<br />

and<br />

J1(u) =<br />

IE[|J1(u)| 2 ]=<br />

=<br />

�<br />

IE e iJ1(u)�<br />

=<br />

∞�<br />

ξn〈u, en〉.<br />

n=0<br />

∞�<br />

k=0<br />

|〈u, en〉| 2 IE[|ξn| 2 ] (4.2)<br />

∞�<br />

|〈u, en〉| 2<br />

k=0<br />

= �u� 2 L 2 (R+) ,<br />

=<br />

∞�<br />

n=0<br />

∞�<br />

n=0<br />

=exp<br />

�<br />

IE e iξn〈u,en〉�<br />

1 − 2<br />

e 〈u,en〉2<br />

L2 (R + )<br />

�<br />

− 1<br />

2 �u�2L 2 �<br />

(R+) ,<br />

hence J1(u) is a centered Gaussian random variable with variance �u�2 L2 (R+) .<br />

Next is a constructive approach to the def<strong>in</strong>ition of Brownian motion, us<strong>in</strong>g<br />

the decomposition<br />

∞�<br />

1 [0,t] =<br />

� t<br />

en(s)ds.<br />

n=0<br />

en<br />

0


34 N. Privault<br />

B t<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5<br />

-1<br />

-1.5<br />

-2<br />

0 0.2 0.4 0.6 0.8 1<br />

Fig. 8 Sample paths of one-dimensional Brownian motion.<br />

Def<strong>in</strong>ition 4.5. For all t ∈ R+, let<br />

Bt(ω) :=J1(1 [0,t]) =<br />

∞�<br />

ξn(ω)<br />

n=0<br />

� t<br />

0<br />

en(s)ds. (4.3)<br />

Clearly, Bt − Bs = J1(1 [s,t]) is a Gaussian centered random variable with<br />

variance:<br />

IE[(Bt − Bs) 2 ]=IE[|J1(1 [s,t])| 2 ]=�1 [s,t]� 2 L2 (R+) = t − s, (4.4)<br />

cf. Figure 8. Moreover, the isometry formula (4.2) shows that if u1,...,un are<br />

orthogonal <strong>in</strong> L 2 (R+) thenJ1(u1),...,J1(un) are also mutually orthogonal<br />

<strong>in</strong> L 2 (Ω), hence from Corollary 16.1 of Jacod and Protter [JP00], we get the<br />

follow<strong>in</strong>g.<br />

Proposition 4.6. Let u1,...,un be an orthogonal family <strong>in</strong> L 2 (R+), i.e.<br />

〈ui,uj〉 L 2 (R+) =0, 1 ≤ i �= j ≤ n.<br />

Then (J1(u1),...,J1(un)) is a vector of <strong>in</strong>dependent Gaussian centered random<br />

variables with respective variances �u1� 2 L 2 (R+) ,...,�u1� 2 L 2 (R+) .<br />

As a consequence of Proposition 4.6, (Bt)t∈R+ has centered <strong>in</strong>dependent <strong>in</strong>crements<br />

hence it is a mart<strong>in</strong>gale.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 35<br />

Moreover, from Relation (4.4) and Remark 4.2 we deduce the follow<strong>in</strong>g<br />

proposition.<br />

Proposition 4.7. The Brownian motion (Bt)t∈R+ is a normal mart<strong>in</strong>gale.<br />

The n-dimensional Brownian motion will be constructed as (B 1 t ,...,B n t )t∈R+<br />

where (B 1 t )t∈R+, ...,(B n t )t∈R+ are <strong>in</strong>dependent copies of (Bt)t∈R+,cf.Figure9<br />

and Figure 10 below.<br />

The compensated Poisson process provides a second example of normal<br />

mart<strong>in</strong>gale. Let now (τn)n≥1 denote a sequence of <strong>in</strong>dependent and identically<br />

exponentially distributed random variables, with parameter λ>0, i.e.<br />

IE[f(τ1,...,τn)] = λ n<br />

� ∞<br />

0<br />

···<br />

� ∞<br />

0<br />

e −λ(s1+···+sn) f(s1,...,sn)ds1 ···dsn,<br />

for all sufficiently <strong>in</strong>tegrable measurable f : R n + → R. Letalso<br />

Tn = τ1 + ···+ τn, n ≥ 1.<br />

We now consider the canonical po<strong>in</strong>t process associated to the family (Tk)k≥1<br />

of jump times.<br />

Def<strong>in</strong>ition 4.8. The po<strong>in</strong>t process (Nt)t∈R+ def<strong>in</strong>ed as<br />

Nt :=<br />

∞�<br />

1 [Tk,∞)(t), t ∈ R+ (4.5)<br />

k=1<br />

is called the standard Poisson po<strong>in</strong>t process.<br />

The process (Nt)t∈R+ has <strong>in</strong>dependent <strong>in</strong>crements which are distributed accord<strong>in</strong>g<br />

to the Poisson law, i.e. for all 0 ≤ t0 ≤ t1 < ···


36 N. Privault<br />

The compensated compound Poisson mart<strong>in</strong>gale def<strong>in</strong>ed as<br />

is a normal mart<strong>in</strong>gale.<br />

4.2 Stochastic Integration<br />

Mt := Xt − λtIE[Y1]<br />

� , t ∈ R+,<br />

λVar[Y1]<br />

In this section we construct the Itô stochastic <strong>in</strong>tegral of square-<strong>in</strong>tegrable<br />

adapted processes with respect to normal mart<strong>in</strong>gales. The filtration (Ft)t∈R+<br />

is generated by (Mt)t∈R+:<br />

Ft = σ(Ms : 0 ≤ s ≤ t), t ∈ R+.<br />

A process (Xt)t∈R+ is said to be Ft-adapted if Xt is Ft-measurable for all<br />

t ∈ R+.<br />

Def<strong>in</strong>ition 4.9. Let L p<br />

ad (Ω×R+), p ∈ [1, ∞], denote the space of Ft-adapted<br />

processes <strong>in</strong> Lp (Ω × R+).<br />

Stochastic <strong>in</strong>tegrals will be first constructed as <strong>in</strong>tegrals of simple predictable<br />

processes.<br />

Def<strong>in</strong>ition 4.10. Let S be a space of random variables dense <strong>in</strong> L2 (Ω,F,P).<br />

Consider the space P of simple predictable processes (ut)t∈R+ of the form<br />

ut =<br />

n�<br />

i=1<br />

Fi1 (t n i−1 ,t n i ](t), t ∈ R+, (4.6)<br />

where Fi is Ftn -measurable, i =1,...,n.<br />

i−1<br />

One easily checks that the set P of simple predictable processes forms a l<strong>in</strong>ear<br />

space. On the other hand, from Lemma 1.1 of Ikeda and Watanabe [IW89],<br />

p. 22 and p. 46, the space P of simple predictable processes is dense <strong>in</strong><br />

L p<br />

ad (Ω × R+) for all p ≥ 1.<br />

Proposition 4.11. The stochastic <strong>in</strong>tegral with respect to the normal mart<strong>in</strong>gale<br />

(Mt)t∈R+, def<strong>in</strong>ed on simple predictable processes (ut)t∈R+ of the form<br />

(4.6) by<br />

� ∞<br />

0<br />

utdMt :=<br />

n�<br />

Fi(Mti − Mti−1), (4.7)<br />

i=1<br />

extends to u ∈ L2 ad (Ω × R+) via the isometry formula<br />

�� ∞ � ∞ � �� ∞ �<br />

IE utdMt vtdMt =IE utvtdt . (4.8)<br />

0<br />

0<br />

0


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 37<br />

Proof. We start by show<strong>in</strong>g that the isometry (4.8) holds for the simple<br />

predictable process u = �n i=1 Gi1 (ti−1,ti], with0=t0


38 N. Privault<br />

= GIE [(Mb − Mt)|Ft]+GIE [(Mt − Ma)|Ft]<br />

= G(Mt − Ma)<br />

=<br />

� ∞<br />

0<br />

1 [0,t](s)usdMs.<br />

ii) If 0 ≤ t ≤ a we have for all bounded Ft-measurable random variable F :<br />

� � ∞ �<br />

IE F usdMs =IE[FG(Mb − Ma)] = 0,<br />

hence<br />

IE<br />

�� ∞<br />

0<br />

0<br />

� �<br />

�<br />

usdMs�Ft<br />

=IE[G(Mb− Ma)|Ft] =0=<br />

� ∞<br />

0<br />

1 [0,t](s)usdMs.<br />

This statement is extended by l<strong>in</strong>earity and density, s<strong>in</strong>ce from the cont<strong>in</strong>uity<br />

of the conditional expectation on L2 we have:<br />

��� t<br />

�� ∞ � ��2<br />

�<br />

IE usdMs − IE usdMs�Ft<br />

0<br />

0<br />

�<br />

= lim<br />

n→∞ IE<br />

��� t<br />

u<br />

0<br />

n �� ∞ � ��2<br />

�<br />

s dMs − IE usdMs�Ft<br />

0<br />

�<br />

= lim<br />

n→∞ IE<br />

�� �� ∞<br />

IE u<br />

0<br />

n � ∞ � �� �<br />

2<br />

�<br />

s dMs − usdMs�Ft<br />

0<br />

≤ lim<br />

n→∞ IE<br />

� ��� ∞<br />

IE u<br />

0<br />

n s dMs<br />

� ∞ � ��<br />

2 ���Ft<br />

− usdMs<br />

0<br />

≤ lim<br />

n→∞ IE<br />

��� ∞<br />

(u<br />

0<br />

n � �<br />

2<br />

s − us)dMs<br />

= lim<br />

n→∞ IE<br />

�� ∞<br />

|u<br />

0<br />

n s − us| 2 �<br />

ds<br />

=0.<br />

In particular, s<strong>in</strong>ce F0 = {∅,Ω}, the Itô <strong>in</strong>tegral is a centered random variable<br />

by Proposition 4.12, i.e.<br />

�� ∞ �<br />

IE usdMs =0. (4.9)<br />

0<br />

The follow<strong>in</strong>g is an immediate corollary of Proposition 4.12.<br />

⊓⊔


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 39<br />

�� t<br />

Corollary 4.13. The <strong>in</strong>def<strong>in</strong>ite stochastic <strong>in</strong>tegral 0 usdMs<br />

�<br />

of u ∈<br />

t∈R+<br />

L2 ad (Ω × R+) is a mart<strong>in</strong>gale, i.e.:<br />

�� t � � � s<br />

�<br />

IE urdMr�Fs<br />

= urdMr, 0 ≤ s ≤ t.<br />

0<br />

0<br />

Clearly, from Def<strong>in</strong>itions 4.5 and (4.7), J1(u) co<strong>in</strong>cides with the s<strong>in</strong>gle stochastic<br />

<strong>in</strong>tegral I1(u) with respect to (Bt)t∈R+.<br />

On the other hand, s<strong>in</strong>ce the standard compensated Poisson mart<strong>in</strong>gale<br />

(Mt)t∈R+ =(Nt− t)t∈R+ is a normal mart<strong>in</strong>gale, the <strong>in</strong>tegral<br />

� T<br />

utdMt<br />

0<br />

is also def<strong>in</strong>ed <strong>in</strong> the Itô sense, i.e. as an L2 (Ω)-limit of stochastic <strong>in</strong>tegrals<br />

of simple adapted processes.<br />

4.3 Quadratic Variation<br />

We now <strong>in</strong>troduce the notion of quadratic variation for normal mart<strong>in</strong>gales.<br />

Def<strong>in</strong>ition 4.14. The quadratic<br />

([M,M]t)t∈R+ def<strong>in</strong>ed as<br />

variation of (Mt)t∈R+ is the process<br />

Let now<br />

[M,M]t = M 2 t<br />

− 2<br />

� t<br />

0<br />

MsdMs, t ∈ R+. (4.10)<br />

= t}<br />

π n = {0 =t n 0


40 N. Privault<br />

and<br />

hence<br />

[M,M]t n i − [M,M]t n i−1 = M 2 tn i − M 2 tn � n<br />

ti − 2<br />

i−1<br />

tn MsdMs<br />

i−1<br />

=(Mtn i − Mtn i−1 )2 � n<br />

ti +2<br />

tn (Mt<br />

i−1<br />

n i−1<br />

⎡�<br />

IE ⎣<br />

[M,M]t −<br />

⎡�<br />

n�<br />

=IE⎣<br />

i=1<br />

⎡�<br />

n�<br />

=4IE⎣<br />

=4IE<br />

=4IE<br />

≤ 4t|π|.<br />

n�<br />

i=1<br />

(Mt n i − Mt n i−1 )2<br />

� ⎤<br />

2<br />

⎦<br />

[M,M]t n i − [M,M]t n i−1 − (Mt n i − Mt n i−1 )2<br />

� t<br />

1 (tn i−1 ,t<br />

i=1 0<br />

n i ](s)(Ms − Mtn i−1 )dMs<br />

� n<br />

ti i=1 tn (Ms − Mt<br />

i−1<br />

n i−1 )2 �<br />

ds<br />

� n<br />

ti i=1 tn (s − t<br />

i−1<br />

n i−1) 2 �<br />

ds<br />

� n�<br />

� n�<br />

� ⎤<br />

2<br />

⎦<br />

− Ms)dMs,<br />

� ⎤<br />

2<br />

⎦<br />

Proposition 4.16. The quadratic variation of Brownian motion (Bt)t∈R+ is<br />

[B,B]t = t, t ∈ R+.<br />

Proof. (cf. e.g. Protter [Pro05], Theorem I-28). For every subdivision {0 =<br />

tn 0 < ··· < tnn = t = t} we have, by <strong>in</strong>dependence of the <strong>in</strong>crements of<br />

Brownian motion:<br />

⎡�<br />

n�<br />

� ⎤<br />

2<br />

IE ⎣ t −<br />

⎦<br />

i=1<br />

⎡�<br />

n�<br />

=IE⎣<br />

=<br />

n�<br />

i=1<br />

i=1<br />

(Bt n i − Bt n i−1 )2<br />

(Bt n i − Bt n i−1 )2 − (t n i − t n i−1)<br />

(t n i − t n i−1) 2 IE<br />

⎡<br />

⎣<br />

� (Bt n i − Bt n i−1 )2<br />

t n i − tn i−1<br />

� ⎤<br />

2<br />

⎦<br />

� ⎤<br />

2<br />

− 1 ⎦<br />

⊓⊔


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 41<br />

=IE[(Z 2 − 1) 2 ]<br />

n�<br />

i=0<br />

≤ t|π|IE[(Z 2 − 1) 2 ],<br />

(t n i − t n i−1) 2<br />

where Z is a standard Gaussian random variable. ⊓⊔<br />

A simple analysis of the Poisson paths shows that the quadratic variation of<br />

the compensated Poisson process (Mt)t∈R+ =(Nt − t)t∈R+ is<br />

[M,M]t = Nt, t ∈ R+.<br />

Similarly for the compensated compound Poisson mart<strong>in</strong>gale<br />

we have<br />

Mt := Xt − λtIE[Y1]<br />

� , t ∈ R+,<br />

λVar[Y1]<br />

Nt �<br />

[M,M]t = |Yk| 2 , t ∈ R+.<br />

k=1<br />

Def<strong>in</strong>ition 4.17. The angle bracket 〈M,M〉t is the unique <strong>in</strong>creas<strong>in</strong>g process<br />

such that<br />

M 2 t −〈M,M〉t, t ∈ R+,<br />

is a mart<strong>in</strong>gale.<br />

As a consequence of Remark 4.3 we have<br />

〈M,M〉t = t, t ∈ R+,<br />

for every normal mart<strong>in</strong>gale. Moreover,<br />

[M,M]t −〈M,M〉t, t ∈ R+,<br />

is also a mart<strong>in</strong>gale as a consequence of Remark 4.3 and Proposition 4.12,<br />

s<strong>in</strong>ce by Def<strong>in</strong>ition 4.14 we have<br />

[M,M]t −〈M,M〉t =[M,M]t − t = M 2 � t<br />

t − t − 2<br />

0<br />

MsdMs, t ∈ R+.<br />

(4.11)<br />

Def<strong>in</strong>ition 4.18. We say that the mart<strong>in</strong>gale (Mt)t∈R+ has the predictable<br />

representation property if any square-<strong>in</strong>tegrable mart<strong>in</strong>gale (Xt)t∈R+ with<br />

respect to (Ft)t∈R+ can be represented as<br />

� t<br />

Xt = X0 +<br />

0<br />

usdMs, t ∈ R+, (4.12)


42 N. Privault<br />

where (ut)t∈R+ ∈ L2 ad (Ω × R+) is an adapted process such that u1 [0,T ] ∈<br />

L2 (Ω × R+) for all T>0.<br />

It is known that Brownian motion and the compensated Poisson process have<br />

the predictable representation property. This is however not true of compound<br />

Poisson processes <strong>in</strong> general.<br />

Def<strong>in</strong>ition 4.19. An equation of the form<br />

[M,M]t = t +<br />

� t<br />

0<br />

φsdMs, t ∈ R+, (4.13)<br />

where (φt)t∈R+ is a square-<strong>in</strong>tegrable adapted process, is called a structure<br />

equation, cf. Emery [ É90].<br />

As a consequence of (4.11) and (4.12) we have the follow<strong>in</strong>g proposition.<br />

Proposition 4.20. Assume that (Mt)t∈R+ is <strong>in</strong> L 4 (Ω) and has the predictable<br />

representation property. Then (Mt)t∈R+ satisfies the structure equation<br />

(4.13), i.e. there exists a square-<strong>in</strong>tegrable adapted process (φt)t∈R+ such<br />

that<br />

[M,M]t = t +<br />

� t<br />

0<br />

φsdMs, t ∈ R+.<br />

Proof. S<strong>in</strong>ce ([M,M]t −t)t∈R+ is a mart<strong>in</strong>gale, the predictable representation<br />

property shows the existence of a square-<strong>in</strong>tegrable adapted process (φt)t∈R+<br />

such that<br />

� t<br />

[M,M]t − t = φsdMs, t ∈ R+.<br />

0<br />

⊓⊔<br />

In particular,<br />

a) the Brownian motion (Bt)t∈R+ satisfies the structure equation (4.13)<br />

with φt = 0, s<strong>in</strong>ce the quadratic variation of (Bt)t∈R+ is [B,B]t = t,<br />

t ∈ R+. Informally we have ∆Bt = ± √ ∆t with equal probabilities 1/2.<br />

b) The compensated Poisson mart<strong>in</strong>gale (Mt)t∈R+ = λ(Nt − t/λ2 )t∈R+,<br />

where (Nt)t∈R+ is a standard Poisson process with <strong>in</strong>tensity 1/λ2 satisfies<br />

the structure equation (4.13) with φt = λ ∈ R, t ∈ R+, s<strong>in</strong>ce<br />

[M,M]t = λ 2 Nt = t + λMt, t ∈ R+.<br />

In this case, ∆Mt ∈{0,λ} with respective probabilities 1 − λ −2 ∆t and<br />

λ −2 ∆t.<br />

The Azéma mart<strong>in</strong>gales correspond to φt = βMt, β ∈ [−2, 0), and provide<br />

examples of processes hav<strong>in</strong>g the chaos representation property, but whose<br />

<strong>in</strong>crements are not <strong>in</strong>dependent, cf. Emery [ É90]. Note that not all normal<br />

mart<strong>in</strong>gales have the predictable representation property. For <strong>in</strong>stance, the<br />

compensated compound Poisson mart<strong>in</strong>gales do not satisfy a structure equation<br />

and do not have the predictable representation property <strong>in</strong> general.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 43<br />

4.4 Itô’s Formula<br />

We consider a normal mart<strong>in</strong>gale (Mt)t∈R+ satisfy<strong>in</strong>g the structure equation<br />

d[M,M]t = dt + φtdMt.<br />

Suchanequationissatisfied<strong>in</strong>particularif(Mt)t∈R+ has the predictable<br />

representation property, cf. Proposition 4.20.<br />

The follow<strong>in</strong>g is a statement of Itô’s formula for normal mart<strong>in</strong>gales, cf.<br />

Emery [ É90], Proposition 2, p. 70.<br />

Proposition 4.21. Assume that φ ∈ L∞ ad (R+×Ω). Let(Xt)t∈R+ be a process<br />

given by<br />

� t � t<br />

Xt = X0 + usdMs + vsds, (4.14)<br />

0<br />

0<br />

where (us)s∈R+, (vs)s∈R+ are adapted processes <strong>in</strong> L2 ad ([0,t]×Ω) for all t>0.<br />

We have for f ∈C1,2 (R+ × R):<br />

f(t, Xt) − f(0,X0) =<br />

If φs =0, the terms<br />

and<br />

� t<br />

0<br />

f(s, Xs− + φsus) − f(s, Xs−) dMs (4.15)<br />

φs<br />

∂f<br />

f(s, Xs + φsus) − f(s, Xs) − φsus (s, Xs)<br />

∂x ds<br />

� t<br />

+<br />

0<br />

� t<br />

∂f<br />

+<br />

0 ∂x (s, Xs)vsds +<br />

� t<br />

f(X s − + φsus) − f(X s −)<br />

φs<br />

0<br />

φ 2 s<br />

∂f<br />

(s, Xs)ds.<br />

∂s<br />

f(Xs + φsus) − f(Xs) − φsusf ′ (Xs)<br />

φ 2 s<br />

can be replaced by their respective limits usf ′ (X s −) and 1<br />

2 u2 s f ′′ (X s −) as<br />

φs → 0.<br />

Examples<br />

i) For the d-dimensional Brownian motion (Bt)t∈R+, φ =0andtheItô<br />

formula reads<br />

� t<br />

f(Bt) =f(B0)+ 〈∇f(Bs),dBs〉 + 1<br />

� t<br />

∆f(Bs)ds,<br />

2<br />

0<br />

0


44 N. Privault<br />

for all C 2 functions f, hence<br />

Ttf(x) =IEx[f(Bt)]<br />

� � t<br />

=IEx f(x)+ 〈∇f(Bs),dBs〉 +<br />

0<br />

1<br />

2<br />

�<br />

=IEx f(x)+ 1<br />

� t �<br />

∆f(Bs)ds<br />

2<br />

= f(x)+ 1<br />

2<br />

= f(x)+ 1<br />

2<br />

� t<br />

0<br />

� t<br />

0<br />

0<br />

IE x [∆f(Bs)] ds<br />

Ts∆f(x)ds,<br />

� t<br />

0<br />

�<br />

∆f(Bs)ds<br />

and as a Markov process, (Bt)t∈R+ has generator 1<br />

2 ∆.<br />

ii) For the compensated Poisson process (Nt − t)t∈R+ we have φs =1,<br />

s ∈ R+, hence<br />

f(Nt − t) =f(0) +<br />

+<br />

� t<br />

0<br />

� t<br />

(f(1 + Ns− − s) − f(Ns− − s))d(Ns − s)<br />

0<br />

(f(1 + Ns − s) − f(Ns − s) − f ′ (Ns − s))ds,<br />

which can actually be recovered by elementary calculus. Hence the generator<br />

of the compensated Poisson process is<br />

Lf(x) =f(x +1)− f(x) − f ′ (x).<br />

From Itô’s formula we have for any stopp<strong>in</strong>g time τ and C2 function u, under<br />

suitable <strong>in</strong>tegrability conditions:<br />

�� τ<br />

�<br />

IE x[u(Bτ)] = u(x)+IEx 〈∇u(Bs),dBs〉 +<br />

0<br />

1<br />

2 IE �� τ �<br />

x ∆u(Bs)ds<br />

0<br />

�� ∞<br />

�<br />

= u(x)+IEx 1 {s≤τ }〈∇u(Bs),dBs〉 +<br />

0<br />

1<br />

2 IE �� τ �<br />

x ∆u(Bs)ds<br />

0<br />

= u(x)+ 1<br />

2 IE �� τ �<br />

x ∆u(Bs)ds , (4.16)<br />

0<br />

which is called Dynk<strong>in</strong>’s formula, cf. Dynk<strong>in</strong> [Dyn65], Theorem 5.1. Let now<br />

σ : R+ × R n → R d ⊗ R n<br />

and<br />

b : R+ × R n → R<br />

satisfy the global Lipschitz condition


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 45<br />

�σ(t, x) − σ(t, y)� 2 + �b(t, x) − b(t, y)� 2 ≤ K 2 �x − y� 2 ,<br />

t ∈ R+, x, y ∈ R n . Then there exists (cf. e.g. [Pro05], Theorem V-7) a unique<br />

strong solution to the stochastic differential equation<br />

� t<br />

Xt = X0 +<br />

0<br />

σ(s, Xs)dBs +<br />

� t<br />

and (Xt)t∈R+ is a Markov process with generator<br />

where a = σ T σ.<br />

L = 1<br />

2<br />

i,j=1<br />

4.5 Killed Brownian Motion<br />

0<br />

i=1<br />

b(s, Xs)ds,<br />

n�<br />

∂<br />

ai,j(t, x)<br />

2 n�<br />

+ bi(t, x)<br />

∂xi∂xj<br />

∂<br />

,<br />

∂xi<br />

Let D be a doma<strong>in</strong> <strong>in</strong> R 2 . The transition operator of the Brownian motion<br />

(B D t ) t∈[0,τ∂D] killed on ∂D is def<strong>in</strong>ed as<br />

q D t (x, A) =Px(B D t ∈ A, τ∂D >t),<br />

where<br />

τ∂D =<strong>in</strong>f{t >0 : Bt ∈ ∂D}<br />

is the first hitt<strong>in</strong>g time of ∂D by (Bt)t∈R+. By the strong Markov property<br />

of Def<strong>in</strong>ition 3.4 we have<br />

Px(Bt ∈ A) =Px(Bt ∈ A, t < τ∂D)+Px(Bt ∈ A, t ≥ τ∂D)<br />

hence<br />

= q D t (x, A)+Px(Bt ∈ A, t ≥ τ∂D)<br />

= q D t (x, A)+IEx[1 {Bt∈A}1 {t≥τ∂D}]<br />

= q D t (x, A)+IEx[IE[1 {Bt∈A}1 {t≥τ∂D} |Fτ∂D ]]<br />

= q D t (x, A)+IEx[1 {t≥τ∂D}IE[1 {Bt∈A}1 {t≥τ∂D} |Fτ∂D ]]<br />

= q D t (x, A)+IEx[1 {t≥τ∂D}IE[1 {Bt−τ∂D ∈A} | B0 = x]x=Bτ ∂D ]<br />

= q D t (x, A)+IEx[Tt−τ∂D 1A(B D τ∂D )1 {t≥τ∂D}],<br />

q D t (x, A) =Px(Bt ∈ A) − IE x[Tt−τ∂D 1A(B D τ∂D )1 {t≥τ∂D}],<br />

and the killed process has the transition densities<br />

p D t (x, y) :=pt(x, y) − IE x[pt−τ∂D (BD τ∂D ,y)1 {t≥τ∂D}], x,y ∈ D, t > 0.<br />

(4.17)


46 N. Privault<br />

The Green kernel is def<strong>in</strong>ed as<br />

gD(x, y) :=<br />

and the associated Green potential is<br />

with <strong>in</strong> particular<br />

GDµ(x) =<br />

� ∞<br />

p<br />

0<br />

D t (x, y)dt,<br />

� ∞<br />

0<br />

gD(x, y)µ(dy),<br />

�� τ∂D �<br />

GDf(x) :=IE f(Bt)dt<br />

0<br />

when µ(dx) =f(x)dx has density f with respect to the Lebesgue measure.<br />

When D = Rn we have τ∂D = ∞ a.s. hence GRn = G.<br />

Theorem 4.22. The function (x, y) ↦→ gD(x, y) is symmetric and cont<strong>in</strong>uous<br />

on D2 ,andx↦→ gD(x, y) is harmonic on D\{y}, y ∈ D.<br />

Proof. For all bounded doma<strong>in</strong>s A <strong>in</strong> Rn , the function GD1A def<strong>in</strong>ed as<br />

�<br />

�� �<br />

τ∂D<br />

GD1A(x) = gD(x, y)1A(y)dy =IEx 1A(Bt)dt<br />

R n<br />

has the mean value property <strong>in</strong> D\ Ā, and the property extends to gD by<br />

l<strong>in</strong>ear comb<strong>in</strong>ations and an limit<strong>in</strong>g argument. ⊓⊔<br />

From (4.17), the associated λ-potential kernel (3.4) is given by<br />

where<br />

and<br />

g λ (x, y) =<br />

� ∞<br />

0<br />

e −λt pt(x, y)dt = g λ �<br />

D (x, y)+<br />

g λ D (x, y) :=<br />

� ∞<br />

e<br />

0<br />

−λt q D t<br />

0<br />

g λ (z,y)h λ D (x, dz), (4.18)<br />

(x, y)dt,<br />

h λ D (x, A) =IEx[e −λτ∂D 1{B D τ∂D ∈A}1 {τ∂D


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 47<br />

that harmonic functions satisfy the mean value property, can be recovered<br />

us<strong>in</strong>g stochastic calculus. Let<br />

τr =<strong>in</strong>f{t ∈ R+ : Bt ∈ S(x, r)}<br />

denote the first exit time of (Bt)t∈R+ from the open ball B(x, r).<br />

Due to the spatial symmetry of Brownian motion, Bτr is uniformly distributed<br />

on S(x, r), hence<br />

�<br />

IE x[u(Bτr)] = u(x)σ r x (dy),<br />

S(x,r)<br />

where σ r x denotes the uniform surface measure on the sphere S(x, r).<br />

On the other hand, from Dynk<strong>in</strong>’s formula (4.16) we have<br />

IE x[u(Bτr)] = u(x)+ 1<br />

2 IE �� �<br />

τr<br />

x ∆u(Bs)ds .<br />

Hence the condition ∆u = 0 implies the mean value property<br />

�<br />

u(x) =<br />

S(x,r)<br />

0<br />

u(y)σ r y (dy), (5.1)<br />

and similarly the condition ∆u ≤ 0 implies<br />

�<br />

u(x) ≥ u(y)σ r y(dy). (5.2)<br />

S(x,r)<br />

From Remark 3, page 134 of Dynk<strong>in</strong> [Dyn65], for all n ≥ 1wehave<br />

1<br />

∆u(x) = lim<br />

2 n→∞<br />

IE x[u(Bτ )] − u(x)<br />

1/n<br />

, (5.3)<br />

IE x[τ1/n] which shows conversely that (5.1), resp. (5.2), implies ∆u ≤ 0, resp ∆u =0,<br />

which recovers Proposition 2.2.<br />

Next, we recover the superharmonicity property of the potential<br />

� ∞<br />

R0f(x) = Ttf(x)dt (5.4)<br />

0<br />

�� ∞ �<br />

=IEx<br />

0<br />

� ∞ �<br />

f(Bt)dt<br />

=<br />

0 Rn f(y)pt(x, y)dtdy<br />

= n − 2<br />

�<br />

2<br />

hx(y)f(y)dy, x ∈ R n .<br />

R n


48 N. Privault<br />

Proposition 5.1. Let f be a non-negative function on Rn . Then the potential<br />

R0f given <strong>in</strong> (5.4) is a superharmonic function provided it is C2 on Rn .<br />

Proof. For all r>0, us<strong>in</strong>g the strong Markov property (3.5) we have<br />

�� � �� τr<br />

∞ �<br />

g(x) =IEx f(Bt)dt +IEx f(Bt)dt<br />

0<br />

τr<br />

�� � � �� τr<br />

∞ � ��<br />

�<br />

=IEx f(Bt)dt +IEx IE f(Bt)dt�Bτr<br />

=IEx<br />

0<br />

�� τr<br />

0<br />

≥ IE x[g(Bτr)]<br />

�<br />

=<br />

τr<br />

�<br />

f(Bt)dt +IEx[g(Bτr)]<br />

g(y)σ<br />

S(x,r)<br />

r x (dy),<br />

which shows that g is ∆-superharmonic from Proposition 2.2. ⊓⊔<br />

Other non-negative superharmonic functionals can also be constructed by<br />

convolution, i.e. if f is ∆-superharmonic and g is non-negative and sufficiently<br />

<strong>in</strong>tegrable, then<br />

�<br />

x ↦→ g(y)f(x − y)dy<br />

R n<br />

is non-negative and ∆-superharmonic.<br />

We now turn to an example <strong>in</strong> discrete time, with the notation of (3.1).<br />

Here a function f is called harmonic when (I − P )f = 0, and superharmonic<br />

if (I − P )f ≥ 0.<br />

Proposition 5.2. Afunctionfissuperharmonic if and only if the sequence<br />

(f(Xn))n∈N is a supermart<strong>in</strong>gale.<br />

Proof. We have<br />

5.2 Dirichlet Problem<br />

IE[f(Xm) |Fn] =IE[f(Xm−n) | X0 = x]x=Xn<br />

= Pm−nf(Xn)<br />

≤ f(Xn).<br />

In this section we revisit the Dirichlet problem us<strong>in</strong>g probabilistic tools.<br />

Theorem 5.3. Consider an open doma<strong>in</strong> <strong>in</strong> R n and f afunctiononR n ,and<br />

assume that the Dirichlet problem<br />

⊓⊔


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 49<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5<br />

-1<br />

-1.5<br />

-2<br />

-2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5<br />

Fig. 9 Sample paths of a two-dimensional Brownian motion.<br />

has a C 2 solution u. Then we have<br />

where<br />

⎧<br />

⎨ ∆u =0, x ∈ D,<br />

⎩<br />

u(x) =f(x), x ∈ ∂D,<br />

u(x) =IEx[f(Bτ∂D )], x ∈ ¯ D,<br />

τ∂D =<strong>in</strong>f{t >0 : Bt ∈ ∂D}<br />

is the first hitt<strong>in</strong>g time of ∂D by (Bt)t∈R+.<br />

Proof. For all r>0 such that B(x, r) ⊂ D we have<br />

u(x) =IEx[f(Bτ∂D )]<br />

=IEx[IE[f(Bτ∂D ) | Bτr]]<br />

=IEx[u(Bτr)]<br />

�<br />

= u(y)σ r x(dy), x ∈ ¯ D,<br />

S(x,r)<br />

hence u has the mean value property, thus ∆u = 0 by Proposition 2.2. ⊓⊔


50 N. Privault<br />

5.3 Poisson Equation<br />

In the next theorem we show that the resolvent<br />

�� ∞<br />

Rλf(x) =IEx e −λt �<br />

f(Bt)dt<br />

=<br />

� ∞<br />

0<br />

0<br />

e −λt Ttf(x)dt, x ∈ R n ,<br />

solves the Poisson equation. This is consistent with the fact that Rλ =(λI −<br />

∆/2) −1 , cf. Relation (3.3) <strong>in</strong> Theorem 3.2.<br />

Theorem 5.4. Let λ>0 and f a non-negative function on R n , and assume<br />

that the Poisson equation<br />

has a C 2 b<br />

1<br />

2 ∆u(x) − λu(x) =−f(x), x ∈ Rn , (5.5)<br />

solution u. Then we have<br />

�� ∞<br />

u(x) =IEx e −λt �<br />

f(Bt)dt , x ∈ R n .<br />

Proof. By Itô’s formula,<br />

e −λt u(Bt) =u(B0)+<br />

+ 1<br />

2<br />

� t<br />

0<br />

= u(B0)+<br />

� t<br />

0<br />

0<br />

e −λs 〈∇u(Bs),dBs〉<br />

e −λs ∆u(Bs)ds − λ<br />

� t<br />

0<br />

� t<br />

e −λs 〈∇u(Bs),dBs〉− 1<br />

2<br />

0<br />

0<br />

e −λs u(Bs)ds<br />

� t<br />

0<br />

e −λs f(Bs)ds,<br />

hence by tak<strong>in</strong>g expectations on both sides and us<strong>in</strong>g (4.9) we have<br />

e −λt IE x[u(Bt)] = u(x) − 1<br />

2 IE �� t<br />

x e<br />

0<br />

−λs �<br />

f(Bs)ds ,<br />

and lett<strong>in</strong>g t tend to <strong>in</strong>f<strong>in</strong>ity we get<br />

0=u(x) − 1<br />

2 IE �� ∞<br />

x e −λs �<br />

f(Bs)ds .<br />

When λ = 0, a similar result holds under the addition of a boundary condition<br />

on a smooth doma<strong>in</strong> D.<br />

⊓⊔


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 51<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5<br />

-1<br />

-1.5<br />

-2 -2 -1.5<br />

-1<br />

-0.5 0<br />

0.5 1 1.5 2<br />

Fig. 10 Sample paths of a three-dimensional Brownian motion.<br />

2<br />

1<br />

1.5<br />

-2<br />

-1.5<br />

-1<br />

-0.5<br />

0<br />

0.5<br />

Theorem 5.5. Let f a non-negative function on Rn , and assume that the<br />

Poisson equation ⎧<br />

⎪⎨ 1<br />

∆u(x) =−f(x), x ∈ D,<br />

2<br />

⎪⎩<br />

u(x) =0, x ∈ ∂D,<br />

(5.6)<br />

has a C 2 b<br />

solution u. Then we have<br />

u(x) =IEx<br />

�� τ∂D<br />

0<br />

�<br />

f(Bt)dt , x ∈ D.<br />

Proof. Similarly to the proof of Theorem 5.4 we have<br />

u(Bt) =u(B0)+<br />

= u(B0)+<br />

� t<br />

0<br />

� t<br />

0<br />

〈∇u(Bs),dBs〉 + 1<br />

2<br />

〈∇u(Bs),dBs〉−<br />

� t<br />

0<br />

� t<br />

0<br />

∆u(Bs)ds<br />

f(Bs)ds,<br />

hence<br />

0=IEx[u(Bτ∂D )] = u(x) − IE � � �<br />

τ∂D<br />

x l f(Bs)ds .<br />

0<br />

We easily check that the boundary condition u(x) =0issatisfiedonx∈∂D s<strong>in</strong>ce τ∂D = 0 a.s. given that B0 ∈ ∂D. ⊓⊔<br />

The probabilistic <strong>in</strong>terpretation of the solution of the Poisson equation can<br />

also be formulated <strong>in</strong> terms of a Brownian motion killed at the boundary ∂D,<br />

cf. Section 4.5.


52 N. Privault<br />

Proposition 5.6. The solution u of (5.6) is given by<br />

�<br />

u(x) =GDf(x) = gD(x, y)f(y)dy, x ∈ D,<br />

R n<br />

where gD and GDf are the Green kernel and the Green potential of Brownian<br />

motion killed on ∂D.<br />

Proof. We first need to show that the function gD is the fundamental solution<br />

of the Poisson equation. We have<br />

�� � �� τ∂D<br />

τ∂D<br />

IE x 1A(Bt)dt =IEx 1A(B<br />

0<br />

0<br />

D t )dt<br />

�<br />

�<br />

hence<br />

u(x) =IEx<br />

�� τ∂D<br />

0<br />

�� τ∂D<br />

0<br />

�� ∞<br />

1 {BD t ∈A}dt<br />

�<br />

1 {BD t ∈A}dt<br />

0<br />

� ∞<br />

Px(B<br />

0<br />

D t ∈ A)dt<br />

� ∞<br />

q<br />

0<br />

D t (x, A)dt<br />

� ∞ �<br />

p<br />

0 A<br />

D t (x, y)dydt<br />

=IEx<br />

=IEx<br />

=<br />

=<br />

=<br />

= GD1A(x)<br />

�<br />

= gD(x, y)1A(y)dy,<br />

R n<br />

f(B D �<br />

�<br />

t )dt = GDf(x) =<br />

Rn gD(x, y)f(y)dy.<br />

The solution of (5.6) can also be expressed us<strong>in</strong>g the λ-potential kernel g λ .<br />

Proposition 5.7. The solution u of (5.5) can be represented as<br />

� ∞<br />

u(x) =<br />

0<br />

e −λt<br />

�<br />

Rn q D �<br />

t (x, y)f(y)dydt<br />

�<br />

+ f(y)<br />

x ∈ D.<br />

R n<br />

Rn g λ (z,y)IEx[e −λτ∂D1{BD τ∂D∈dz}1 {τ∂D


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 53<br />

Proof. From (4.18) we have<br />

u(x) =Rλf(x)<br />

�<br />

=<br />

Rn g λ (x, y)f(y)dy<br />

�<br />

=<br />

Rn g λ �<br />

D (x, y)f(y)dy +<br />

� ∞<br />

= e<br />

0<br />

−λt<br />

�<br />

Rn � �<br />

+ f(y)<br />

R n<br />

R n<br />

q D t (x, y)f(y)dydt<br />

�<br />

f(y)<br />

Rn g λ (z,y)h λ D (x, dz)dy<br />

Rn g λ (z,y)IEx[e −λτ∂D1{BD τ∂D∈dz}1 {τ∂D


54 N. Privault<br />

5.4 Cauchy Problem<br />

This section presents a version of the Feynman-Kac formula. Consider the<br />

partial differential equation (PDE)<br />

⎧<br />

⎪⎨<br />

∂u<br />

(t, x) =1 ∆u(t, x) − V (x)u(t, x)<br />

∂t 2<br />

⎪⎩<br />

u(0,x)=f(x).<br />

(5.7)<br />

Proposition 5.8. Assume that f,V ∈Cb(R n ) and V is non-negative. Then<br />

the solution of (5.7) is given by<br />

u(t, x) =IEx<br />

Proof. Let<br />

We have<br />

˜Ttf(x) =IEx<br />

˜Tt+sf(x) =IEx<br />

� �<br />

exp −<br />

� t<br />

0<br />

� �<br />

exp −<br />

� �<br />

exp −<br />

� �<br />

V (Bs)ds f(Bt) , t ∈ R+, x ∈ R n . (5.8)<br />

� t<br />

0<br />

� t+s<br />

0<br />

� �<br />

V (Bs)ds f(Bt) , t ∈ R+, x ∈ R n .<br />

� �<br />

V (Bu)du f(Bt+s)<br />

� � � t � �<br />

=IEx exp − V (Bu)du exp −<br />

� �<br />

0<br />

� � t+s �<br />

t<br />

=IEx IE exp − V (Bu)du<br />

� � � t<br />

0<br />

� � �<br />

=IEx exp − V (Bu)du IE exp −<br />

� �<br />

0<br />

� t � � �<br />

=IEx exp − V (Bu)du IE exp −<br />

0<br />

� � � t � �<br />

=IEx exp − V (Bu)du ˜Tsf(Bt)<br />

= ˜ Tt ˜ Tsf(x),<br />

0<br />

� t+s<br />

f(Bt+s)<br />

� �<br />

V (Bu)du f(Bt+s)<br />

� ��<br />

�<br />

�Ft<br />

� � ��<br />

�<br />

V (Bu)du f(Bt+s) �Ft<br />

� � �<br />

�<br />

V (Bu)du f(Bs) �B0 = x<br />

� t+s<br />

t<br />

� s<br />

hence ( ˜ Tt)t∈R+ has the semigroup property. Next we have<br />

˜Ttf(x) − f(x)<br />

t<br />

= 1<br />

� � �<br />

IE x exp −<br />

t<br />

= 1<br />

t (IEx[f(Bt)] − f(x)) + 1<br />

t IE x<br />

� t<br />

0<br />

�<br />

exp<br />

0<br />

� � �<br />

V (Bu)du f(Bt) − f(x)<br />

�<br />

−<br />

� t<br />

0<br />

x=Bt<br />

� �<br />

V (Bu)du f(Bt) + o(t),<br />


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 55<br />

hence<br />

d ˜ Tt<br />

dt |t=0f(x) = 1<br />

∆f(x) − V (x)f(x),<br />

2<br />

x ∈ R, (5.9)<br />

and u given by (5.8) satisfies<br />

∂u<br />

∂t (t, x) = d ˜ Tt<br />

dt f(x)<br />

Moreover, (5.9) also yields<br />

�<br />

Rn f(y)pt(x, y)dy = f(x)+ 1<br />

2<br />

= f(x)+ 1<br />

2<br />

= d ˜ Tt<br />

dt |t=0 ˜ Ttf(x)<br />

� �<br />

1<br />

= ∆ − V (x) ˜Ttf(x)<br />

2<br />

= 1<br />

∆u(t, x) − V (x)u(t, x).<br />

2<br />

� t<br />

0<br />

� t<br />

0<br />

�<br />

�<br />

R n<br />

R n<br />

∆xf(y)ps(x, y)dyds<br />

f(y)∆xps(x, y)dyds,<br />

for all sufficiently regular functions f on R n ,wherept(x, y) is the Gaussian<br />

density kernel. Hence we get, after differentiation with respect to t,<br />

∂pt<br />

(x, y) =1<br />

∂t 2 ∆xpt(x, y). (5.10)<br />

From (5.10) we recover the harmonicity of hy on Rn \{y}:<br />

∆xhy(x) =∆x<br />

=<br />

=2<br />

� ∞<br />

0<br />

� ∞<br />

0<br />

� ∞<br />

0<br />

pt(x, y)dt<br />

∆xpt(x, y)dt<br />

∂pt<br />

(x, y)dt<br />

∂t<br />

= 2 lim<br />

t→∞ pt(x, y) − 2 lim<br />

t→0 pt(x, y)<br />

=0,<br />

provided x �= y. The backward Kolmogorov partial differential equation<br />

⎧<br />

⎪⎨ −<br />

⎪⎩<br />

∂u<br />

(t, x) =1 ∆u(t, x) − V (x)u(t, x),<br />

∂t 2<br />

(5.11)<br />

u(T,x)=f(x),<br />

can be similarly solved as follows.<br />

⊓⊔


56 N. Privault<br />

Proposition 5.9. Assume that f,V ∈Cb(Rn ) and V is non-negative. Then<br />

the solution of (5.11) is given by<br />

� � � �<br />

�<br />

T<br />

�<br />

�<br />

u(t, x) =IE exp − V (Bs)ds f(BT ) �Bt = x , (5.12)<br />

t ∈ R+, x ∈ R n .<br />

Proof. Lett<strong>in</strong>g u(t, x) be def<strong>in</strong>ed by (5.12) we have<br />

�<br />

exp −<br />

� t<br />

�<br />

V (Bs)ds u(t, Bt)<br />

0<br />

� t<br />

0<br />

t<br />

�<br />

� � � � � �<br />

T<br />

�<br />

�<br />

=exp − V (Bs)ds IE exp − V (Bs)ds f(BT ) �Ft<br />

0<br />

t<br />

� � � � �<br />

T<br />

�<br />

�<br />

=IE exp − V (Bs)ds f(BT ) �Ft ,<br />

t ∈ R+, x ∈ R n , which is a mart<strong>in</strong>gale by construction. Apply<strong>in</strong>g Itô’s formula<br />

to this process we get<br />

�<br />

exp −<br />

� t<br />

0<br />

= u(0,B0)+<br />

+ 1<br />

2<br />

−<br />

+<br />

� t<br />

0<br />

� t<br />

0<br />

� t<br />

0<br />

�<br />

V (Bs)ds u(t, Bt)<br />

exp<br />

� t<br />

0<br />

�<br />

−<br />

V (Bs)exp<br />

�<br />

exp −<br />

�<br />

exp −<br />

� s<br />

� s<br />

0<br />

0<br />

�<br />

−<br />

� s<br />

0<br />

V (Bτ)dτ<br />

� s<br />

0<br />

V (Bτ )dτ<br />

�<br />

V (Bτ )dτ 〈∇xu(s, Bs),dBs〉<br />

�<br />

∆xu(s, Bs)ds<br />

�<br />

V (Bτ )dτ u(s, Bs)ds<br />

�<br />

∂u<br />

u(s, Bs)ds.<br />

∂s<br />

The mart<strong>in</strong>gale property shows that the absolutely cont<strong>in</strong>uous f<strong>in</strong>ite variation<br />

terms vanish (see e.g. Corollary II-1 of Protter [Pro05]), hence<br />

1<br />

2 ∆xu(s, Bs) − V (Bs)u(s, Bs)+ ∂u<br />

∂s (s, Bs) =0, s ∈ R+,<br />

which leads to (5.11). ⊓⊔


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 57<br />

5.5 Mart<strong>in</strong> Boundary<br />

Our aim is now to provide a probabilistic <strong>in</strong>terpretation of the Mart<strong>in</strong> boundary<br />

<strong>in</strong> discrete time. Namely we use the Mart<strong>in</strong> boundary theory to study<br />

the way a Markov cha<strong>in</strong> with transition operator P leaves the space E, <strong>in</strong><br />

particular when E is a union<br />

E =<br />

∞�<br />

En,<br />

n=0<br />

of transient sets En, n ∈ N. We assume that E is a metric space with distance<br />

δ and that the Cauchy completion of E co<strong>in</strong>cides with its Alexandrov<br />

compactification (E, ˆx).<br />

For simplicity we will assume that the transition operator P is self-adjo<strong>in</strong>t<br />

with respect to a reference measure m. LetnowG denote the potential<br />

Gf(x) :=<br />

with kernel g(·, ·), satisfy<strong>in</strong>g<br />

�<br />

Gf(x) =<br />

E<br />

∞�<br />

P n f(x), x ∈ E.<br />

n=0<br />

f(y)g(x, y)m(dy), x ∈ E,<br />

for a given reference measure m. Fixx0 ∈ E1. Forf <strong>in</strong> the space Cc(E) of<br />

compactly supported function on E, def<strong>in</strong>e the kernel<br />

and the operator<br />

kx0(x, z) =<br />

Kx0f(x) := Gf(x)<br />

g(x, x0) =<br />

g(x, z)<br />

, x,z ∈ E,<br />

g(x, x0)<br />

�<br />

E<br />

kx0(x, z)f(z)m(dz), x ∈ E.<br />

Consider a sequence (fn)n∈N dense <strong>in</strong> Cc(E) and the metric def<strong>in</strong>ed by<br />

d(x, y) :=<br />

∞�<br />

ζn|Kx0fn(x) − Kx0fn(y)|,<br />

n=1<br />

where (ζn)n∈N is a sequence of non-negative numbers such that<br />

∞�<br />

ζn�Kx0fn�∞ < ∞.<br />

n=1


58 N. Privault<br />

The Mart<strong>in</strong> space Ê for X started with distribution rm is constructed as the<br />

Cauchy completion of (E,δ + d). Then Kx0f, f ∈Cc(E), can be extended<br />

by cont<strong>in</strong>uity to Ê s<strong>in</strong>ce for all ε>0thereexistsn∈Nsuch that for all<br />

x, y ∈ E,<br />

|Kx0f(x) − Kx0f(y)|<br />

≤|Kx0f(x) − Kx0fn(x)| + |Kx0fn(x) − Kx0fn(y)| + |Kx0fn(y) − Kx0f(y)|<br />

≤ ε + ζnd(x, y),<br />

and, from Proposition 2.3 of Revuz [Rev75], the sequence (Kx0fn)n∈N is dense<br />

<strong>in</strong> {Kx0f : f ∈Cc(E)}.<br />

If x ∈ ∆E, a sequence (xn)n∈N ⊂ E converges to x if and only if it<br />

converges to po<strong>in</strong>t at <strong>in</strong>f<strong>in</strong>ity ˆx for the metric δ and (Kx0f(xn))n∈N converges<br />

to Kx0f(x) for every f ∈Cc(E). F<strong>in</strong>ally we have the follow<strong>in</strong>g result, for which<br />

we refer to [Rev75].<br />

Theorem 5.10. The sequence (Xn)n∈N converges Px0-a.s. <strong>in</strong> Ê and the law<br />

of X∞ under Px0 is carried by ∆E.<br />

References<br />

[Bas98] R.F. Bass. Diffusions and elliptic operators. <strong>Probability</strong> and its Applications<br />

(New York). Spr<strong>in</strong>ger-Verlag, New York, 1998.<br />

[BH91] N. Bouleau and F. Hirsch. Dirichlet Forms and Analysis on Wiener Space. de<br />

Gruyter, 1991.<br />

[Bia93] Ph. Biane. Calcul stochastique non-commutatif. In Ecole d’Eté deProbabilités<br />

de Sa<strong>in</strong>t-Flour, volume 1608 of Lecture Notes <strong>in</strong> Mathematics. Spr<strong>in</strong>ger-Verlag,<br />

1993.<br />

[Bre65] M. Brelot. Éléments de la théorie classique du potentiel. 3e édition. Les cours<br />

de Sorbonne. 3e cycle. Centre de Documentation Universitaire, Paris, 1965.<br />

Re-edited by Association Laplace-Gauss, 1997.<br />

[Chu95] K.L. Chung. Green, Brown, and probability. World Scientific, 1995.<br />

[Doo84] J.L. Doob. <strong>Classical</strong> potential theory and its probabilistic counterpart. Spr<strong>in</strong>ger-<br />

Verlag, Berl<strong>in</strong>, 1984.<br />

[dP70] N. du Plessis. An <strong>in</strong>troduction to potential theory. Oliver & Boyd, 1970.<br />

[Dyn65] E.B. Dynk<strong>in</strong>. Markov processes. Vols. I, II. Die Grundlehren der Mathematischen<br />

Wissenschaften. Academic Press Inc., New York, 1965.<br />

[ É90] M. Émery. On the Azéma mart<strong>in</strong>gales. In Sém<strong>in</strong>aire de Probabilités XXIII,<br />

[EK86]<br />

volume 1372 of Lecture Notes <strong>in</strong> Mathematics, pages 66–87. Spr<strong>in</strong>ger Verlag,<br />

1990.<br />

S.N. Ethier and T.G. Kurtz. Markov processes, characterization and convergence.<br />

John Wiley & Sons Inc., New York, 1986.<br />

[FOT94] M. Fukushima, Y. Oshima, and M. Takeda.<br />

Markov Processes. de Gruyter, 1994.<br />

Dirichlet Forms and Symmetric<br />

[Hel69] L.L. Helms. Introduction to potential theory, volume XII of Pure and applied<br />

mathematics. Wiley, 1969.<br />

[HL99] F. Hirsch and G. Lacombe. Elements of functional analysis, volume 192 of<br />

Graduate Texts <strong>in</strong> Mathematics. Spr<strong>in</strong>ger-Verlag, New York, 1999.


<strong>Potential</strong> <strong>Theory</strong> <strong>in</strong> <strong>Classical</strong> <strong>Probability</strong> 59<br />

[IW89] N. Ikeda and S. Watanabe. Stochastic Differential Equations and Diffusion<br />

Processes. North-Holland, 1989.<br />

[JP00] J. Jacod and Ph. Protter. <strong>Probability</strong> essentials. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 2000.<br />

[Kal02] O. Kallenberg. Foundations of modern probability. <strong>Probability</strong> and its Applications.<br />

Spr<strong>in</strong>ger-Verlag, New York, second edition, 2002.<br />

[MR92] Z.M. Ma and M. Röckner. Introduction to the theory of (nonsymmetric) Dirichlet<br />

forms. Universitext. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 1992.<br />

[Pro05] Ph. Protter. Stochastic <strong>in</strong>tegration and differential equations. Stochastic Modell<strong>in</strong>g<br />

and Applied <strong>Probability</strong>, Vol. 21. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 2005.<br />

[PS78] S.C. Port and C.J. Stone. Brownian motion and classical potential theory.<br />

Academic Press, 1978.<br />

[Rev75] D. Revuz. Markov cha<strong>in</strong>s. North Holland, 1975.


Introduction to Random Walks<br />

on Noncommutative Spaces<br />

Philippe Biane<br />

Abstract We <strong>in</strong>troduce several examples of random walks on noncommutative<br />

spaces and study some of their probabilistic properties. We emphasize<br />

connections between classical potential theory and group representations.<br />

1 Introduction<br />

Whereas random walks form one of the most <strong>in</strong>vestigated class of stochastic<br />

processes, their noncommutative analogues have been studied only recently.<br />

In these lectures I will present some results on random walks which take their<br />

values <strong>in</strong> noncommutative spaces. The notion of a noncommutative space has<br />

emerged progressively from the development of quantum physics, see e.g. [C].<br />

The key idea is to consider not the space itself but the set of real, or complex<br />

functions on it. For a usual space, this forms an algebra, which is commutative<br />

by nature. A noncommutative space is given by a noncommutative<br />

(usually complex) algebra which is to be thought of as the algebra of complex<br />

functions on the space. We shall expla<strong>in</strong> this idea <strong>in</strong> more details <strong>in</strong><br />

section 2, and <strong>in</strong> particular def<strong>in</strong>e noncommutative probability spaces. Once<br />

noncommutative spaces have been def<strong>in</strong>ed <strong>in</strong> this way it is easy to def<strong>in</strong>e random<br />

variables, and stochastic processes tak<strong>in</strong>g their values <strong>in</strong> these spaces.<br />

Rather than start<strong>in</strong>g an abstract theory, these lectures will consist ma<strong>in</strong>ly <strong>in</strong><br />

a collection of examples, which I th<strong>in</strong>k show that this notion is <strong>in</strong>terest<strong>in</strong>g<br />

and worth study<strong>in</strong>g. We shall beg<strong>in</strong> with the most simple stochastic process<br />

namely the Bernoulli random walk. We shall show how to quantize it <strong>in</strong><br />

order to construct the quantum Bernoulli random walk. Simple as it is this<br />

P. Biane<br />

Institut d’électronique et d’<strong>in</strong>formatique Gaspard-Monge, 77454 Marne-la-Vallée Cedex 2,<br />

France<br />

e-mail: biane@univ-mlv.fr<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 61<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


62 P. Biane<br />

noncommutative stochastic process exhibits quite deep properties, related to<br />

group representation and potential theory. Actually <strong>in</strong>terpret<strong>in</strong>g it as a random<br />

walk with values <strong>in</strong> a noncommutative space, the dual of SU(2), we will<br />

be lead to def<strong>in</strong>e random walks with values <strong>in</strong> duals of compact groups. The<br />

study of such random walks <strong>in</strong> the case of special unitary groups uncovers<br />

connections with potential theory, <strong>in</strong> particular with the Mart<strong>in</strong> boundary.<br />

We will <strong>in</strong>vestigate more on these connections. From the Bernoulli random<br />

walk we can take limit objects, as <strong>in</strong> the central limit theorem. One of these<br />

objects is a noncommutative Brownian motion which we construct as a family<br />

of operators on a Fock space and <strong>in</strong>terpret then as a cont<strong>in</strong>uous time<br />

stochastic process with <strong>in</strong>dependent <strong>in</strong>crements, with values <strong>in</strong> the dual of<br />

the Heisenberg group. We then extend this construction to more general<br />

noncompact locally compact groups. F<strong>in</strong>ally we will also start to consider<br />

quantum groups <strong>in</strong> the last section.<br />

The next section consists <strong>in</strong> prelim<strong>in</strong>aries about C ∗ and von Neumann<br />

algebras and noncommutative spaces.<br />

2 Noncommutative Spaces and Random Variables<br />

2.1 What are Noncommutative Spaces?<br />

The random walks that we are go<strong>in</strong>g to study take their values <strong>in</strong> noncommutative<br />

spaces, so we should start by mak<strong>in</strong>g this notion more precise. In<br />

many parts of mathematics, one studies spaces through the set of functions<br />

def<strong>in</strong>ed on them. There can be many k<strong>in</strong>d of functions, e.g. measurable, <strong>in</strong>tegrable,<br />

cont<strong>in</strong>uous, bounded, differentiable, and so on. Each property of the<br />

functions reflects a property of the space on which they are def<strong>in</strong>ed. Sometimes,<br />

<strong>in</strong> probability theory for example, one is even not <strong>in</strong>terested at all <strong>in</strong><br />

the space, but only <strong>in</strong> the functions themselves, the random variables. Also<br />

very often the set of complex functions considered determ<strong>in</strong>es completely the<br />

underly<strong>in</strong>g space. This is the case for example for compact topological spaces,<br />

determ<strong>in</strong>ed by their algebra of cont<strong>in</strong>uous functions, or differentiable manifolds<br />

which are determ<strong>in</strong>ed by their smooth functions. The common feature<br />

shared by these situations is that all these spaces of complex valued functions<br />

are commutative algebras. It has been realized, s<strong>in</strong>ce the beg<strong>in</strong>n<strong>in</strong>g of<br />

quantum mechanics that one can obta<strong>in</strong> a better description of nature by<br />

relax<strong>in</strong>g this commutativity hypothesis. Henceforth we shall consider a non<br />

commutative space as given by a complex algebra, which plays the role of<br />

space of functions on the space. We will see many examples throughout these<br />

lectures. Some algebras may come equipped with a supplementary structure,<br />

for example an antil<strong>in</strong>ear <strong>in</strong>volution, a norm, a preferred l<strong>in</strong>ear form,<br />

a topology, etc... Actually most of the times these algebras will be algebras


Introduction to Random Walks on Noncommutative Spaces 63<br />

of operators on some complex Hilbert space H, and the <strong>in</strong>volution will be<br />

given by the adjo<strong>in</strong>t operation. We shall decribe the k<strong>in</strong>d of algebras we will<br />

consider, ma<strong>in</strong>ly C ∗ and von Neumann algebras. We will use the language<br />

and some basic results <strong>in</strong> the theory of these objects, but we will need no<br />

deep knowledge of them. We will only assume that the reader is familiar with<br />

the spectral theorem for selfadjo<strong>in</strong>t operators on a Hilbert space. We refer<br />

for example to the treatises [D1], [D2] or [T] for more details.<br />

2.2 C ∗ Algebras<br />

A C ∗ algebra is a normed ∗-algebra which is isometric with a subalgebra of<br />

the algebra B(H) of all bounded operators on some complex Hilbert space H,<br />

stable under tak<strong>in</strong>g the adjo<strong>in</strong>t, and closed for the operator norm topology.<br />

Elements <strong>in</strong> a C ∗ -algebra of the form aa ∗ for some a ∈ A are called positive.<br />

Positive elements are exactly the selfadjo<strong>in</strong>t positive operators which belong<br />

to the algebra.<br />

Let X be a locally compact topological space, then the algebra of complex<br />

cont<strong>in</strong>uous functions on X, vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity, is a C ∗ -algebra, and the famous<br />

Gelfand-Naimark theorem states that any commutative C ∗ -algebra is<br />

isomorphic to such an algebra. The topological space is compact if and only if<br />

the algebra has a unit, and there is a one to one correspondence between the<br />

po<strong>in</strong>ts of the space and the characters of the algebra, that is, the cont<strong>in</strong>uous<br />

algebra homomorphisms with values <strong>in</strong> the complex numbers, or equivalently<br />

with the maximal closed ideals, therefore the space is unique up to homeomorphism<br />

and can be recovered from the algebra. It is usually denoted<br />

by spec(A) ifA is the C ∗ -algebra. Thus we should th<strong>in</strong>k of a C ∗ -algebra<br />

as provid<strong>in</strong>g the algebra of cont<strong>in</strong>uous functions on some noncommutative<br />

space. Note that C ∗ algebras are closed under cont<strong>in</strong>uous functional calculus,<br />

namely if a is a self-adjo<strong>in</strong>t element <strong>in</strong> a C ∗ algebra, and f a cont<strong>in</strong>uous<br />

functions on its spectrum, then the operator f(a) also belongs to the algebra<br />

C. This can be easily seen by approximat<strong>in</strong>g uniformly f by polynomials on<br />

the spectrum of a.<br />

If A ⊂ B(H) isaC ∗ -algebra, then the multiplier algebra M(A) ofA is the<br />

set of all operators x such that xA ⊂ A and Ax ⊂ A. ItisaC ∗ algebra with<br />

a unit, conta<strong>in</strong><strong>in</strong>g A. Itco<strong>in</strong>cideswithA if and only if A has a unit. If A is<br />

abelian, then M(A) is just the algebra of all bounded cont<strong>in</strong>uous functions on<br />

spec(A). In the noncommutative situation, it corresponds to the Stone-Cech<br />

compactification of the topological space underly<strong>in</strong>g the algebra.<br />

Cont<strong>in</strong>uous positive l<strong>in</strong>ear functionals on a C ∗ algebra play the role of<br />

positive bounded measures. Here positivity for a functional means that it<br />

is positive on positive elements. Aga<strong>in</strong> <strong>in</strong> the commutative case, by Riesz’<br />

theorem, such l<strong>in</strong>ear functionals correspond to f<strong>in</strong>ite positive Borel measures<br />

on the underly<strong>in</strong>g topological space. Positive l<strong>in</strong>ear functionals of norm one


64 P. Biane<br />

are called states, and correspond to probability measures <strong>in</strong> the commutative<br />

case. A large supply of states is given by unit vectors <strong>in</strong> the Hilbert space on<br />

which the C∗ algebra acts. Indeed any such vector ψ def<strong>in</strong>es a state by the<br />

formula<br />

ωψ(a) =〈aψ, ψ〉 a ∈ A<br />

Given a self adjo<strong>in</strong>t element a <strong>in</strong> a C∗ algebra A, and a state σ on A, there<br />

exists a unique measure on R, with compact support, such that<br />

�<br />

σ(f(a)) = f(x)dµ(x) for all cont<strong>in</strong>uous f on R.<br />

The support of this measure is <strong>in</strong>cluded <strong>in</strong> the spectrum of a. TheGNS<br />

construction assigns to every C ∗ algebra, with a cont<strong>in</strong>uous positive l<strong>in</strong>ear<br />

functional σ, a representation of the algebra on a Hilbert space. A l<strong>in</strong>ear<br />

functional is called tracial if for any a, b ∈ A one has τ(ab) =τ(ba).<br />

Each cont<strong>in</strong>uous map between topological spaces f : X → Y gives rise<br />

to a cont<strong>in</strong>uous algebra morphism Φf : C0(Y ) → C0(X); h ↦→ h ◦ f, and<br />

conversely any such algebra morphism comes from a cont<strong>in</strong>uous map, therefore<br />

one can th<strong>in</strong>k of a homomorphism between C ∗ algebras as a cont<strong>in</strong>uous<br />

map between the underly<strong>in</strong>g noncommutative spaces (with the direction of<br />

the arrows reversed). One must note however that there may exist very few<br />

morphisms between two C ∗ algebras. For example there does not exist any<br />

nonzero homomorphism from the f<strong>in</strong>ite dimensional C ∗ algebra Mn(C) to<br />

Mm(C) ifn > m. Indeed this is a purely algebraic fact, s<strong>in</strong>ce Mn(C) is<br />

a simple algebra, if a homomorphism from Mn(C) is not <strong>in</strong>jective, then it<br />

must be 0.<br />

2.3 von Neumann Algebras<br />

Let S be a subset of B(H), then its commutant S ′ is the set of bounded<br />

operators which commute with every element of S. A von Neumann algebra<br />

is a subalgebra of B(H) which is closed under under tak<strong>in</strong>g the adjo<strong>in</strong>t,<br />

and is equal to its bicommutant, i.e. the commutant of its commutant. By<br />

the von Neumann bicommutant theorem the von Neumann algebras are the<br />

∗-subalgebra of B(H), conta<strong>in</strong><strong>in</strong>g the identity operator, and closed for the<br />

strong topology. S<strong>in</strong>ce the strong topology is weaker than the operator norm<br />

topology any von Neumann algebra is also a unital C ∗ algebra, although<br />

generally too large to be <strong>in</strong>terest<strong>in</strong>g as such.<br />

A von Neumann algebra is closed under Borel functional calculus, namely<br />

if a ∈ M is self-adjo<strong>in</strong>t and f is a bounded Borel function on the spectrum of<br />

a, thenf(a) belongs to M, and aga<strong>in</strong>, the same is true for f(a1,...,an)where<br />

a1,...,an are commut<strong>in</strong>g selfadjo<strong>in</strong>t operators <strong>in</strong> the von Neumann algebra,<br />

and f is a bounded Borel function def<strong>in</strong>ed on the product of their spectra.


Introduction to Random Walks on Noncommutative Spaces 65<br />

The nuance between C ∗ algebras and von Neumann algebras can be grasped<br />

by look<strong>in</strong>g at the commutative case. Indeed commutative von Neumann algebras<br />

correspond to measure spaces, more precisely any commutative von<br />

Neumann algebra is isomorphic to the algebra L ∞ (X, m) whereX is a measure<br />

space and m a positive measure (the algebra actually depends only on<br />

the class of the measure). This statement can be seen as a reformulation<br />

of the spectral theorem for commut<strong>in</strong>g self-adjo<strong>in</strong>t operators on a Hilbert<br />

space. Therefore it is natural to th<strong>in</strong>k of von Neumann algebras as “algebras<br />

of noncommutative random variables”. A normal state on a von Neumann<br />

algebra is a positive l<strong>in</strong>ear form which is cont<strong>in</strong>uous for the σ-weak topology<br />

and takes the value 1 on the unit. It corresponds, <strong>in</strong> the commutative case,<br />

to a probability measure, which is absolutely cont<strong>in</strong>uous with respect to the<br />

measure m. We shall sometimes call a von Neumann algebra, with a normal<br />

state, a “non commutative probability space”.<br />

A weight on a von Neumann algebra is a map ϕ from the cone of positive<br />

elements of the von Neumann algebra to [0, +∞], which is additive, and<br />

homogeneous, i.e. ϕ(λx) =λϕ(x) forx positive and real λ>0. A weight is<br />

called normal if sup i∈I ϕ(xi) =ϕ(sup(xi)i∈I) for every bounded <strong>in</strong>creas<strong>in</strong>g<br />

net (xi)i∈I. Com<strong>in</strong>g back to the commutative case, weights are positive, possibly<br />

unbounded measures, <strong>in</strong> the measure class of m. Aweightµ is called<br />

f<strong>in</strong>ite if µ(1) < ∞, <strong>in</strong>thiscaseµ is a multiple of a state.<br />

Given a selfadjo<strong>in</strong>t element, a ∈ M and a normal state σ on M, wedenote<br />

by µa the distribution of a, namely the measure such that<br />

�<br />

σ(f(a)) = f(x)dµa(x)<br />

for all bounded Borel functions on spec(a). More generally if a1,...,an is a<br />

family of commut<strong>in</strong>g self-adjo<strong>in</strong>t operators <strong>in</strong> M, their jo<strong>in</strong>t distribution is<br />

the unique probability measure µa1,...,an on Rn such that<br />

�<br />

σ(f(a1,...,an)) = f(x)dµa1,...,an(x)<br />

for all bounded Borel function f on R n .<br />

Let N ⊂ M be a von Neumann subalgebra, and σ a state on M, thena<br />

conditional expectation of M onto N is a norm one projection σ(.|N) such<br />

that σ(a|N) =a for all a ∈ N, σ(abc|N) =aσ(b|N)c for all a, c ∈ N,b ∈ M,<br />

and σ(σ(b|N)) = σ(b) for all b ∈ M. GivenM,N and σ, such a map need<br />

not exist, but it always exists, and is unique, if σ is tracial.<br />

We will consider spatial tensor products of von Neumann algebras. If<br />

A ⊂ B(H) andB ⊂ B(K) are two von Neumann algebras, their algebraic<br />

tensor product acts on the Hilbert space H ⊗ K, and the spatial tensor product<br />

of A and B is def<strong>in</strong>ed as the von Neumann algebra generated by this<br />

tensor product. Given an <strong>in</strong>f<strong>in</strong>ite family of von Neumann algebras (Ai; i ∈ I)<br />

equipped with normal states ωi it is possible to construct an <strong>in</strong>f<strong>in</strong>ite tensor


66 P. Biane<br />

product ⊗i(Ai,ωi) which is a von Neumann algebra with a state ⊗iωi. One<br />

considers operators of the form ⊗i∈Iai where ai ∈ Ai and ai �= Id only for<br />

a f<strong>in</strong>ite number of i ∈ I. These generate an algebra, which is the algebraic<br />

tensor product of the Ai. One can def<strong>in</strong>e a positive l<strong>in</strong>ear functional on this<br />

algebraic tensor product by ω(⊗i∈Iai) = � ωi(ai). The GNS construction<br />

then yields a Hilbert space H, with a pure state on B(H), and the von Neumann<br />

algebra tensor product is the von Neumann algebra <strong>in</strong> B(H) generated<br />

by this algebraic tensor product.<br />

2.4 Random Variables, Stochastic Processes<br />

with Values <strong>in</strong> some Noncommutative Space<br />

Given a von Neumann algebra M equipped with a normal state σ, andaC ∗<br />

algebra C, a random variable with values <strong>in</strong> C (or, more appropriately, <strong>in</strong> the<br />

noncommutative space underly<strong>in</strong>g C), is a norm cont<strong>in</strong>uous morphism from<br />

C to M. The distribution of the random variable ϕ : C → M is the state<br />

on C given by σ ◦ ϕ. If the algebra C is commutative, then it corresponds to<br />

some topological space, and the state σ ◦ ϕ to a probability measure on this<br />

space. When C = C0(R) there exists a self-adjo<strong>in</strong>t element a ∈ M such that<br />

ϕ(f) =f(a) for all f ∈ C, and we are back to the situation <strong>in</strong> the preced<strong>in</strong>g<br />

section, where the distribution of a was def<strong>in</strong>ed. We will call the state σ ◦ ϕ<br />

the distribution of the random variable ϕ. More generally a family of random<br />

variables with values <strong>in</strong> some noncommutative space, <strong>in</strong>dexed by some set, is<br />

a stochastic process.<br />

If A and B are two C ∗ -algebras, a positive map Φ : A → B is a l<strong>in</strong>ear<br />

map such that Φ(a) is positive for each positive a ∈ A. WhenA and B are<br />

commutative, thus A = C0(X) andB = C0(Y ), such a l<strong>in</strong>ear map can be<br />

realized as a measure kernel k(y, dx) whereforeachy ∈ Y one has a f<strong>in</strong>ite<br />

positive measure k(y, dx) onX. IfA and B are unital, and Φ(I) =I then this<br />

kernel is a Markov kernel, i.e. all measures are probability measures. Thus<br />

we see that the generalization of a Markov kernel to the non-commutative<br />

context can be given by the notion of positive maps. It turns out however<br />

that this notion is slightly too general to be useful and it is necessary to<br />

restrict oneself to a particular class called completely positive maps.<br />

Def<strong>in</strong>ition 2.1. l<strong>in</strong>ear map between two C ∗ algebras A and B is called completely<br />

positive if, for all n ≥ 0, the map Φ ⊗ Id : A ⊗ Mn(C) → B ⊗ Mn(C)<br />

is positive. It is called unit preserv<strong>in</strong>g if furthermore Φ(Id)=Id.<br />

We shall consider semigroups of unit preserv<strong>in</strong>g, completely positive maps on<br />

a C ∗ algebra C. These will be <strong>in</strong>dexed by a set of times which will be either<br />

the nonnegative <strong>in</strong>tegers (discrete times) or the positive real l<strong>in</strong>e (cont<strong>in</strong>uous<br />

time). Thus a discrete time semigroup of unit preserv<strong>in</strong>g, completely positive<br />

maps on a C ∗ algebra C will be a family (Φn : C → C)n≥0 of completely


Introduction to Random Walks on Noncommutative Spaces 67<br />

positive maps, such that Φn ◦ Φm = Φn+m. In cont<strong>in</strong>uous time we will have<br />

a family (Φt)t∈R+ which satisfies Φt ◦ Φs = Φt+s. In the discrete time sett<strong>in</strong>g<br />

one has Φn =(Φ1) n and the semigroup is deduced from the value at time 1.<br />

We shall denote generally the time set by T when we do not specify whether<br />

we are <strong>in</strong> discrete or cont<strong>in</strong>uous time.<br />

Def<strong>in</strong>ition 2.2. Let C be a C ∗ algebra, then a dilation of a semigroup<br />

(Φt)t∈T of completely positive maps on C is given by a von Neumann algebra<br />

M, with a normal state ω, an <strong>in</strong>creas<strong>in</strong>g family of von Neumann subalgebras<br />

Mt; t ∈ T , with conditional expectations ω(.|Mt), and a family of morphisms<br />

jt : C → (M,ω) such that for any t ∈ T and a ∈ C, one has jt(a) ∈ Mt and<br />

for all s


68 P. Biane<br />

where t1,...,tn is an arbitrary sequence of times (i.e. not necessarily <strong>in</strong>creas<strong>in</strong>g)<br />

and a1,...,an ∈ C. ThusifC is a commutative algebra, then all<br />

commutative dilations with the same <strong>in</strong>itial distribution are equivalent, but<br />

for a given semigroup of completely positive maps there may exist a lot of<br />

non equivalent dilations. Actually we shall encounter <strong>in</strong> these lectures some<br />

natural noncommutative dilations of Markov semigroups on classical spaces!<br />

An important source of such dilations comes from restrictions: if B ⊂ C is<br />

a subalgebra and the completely positive semigroup leaves B <strong>in</strong>variant, then<br />

the restriction of (jt)t∈T to the subalgebra B is a dilation of the restriction<br />

of the completely positive semigroup.<br />

3 Quantum Bernoulli Random Walks<br />

3.1 Quantization of the Bernoulli Random Walk<br />

Our first example of a quantum random walk will be the quantization of<br />

the simple (or Bernoulli) random walk. This is just the random walk whose<br />

<strong>in</strong>dependent <strong>in</strong>crements have values ±1. In order to quantize it we will replace<br />

the set of <strong>in</strong>crements {±1} by its quantum analogue, namely the space of<br />

two by two complex matrices, with its structure of C ∗ -algebra. The subset<br />

of hermitian operators is a four dimensional real subspace, generated by the<br />

identity matrix I as well as the three matrices<br />

� �<br />

01<br />

σx =<br />

10<br />

� �<br />

0 −i<br />

σy =<br />

i 0<br />

σz =<br />

� �<br />

1 0<br />

0 −1<br />

The matrices σx,σy,σz are the Pauli matrices. They satisfy the commutation<br />

relations<br />

[σx,σy] =2iσz; [σy,σz] =2iσx; [σz,σx] =2iσy. (3.1)<br />

The group SU(2) acts by the automorphisms A → UAU ∗ on this C ∗ -algebra.<br />

We observe that this group is much larger than the group of symmetries of the<br />

two po<strong>in</strong>ts space (which consists just of a two elements group). This action<br />

leaves the space generated by I <strong>in</strong>variant, and acts by rotations on the real<br />

three dimensional space generated by the Pauli matrices. Indeed the <strong>in</strong>ner<br />

product on the space of hermitian matrices 〈A, B〉 = Tr(AB) is <strong>in</strong>variant by<br />

unitary conjugation.<br />

A state ω on M2(C) is given by a positive hermitian matrix S with trace<br />

1, by the formula<br />

ω(A) =Tr(AS).<br />

The most general such matrix can be written as<br />

S = 1<br />

� �<br />

1+u v + iw<br />

2 v − iw 1 − u


Introduction to Random Walks on Noncommutative Spaces 69<br />

where (u, v, w) ∈ R 3 satisfies u 2 + v 2 + w 2 ≤ 1. The extreme po<strong>in</strong>ts on the<br />

unit sphere (sometimes called the “Bloch sphere” <strong>in</strong> the physics litterature),<br />

correspond to pure states, when S is a projection on a one dimensional subspace.<br />

Any hermitian operator has a two-po<strong>in</strong>t spectrum, hence <strong>in</strong> a state<br />

ω its distribution is a probability measure on R supported by at most two<br />

po<strong>in</strong>ts. In particular, for each of the Pauli matrices, its distribution <strong>in</strong> the<br />

state ω is a probability measure on {±1}, givenby<br />

P (σx =1)= 1+v<br />

2<br />

P (σy =1)= 1+w<br />

2<br />

P (σz =1)= 1+u<br />

2<br />

(3.2)<br />

Mimick<strong>in</strong>g the construction of a random walk, we use the <strong>in</strong>f<strong>in</strong>ite product<br />

algebra (M2(C),ω) ⊗N (recall the construction of section 2.3). For each Pauli<br />

matrix we build the matrices<br />

xn = I ⊗(n−1) ⊗σx⊗I ⊗∞ yn = I ⊗(n−1) ⊗σy⊗I ⊗∞ , zn = I ⊗(n−1) ⊗σz⊗I ⊗∞<br />

which represent the <strong>in</strong>crements of the process. It is easy to see that, for<br />

example, the operators xn, forn≥1, form a commut<strong>in</strong>g family of operators,<br />

which is distributed, <strong>in</strong> the state ω∞ , as a sequence of <strong>in</strong>dependent Bernoulli<br />

random variables.<br />

Then we put<br />

n�<br />

n�<br />

n�<br />

Xn = xi; Yn = yi Zn = zi.<br />

i=1<br />

i=1<br />

This gives us three families of operators (Xn)n≥1;(Yn)n≥1 and (Zn)n≥1 on<br />

this space.<br />

We observe that each of these three families consists <strong>in</strong> commut<strong>in</strong>g operators,<br />

hence has a jo<strong>in</strong>t distribution. It is not difficult to check that this<br />

distribution is that of a Bernoulli random walk, whose <strong>in</strong>crements have distribution<br />

given by the probability distributions on {±1} of formula (3.2).<br />

The three families of operators, however do not commute. In fact us<strong>in</strong>g the<br />

commutation relations (3.1) one sees that for n, m positive <strong>in</strong>tegers,<br />

[xn,ym] =2izmδnm<br />

and<br />

[Xn,Ym] =2iZn∧m<br />

(3.3)<br />

as well as the similar relations obta<strong>in</strong>ed by cyclic permutation of X, Y, Z.<br />

We will call the family of triples of operators (Xn,Yn,Zn); n ≥ 1 a quantum<br />

Bernoulli random walk. We shall later <strong>in</strong>terpret this noncommutative process<br />

as a random walk with values <strong>in</strong> some noncommutative space, however for the<br />

moment we will study some of its properties related to the automorphisms<br />

of the algebra M2(C).<br />

i=1


70 P. Biane<br />

3.2 The Sp<strong>in</strong> Process<br />

Because of the rotation <strong>in</strong>variance of the commutation relations (3.1), we see<br />

that for any unitary matrix U we can def<strong>in</strong>e conjugated variables<br />

and<br />

X U n =<br />

x U n = I⊗(n−1) ⊗ UσxU ∗ ⊗ I ⊗∞<br />

y U n = I ⊗(n−1) ⊗ UσyU ∗ ⊗ I ⊗∞<br />

z U n = I⊗(n−1) ⊗ UσzU ∗ ⊗ I ⊗∞<br />

n�<br />

i=1<br />

x U i ; Y U n =<br />

n�<br />

i=1<br />

y U i<br />

Z U n =<br />

then this new stochastic process is obta<strong>in</strong>ed from the orig<strong>in</strong>al quantum<br />

Bernoulli random walk by a rotation matrix. It follows by a simple computation,<br />

us<strong>in</strong>g the commutation relations, that X2 n + Y 2 n + Z2 n is <strong>in</strong>variant<br />

under conjugation, namely one has<br />

for any unitary matrix U.<br />

n�<br />

i=1<br />

X 2 n + Y 2 n + Z 2 n =(X U n ) 2 +(Y U n ) 2 +(Z U n ) 2<br />

Lemma 3.1. For all m, n ≥ 1 one has<br />

[X 2 n + Y 2 n + Z2 n ,X2 m + Y 2 m + Z2 m ]=0<br />

Actually we shall prove that [Xn,X2 m + Y 2 m + Z2 m ]=0ifm ≤ n. This follows<br />

from the computation<br />

[Xn,Y 2 m ]=[Xn,Ym]Ym + Ym[Xn,Ym] =2i(ZmYm + YmZm)<br />

[Xn,Z 2 m ]=[Xn,Zm]Zm + Zm[Xn,Ym] =−2i(ZmYm + YmZm)<br />

Us<strong>in</strong>g <strong>in</strong>variance of the commutation relations by cyclic permutation of<br />

X, Y, Z we also have [Yn,X2 m + Y 2 m + Z2 m ]=[Zn,X2 m + Y 2 m + Z2 m ] = 0, and<br />

the result follows. ⊓⊔<br />

We deduce from this that the family of operators (X 2 n + Y 2 n + Z2 n ); n ≥ 1is<br />

commutative, and therefore def<strong>in</strong>es a classical process. We shall compute its distribution<br />

now. For this we <strong>in</strong>troduce the operators Sn = � X2 n + Y 2 n + Z2 n + I.<br />

By the preced<strong>in</strong>g Lemma, the family (Sn)n≥0 is a commut<strong>in</strong>g family of operators,<br />

therefore one can consider their jo<strong>in</strong>t distribution.<br />

Theorem 3.2. Let us take for ω the tracial state, then the operators (Sn;<br />

n ≥ 1) form a Markov cha<strong>in</strong>, with values <strong>in</strong> the positive <strong>in</strong>tegers, such that<br />

P (S1 =2)=1, and transition probabilities<br />

p(k, k +1)=<br />

z U i<br />

k +1<br />

k − 1<br />

; p(k, k − 1) =<br />

2k 2k .


Introduction to Random Walks on Noncommutative Spaces 71<br />

In order to prove the theorem, it is enough to consider the process up to a<br />

f<strong>in</strong>ite time n. So we shall restrict considerations to a f<strong>in</strong>ite product M2(C) ⊗n<br />

act<strong>in</strong>g on (C 2 ) ⊗n . We remark that the state ω ⊗n on (C 2 ) ⊗n is the unique<br />

tracial state, and it gives, to every orthogonal projection on a subspace V ,the<br />

dim(V )<br />

value 2n . We shall f<strong>in</strong>d a basis of this space consist<strong>in</strong>g of jo<strong>in</strong>t eigenvectors<br />

of the operators (Sk;1 ≤ k ≤ n). For this we analyze the action of the<br />

operators A + n<br />

where<br />

A + n =<br />

n�<br />

j=1<br />

:= 1<br />

2 (Xn − iYn) andA − n<br />

I ⊗(j−1) ⊗ α + ⊗ I ⊗n−j<br />

α + =<br />

� �<br />

00<br />

10<br />

= 1<br />

2 (Xn + iYn). One has<br />

A − n =<br />

α − =<br />

n�<br />

j=1<br />

� �<br />

01<br />

.<br />

00<br />

I ⊗(j−1) ⊗ α − ⊗ I ⊗n−j<br />

Let us call e0,e1 the canonical basis of C2 . An orthonormal basis of (C2 ) ⊗n<br />

is given by the vectors eU; U ⊂{1, 2,...,n} where eU = ei1 ⊗...⊗e<strong>in</strong>, ik =1<br />

if k ∈ U and ik =0ifk/∈ U. In terms of this basis the action of A + n and A−n is given by<br />

A + n eU = �<br />

k/∈U eU∪{k} A− n eU = �<br />

k∈U eU\{k} ZneU =(n− 2|U|)eU<br />

Let us consider the vector e ∅ = e ⊗n<br />

0<br />

and its images by the powers of A − n ,<br />

normalized to have norm one. These are the vectors<br />

ε j n =<br />

�<br />

j!(n − j)!<br />

n!<br />

�<br />

eU j =0, 1,...,n<br />

U⊂{1,2,...,n};|U|=j<br />

and these vectors are orthogonal. The action of the operators A + n ,A − n ,Zn on<br />

these vectors is given by<br />

A + n εjn = � (j +1)(n−j)εj+1 n<br />

A− n εj n = � j(n − j +1)εj−1 n<br />

Znεj n =(n− 2j)εj n<br />

Snεj n =(n +1)εjn In particular we see that these vectors belong to the eigenspace of Sn of<br />

eigenvalue n + 1. We shall generalize this computation to f<strong>in</strong>d the common<br />

eigenspaces of the operators S1,S2,...,Sn.<br />

Lemma 3.3. Let J =(j1,...,jn) be a sequence of <strong>in</strong>tegers such that<br />

• i) j1 =2<br />

• ii) ji ≥ 1 for all i ≤ n<br />

• iii) |ji+1 − ji| =1for all i ≤ n − 1


72 P. Biane<br />

then there exists a subspace HJ of (C 2 ) ⊗n of dimension jn = l +1,which<br />

is an eigenspace of S1,...,Sn, with respective eigenvalues j1,...,jn, andan<br />

orthonormal basis (φ0,...,φl) such that<br />

A + n φj = � (j +1)(l − j)φj+1<br />

A − n φj = � j(l − j +1)φj−1<br />

Znφj =(l − 2j)φj<br />

Furthermore, the spaces HJ are orthogonal and (C 2 ) ⊗n = ⊕JHJ.<br />

Proof of the lemma. We shall use <strong>in</strong>duction on n. The lemma is true<br />

for n = 1, us<strong>in</strong>g φ0 = e0; φ1 = e1. Assume the lemma holds for n, and<br />

let J = (j1,...,jn) be a sequence satisfy<strong>in</strong>g the conditions i),ii), iii), of<br />

the lemma. We shall decompose the space HJ ⊗ C 2 as a direct sum of two<br />

subspaces. Let the vectors ψj and ηj be def<strong>in</strong>ed by<br />

ψj =<br />

�<br />

l − j +1<br />

l +1 φj<br />

�<br />

j<br />

⊗ e0 +<br />

l +1 φj−1 ⊗ e1 j =0,...,l+1<br />

and<br />

�<br />

j +1<br />

ηj =<br />

l +1 φj+1<br />

�<br />

l − j<br />

⊗ e0 −<br />

l +1 φj ⊗ e1 j =0,...,l− 1<br />

It is easy to check that these vectors form an orthonormal basis of the tensor<br />

product HJ ⊗ C2 , and a simple computation us<strong>in</strong>g<br />

Xn+1 = Xn + I ⊗n ⊗ σx,Yn+1 = Yn + I ⊗n ⊗ σy,Zn+1 = Zn + I ⊗n ⊗ σz<br />

shows that these vectors have the right behaviour under these operators. ⊓⊔<br />

We can now prove theorem 3.2, <strong>in</strong>deed for any sequence satisfy<strong>in</strong>g the<br />

hypotheses of lemma 3.3 one has<br />

P (S1 = j1,...,Sn = jn) = jn j1<br />

=<br />

2n 2<br />

j2<br />

...<br />

2j1<br />

jn<br />

2jn−1<br />

where the right hand side is given by the distribution of the Markov cha<strong>in</strong> of<br />

theorem 3.2. ⊓⊔<br />

We shall give another, more conceptual, proof of the former result <strong>in</strong><br />

section 5, us<strong>in</strong>g the representation theory of the group SU(2) and Clebsch-<br />

Gordan formulas. For this we shall <strong>in</strong>terpret the quantum Bernoulli random<br />

walk as a Markov process with values <strong>in</strong> a noncommutative space, but we<br />

need first to give some def<strong>in</strong>itions perta<strong>in</strong><strong>in</strong>g to bialgebras and group algebras,<br />

which we do <strong>in</strong> the next section.


Introduction to Random Walks on Noncommutative Spaces 73<br />

4 Bialgebras and Group Algebras<br />

4.1 Coproducts<br />

Let X be a f<strong>in</strong>ite set, and F(X) be the algebra of complex functions on X.<br />

A composition law on X is a map X × X → X. This gives rise to a unit<br />

preserv<strong>in</strong>g algebra morphism ∆ : F(X) →F(X × X), where F(X × X)<br />

is the algebra of functions on X × X, and one has a natural isomorphism<br />

F(X × X) ∼ F(X) ⊗F(X). Conversely, such an algebra morphism ∆ :<br />

F(X) →F(X) ⊗F(X) comes from a composition law, and many properties<br />

of the composition law can be read on it. For example, associativity translates<br />

<strong>in</strong>to coassociativity of the coproduct which means that<br />

(∆ ⊗ I) ◦ ∆ =(I ⊗ ∆) ◦ ∆<br />

whereas commutativity gives cocommutativity for the coproduct, which<br />

means that<br />

v ◦ ∆ = ∆<br />

where v : F(X) ⊗F(X) →F(X) ⊗F(X) is the flip automorphism v(a ⊗<br />

b) = b ⊗ a. In order to obta<strong>in</strong> an analogue of a composition law <strong>in</strong> the<br />

noncommutative context, one can def<strong>in</strong>e a coproduct for any algebra A as a<br />

morphism ∆ : A → A ⊗ A, however <strong>in</strong> general the algebraic tensor product is<br />

too small for obta<strong>in</strong><strong>in</strong>g <strong>in</strong>terest<strong>in</strong>g examples. Th<strong>in</strong>k for example to the case<br />

A = Cb(X), X a locally compact space, and see that A ⊗ A ⊂ Cb(X × X) is<br />

a small subspace. We shall therefore consider coproducts which take values<br />

<strong>in</strong> a suitable completion of the algebraic tensor product. An algebra endowed<br />

with a coassociative coproduct is called a bialgebra. If the algebra is a C∗ (resp. a von Neumann) algebra and the tensor product is the m<strong>in</strong>imal C∗ product (resp. the von Neumann algebra tensor product), then one has a<br />

C∗ (resp. a von Neumann) bialgebra.<br />

Some further properties of a coproduct are the existence of the dual notion<br />

of the unit element and the <strong>in</strong>verse, which are respectively called a counit,<br />

ε : A → C and an antipode, i : ∆ → ∆.<br />

A Hopf algebra is a bialgebra with a unit and an antipode, satisfy<strong>in</strong>g some<br />

compatibility conditions. I refer for example to [K] for an exposition of Hopf<br />

algebras and quantum groups.<br />

4.2 Some Algebras Associated to a Compact Group<br />

We shall <strong>in</strong>vestigate <strong>in</strong> more details the notions above <strong>in</strong> the case of the group<br />

algebra of a compact group, which we assume for simplicity to be separable.<br />

Recall that every representation of a compact group can be made unitary, and<br />

can be reduced to an orthogonal direct sum of irreducible representations. The


74 P. Biane<br />

right regular representation of a compact group G is the representation of G<br />

on L 2 (G, m)(wherem is a Haar measure on G) by right translations ρgf(h) =<br />

f(hg −1 ) and the left regular representation acts by left translations λgf(h) =<br />

f(gh). A fundamental theorem is the Peter-Weyl theorem. It states that every<br />

irreducible representation of G arises <strong>in</strong> the decomposition of the left (or<br />

right) regular representation, actually one has an orthogonal direct sum<br />

L 2 (G) =⊕ χ∈ ˆ G Eχ<br />

where ˆ G is the (countable) set of equivalence classes of irreducible representations<br />

of G, andforχ ∈ ˆ G, Eχ is the space of coefficients of the representation<br />

i.e. the vector space generated by functions on G of the form<br />

f(g) =〈χ(g)u, v〉<br />

where u, v are vectors <strong>in</strong> the representation space of χ (〈., .〉 be<strong>in</strong>g an <strong>in</strong>variant<br />

hermitian product on the space). The space Eχ is f<strong>in</strong>ite dimensional, its<br />

dimension be<strong>in</strong>g dim(χ) 2 , and it is an algebra for the convolution product<br />

on G, isomorphic to the matrix algebra Mn(C) withn = dim(χ). We shall<br />

denote this space Mχ when we want to emphasize its algebra structure.<br />

We shall describe several algebras associated to G. The first one is the<br />

convolution algebra A 0 (G) generated by the coefficients of the f<strong>in</strong>ite dimensional<br />

representations. As a vector space it is the algebraic direct sum<br />

A 0 (G) =⊕χMχ ∼⊕χM dim(χ)(C). There are larger algebras such as L 1 (G)<br />

the space of <strong>in</strong>tegrable functions (with respect to the Haar measure on G),<br />

and C ∗ (G) theC ∗ algebra generated by L 1 (G). This algebra consists of sequences<br />

(mχ; χ ∈ ˆ G) such that mχ ∈ Mχ; |mχ| →0asχ →∞.<br />

The multiplier algebra of C ∗ (G) co<strong>in</strong>cides with the von Neumann algebra<br />

A(G), which is generated (topologically) by the left translation operators<br />

λg; g ∈ G, it consists of sequences (mχ; χ ∈ ˆ G) such that sup χ |mχ| < ∞. In<br />

both cases, the norm <strong>in</strong> these algebras is sup χ |mχ|. Note that the left and<br />

right translation operators λg and ρg are unitary, and the right translation<br />

operators generate the commutant of A(G).<br />

When the group G is abelian, there is a natural isomorphism, given by<br />

Fourier transform, between the algebra A(G) andL ∞ ( ˆ G) where here ˆ G is<br />

the group of characters of G (this is consistent with our earlier notation<br />

s<strong>in</strong>ce then all irreducible representations of G are one dimensional and are<br />

thus characters). This group of characters is a discrete abelian group, and its<br />

Haar measure is the count<strong>in</strong>g measure.<br />

In the general case, the algebra L ∞ ( ˆG) is isomorphic with the center of<br />

A(G), s<strong>in</strong>ce a bounded function on ˆ G can be identified with a sequence<br />

(mχ) χ∈ ˆ G of scalar operators.<br />

Any closed subgroup of G generates a von Neumann subalgebra A(H) furthermore<br />

the coproduct ∆ restricts to this subalgebra and def<strong>in</strong>es a coproduct<br />

∆ : A(H) → A(H) ⊗ A(H). If the subgroup is abelian, then this subalgebra<br />

is commutative and is isomorphic to the algebra L ∞ ( ˆ H),


Introduction to Random Walks on Noncommutative Spaces 75<br />

F<strong>in</strong>ally we shall also use the algebra Â(G) =�χ<br />

Mχ which consists <strong>in</strong><br />

unbounded operators on L2 (G), with common dense doma<strong>in</strong> ⊕χEχ (algebraic<br />

direct sum), affiliated with the von Neumann algebra A(G). One has natural<br />

<strong>in</strong>clusions<br />

A 0 (G) ⊂ C ∗ (G) ⊂ M(C ∗ (G)) = A(G) ⊂ Â(G)<br />

The algebra Â(G) ⊗ Â(G) is an algebra of operators on the algebraic direct<br />

sum ⊕χ,χ ′Eχ ⊗ Eχ ′ and we denote by Â(G) ˆ⊗ Â(G) its completion for simple<br />

′. One has<br />

convergence on ⊕χ,χ ′Eχ ⊗ Eχ<br />

Â(G) ⊗ Â(G) =� Mχ ⊗ �<br />

Mχ ⊂ Â(G) �<br />

ˆ⊗ Â(G) ∼ Mχ ⊗ Mχ ′<br />

χ<br />

χ<br />

The ∗ algebra structure extends obviously to Â(G),andanelementispositive<br />

if and only if its components are positive.<br />

4.3 The Coproduct<br />

The coproduct formula ∆ : λg → λg ⊗ λg extends by l<strong>in</strong>earity and cont<strong>in</strong>uity<br />

to the von Neumann algebra A(G) if we use the von Neumann algebra tensor<br />

product. It def<strong>in</strong>es a structure of cocommutative von Neumann bialgebra on<br />

A(G). One can also def<strong>in</strong>e an extension of the coproduct<br />

ˆ∆ : Â(G) → Â(G) ˆ⊗ Â(G).<br />

Indeed it is easy to check that for any χ ∈ ˆ G,anda ∈ Mχ, the operator ∆(a)<br />

is nonzero on the space Eχ ′ ⊗ Eχ ′′ if and only if χ has a non zero multiplicity<br />

<strong>in</strong> the decomposition of the tensor product representation Eχ ′ ⊗ Eχ ′′. It<br />

follows that for any sequence (aχ) χ∈Gˆ ∈ �<br />

χ∈ ˆ G Mχ the sum �<br />

χ∈ ˆ G ∆(aχ) is<br />

a f<strong>in</strong>ite sum <strong>in</strong> each component Eχ ⊗ Eχ ′ therefore it def<strong>in</strong>es an element <strong>in</strong><br />

�<br />

Â(G) ˆ⊗ Â(G) ∼ χ,χ ′ Mχ ⊗ Mχ ′.<br />

One can def<strong>in</strong>e the convolution of two f<strong>in</strong>ite weights µ and ν by the formula<br />

µ ∗ ν =(µ ⊗ ν) ◦ ∆<br />

s<strong>in</strong>ce the coproduct is cocommutative and coassociative, one checks that this<br />

is an associative and commutative operation.<br />

4.4 The Case of Lie Groups<br />

For compact Lie groups there is another algebra of <strong>in</strong>terest, which is the<br />

envelopp<strong>in</strong>g algebra of the Lie algebra. Recall that the Lie algebra of a Lie<br />

χ,χ ′


76 P. Biane<br />

group is composed of right <strong>in</strong>variant vector fields on the group. As such it<br />

acts on the L2 space as a family of unbounded operators. Therefore the Lie<br />

algebra is a subspace of the “big” algebra Â(G), and the algebra generated<br />

by this subspace is naturally isomorphic to the envelopp<strong>in</strong>g Lie algebra (for<br />

simply connected groups). The elements of the Lie algebra have the follow<strong>in</strong>g<br />

behaviour with respect to the extended coproduct on Â(G)<br />

∆(X) =X ⊗ I + I ⊗ X<br />

This can be seen by tak<strong>in</strong>g derivatives with respect to s <strong>in</strong> the equation<br />

∆(e sX )=e sX ⊗ e sX<br />

where e sX ; s ∈ R is the one parameter subgroup generated by X.<br />

4.5 States and Weights<br />

A normal state ν on A(G) is determ<strong>in</strong>ed by its value on the generators λg,<br />

thus by the function φν(g) =ν(λg). It is a classical result that a function φ<br />

on G corresponds to a state on C ∗ (G), and to a normal state on A(G), if and<br />

only if it is a cont<strong>in</strong>uous positive def<strong>in</strong>ite function on the group, satisfy<strong>in</strong>g<br />

φ(e) =1.<br />

Every normal weight on A(G) is given by a sequence of weights (νχ) χ∈ ˆ G<br />

on each of the subalgebras Mχ, therefore for a weight ν on A(G) thereexists<br />

a sequence of positive elements fχ ∈ Mχ such that νχ(a) = Tr(afχ) for<br />

all a ∈ Mχ. Conversely any such sequence fχ def<strong>in</strong>es a normal weight, thus<br />

normal weights on A(G) correspond to positive elements <strong>in</strong> Â(G).<br />

5 Random Walk on the Dual of SU(2)<br />

5.1 The Dual of SU(2) as a Noncommutative Space<br />

We shall now <strong>in</strong>terpret the process constructed <strong>in</strong> section 3 as a random<br />

walk on a noncommutative space. For this we consider the group SU(2) of<br />

unitary 2×2 matrices with determ<strong>in</strong>ant 1. It is well known that the irreducible<br />

representations of this group are f<strong>in</strong>ite dimensional, moreover, for each <strong>in</strong>teger<br />

n ≥ 1 there exists, up to equivalence, exactly one irreducible representation<br />

of dimension n, therefore one has<br />

A(SU(2)) = ⊕ ∞ n=1 Mn(C)


Introduction to Random Walks on Noncommutative Spaces 77<br />

as a von Neumann algebra direct sum and similarly the C ∗ algebra of SU(2)<br />

can be identified with the algebra of sequences (Mn)n≥1 where Mn is an<br />

n × n matrix and one has �Mn� →0asn →∞. We shall <strong>in</strong>terpret this<br />

C ∗ -algebra as the space of “functions” vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity on some noncommutative<br />

space. The associated von Neumann algebra corresponds to the<br />

space of “bounded functions” on our noncommutative space. In order to get a<br />

better picture of this space it is desirable to have a geometric understand<strong>in</strong>g<br />

of its structure. For this we first note that this space has a cont<strong>in</strong>uous group<br />

of symmetries, s<strong>in</strong>ce the group SU(2) acts on the algebra by <strong>in</strong>ner automorphisms.<br />

S<strong>in</strong>ce the elements ±I act trivially, this is really an action of the<br />

quotient SU(2)/{±I} which is isomorphic with the group SO(3), therefore<br />

this space has a three dimensional rotational symmetry. We can understand<br />

this symmetry by look<strong>in</strong>g at some special elements <strong>in</strong> the larger algebra<br />

Â(SU(2)), which correspond to “unbounded functions”. Let us consider the<br />

self-adjo<strong>in</strong>t elements correspond<strong>in</strong>g to the Pauli matrices, viewed as elements<br />

of the Lie algebra of SU(2) (or rather the complexified Lie algebra). These<br />

def<strong>in</strong>e unbounded operators on L2 (SU(2)), which we shall denote X, Y, Z,<br />

and which lie <strong>in</strong> Â(SU(2)). A good way to th<strong>in</strong>k about these three functions<br />

is as three “coord<strong>in</strong>ates” on our space, correspond<strong>in</strong>g to three orthogonal<br />

directions. Each of these elements has a spectrum which is exactly the set of<br />

<strong>in</strong>tegers (which you can view as the group dual to the one parameter group<br />

generated by one of these Lie algebra elements). Moreover this is true also of<br />

any l<strong>in</strong>ear comb<strong>in</strong>ation xX + yY + zZ with x2 + y2 + z2 = 1. This means that<br />

if you are <strong>in</strong> this space and try to measure your position, you can measure,<br />

as <strong>in</strong> quantum mechanics, one coord<strong>in</strong>ate <strong>in</strong> some direction (x, y, z) us<strong>in</strong>gthe<br />

operator xX + yY + zZ, and you will always f<strong>in</strong>d an <strong>in</strong>teger. Thus the space<br />

has some discrete feature, <strong>in</strong> that you always get <strong>in</strong>teger numbers for your<br />

coord<strong>in</strong>ates, but also a cont<strong>in</strong>uous rotational symmetry which comes from the<br />

action of SU(2) by automorphisms of the algebra. This is obviously impossible<br />

to obta<strong>in</strong> <strong>in</strong> a classical space. Of course s<strong>in</strong>ce the operators <strong>in</strong> different<br />

directions do not commute, you cannot measure your position <strong>in</strong> different<br />

directions of space simultaneously. What you can do nevertheless is measure<br />

simultaneoulsy one coord<strong>in</strong>ate <strong>in</strong> space, and your distance to the orig<strong>in</strong>. This<br />

last measurement is done us<strong>in</strong>g the operator D = √ I + X2 + Y 2 + Z2 − I<br />

which is <strong>in</strong> the center of the algebra Â(SU(2)), and therefore can be measured<br />

simultaneously with any other operator. Its eigenvalues are the nonnegative<br />

<strong>in</strong>tegers 0, 1, 2 ..., and its spectral projections are the identity elements of<br />

the algebras Mn(C), more precisely, one has <strong>in</strong> Â(SU(2))<br />

D =<br />

∞�<br />

(n − 1)IMn(C) n=1<br />

We thus see that the subalgebra Mn(C) is a k<strong>in</strong>d of “noncommutative<br />

sphere of radius n − 1”, and moreover by look<strong>in</strong>g at the eigenvalues of<br />

the operators xX + yY + zZ <strong>in</strong> the correspond<strong>in</strong>g representation, we see


78 P. Biane<br />

A(SU(2))<br />

Fig. 1 Noncommutative space underly<strong>in</strong>g A(SU(2)).<br />

that on any “radius” of this sphere, correspond<strong>in</strong>g to a direction of space,<br />

the coord<strong>in</strong>ate on this radius can only take the n values n − 1,n− 3,<br />

n − 5,...,−n +1.<br />

If we rescale the noncommutative sphere of large radius to have radius<br />

1, it looks more and more like a classical sphere, see e.g. [Rie] for a precise<br />

statement.<br />

5.2 Construction of the Random Walk<br />

Let ω be a state on M2(C), which we can also consider as a state on<br />

A(SU(2)) by the projection A(SU(2)) → M2(C). Let us consider the <strong>in</strong>f<strong>in</strong>ite<br />

tensor product algebra, with respect to the product state ν = ω ⊗∞ on<br />

N = ⊗ ∞ 1 A(SU(2))).<br />

Let T : N → N be def<strong>in</strong>ed by ∆ ⊗ s where s : A(SU(2)) [2,∞[ →<br />

A(SU(2)) [3,∞[ is the obvious shift isomorphism. Let jn : A(SU(2)) →N<br />

be the morphisms def<strong>in</strong>ed by by jn = T n ◦ i where i(a) =a ⊗ I ∞ is the<br />

GNS representation of A(SU(2)) associated with ω, act<strong>in</strong>g on the first factor.<br />

Note that one has actually N = ⊗ ∞ 1 M2(C) andi(a) =ρ2(a) ⊗ I ∞ where<br />

ρ2 is a two dimensional irreducible representation. The morphisms jn can be<br />

extended to the large algebra Â(SU(2)) and for V <strong>in</strong> the Lie algebra, us<strong>in</strong>g<br />

the formula for the coproduct one checks that<br />

jn(V )=<br />

n�<br />

i=1<br />

I ⊗(i−1) ⊗ V ⊗ I ⊗∞


Introduction to Random Walks on Noncommutative Spaces 79<br />

Thus the quantum Bernoulli random walk is obta<strong>in</strong>ed by tak<strong>in</strong>g the operators<br />

(jn(X),jn(Y ),jn(Z)) where X, Y, Z are the Lie algebra elements correspond<strong>in</strong>gtothePaulimatrices.Forn<br />

≥ 0, we let Nn be the algebra generated by<br />

the first n factors <strong>in</strong> the tensor product. There exists a conditional expectation<br />

ν(.|Nn) with respect to the state ν onto such a subalgebra, it is given<br />

simply by I ⊗ω ∞ on the factorization N = A(SU(2)) ⊗n ⊗A(SU(2)) ⊗[n+1,∞[ .<br />

The family of morphisms (jn)n≥1 form a noncommutative process wih<br />

values <strong>in</strong> the dual of SU(2). The follow<strong>in</strong>g proposition is left to the reader,<br />

as an exercise <strong>in</strong> manipulation of coproducts.<br />

Proposition 5.1. The family of morphisms jn; n ≥ 1, together with the family<br />

of algebras N , Nn; n ≥ 1, and the conditional expectations ν(.|Nn), form<br />

a dilation of the completely positive map A(SU(2)) →A(SU(2)) given by<br />

Φω =(I ⊗ ω) ◦ ∆.<br />

Let us now consider the three one-parameter subgroups generated by the<br />

Pauli matrices, they consist respectively of the matrices<br />

� � � � �<br />

iθ cos θ is<strong>in</strong> θ cos θ s<strong>in</strong> θ e 0<br />

i s<strong>in</strong> θ cos θ − s<strong>in</strong> θ cos θ 0 e−iθ �<br />

θ ∈ [0, 2π[<br />

Each of these subgroups generates a commutative von Neumann subalgebra<br />

of A(SU(2)), which is isomorphic with the group von Neumann algebra of<br />

the group U(1) of complex numbers of modulus 1. Such a von Neumann algebra<br />

is isomorphic, by Pontryag<strong>in</strong> duality, to the algebra of bounded functions<br />

on the dual group, therefore the restriction of the dilation to this subalgebra<br />

provides a random walk on this dual group, which is isomorphic to Z. Thus<br />

we recover, from the abstract considerations on duals of compact groups, our<br />

concrete Bernoulli random walks. In terms of our noncommutative space, we<br />

can observe a particle undergo<strong>in</strong>g this random walk along any fixed direction<br />

of space, and what we see is a Bernoulli random walk (recall that the<br />

coord<strong>in</strong>ate <strong>in</strong> some fixed direction of space can only take <strong>in</strong>teger values).<br />

We now turn to the sp<strong>in</strong> process which can be <strong>in</strong>terpreted <strong>in</strong> terms of the<br />

restriction of the dilation (jn)n≥1 to the center of the group algebra A(G).<br />

This center consists of operators of the form (mχ; χ ∈ ˆ G)whereeachmχ is<br />

ascalaroperator<strong>in</strong>Mχ, it is a commutative algebra, isomorphic with the<br />

algebra of bounded functions on ˆ G (recall that we have assumed that ˆ G is<br />

countable). Equivalently it is the algebra generated by the operator D i.e.<br />

the algebra of operators of the form f(D) wheref is a bounded complex<br />

function on the nonnegative <strong>in</strong>tegers. It is also easy to compute the restriction<br />

of the completely positive map Φω to this center. It is given by the<br />

Clebsch-Gordan formula which computes the decomposition <strong>in</strong>to irreducible<br />

of a tensor product of representations of the group SU(2). What we need is<br />

the formula (where ρn is the n-dimensional irreducible representation)<br />

ρ2 ⊗ ρn = ρn−1 ⊕ ρn+1


80 P. Biane<br />

which tells us that the Markov cha<strong>in</strong> can jump from n to either n−1 orn+1,<br />

furthermore the transition probabilities are proportional to the dimensions<br />

of the targets, s<strong>in</strong>ce we are <strong>in</strong> the trace state. We see that the restriction of<br />

jn to this algebra thus corresponds to the sp<strong>in</strong> process.<br />

F<strong>in</strong>ally there are other commutative algebras which are <strong>in</strong>variant under<br />

the completely positive map. They are generated by the center of A(SU(2))<br />

and by a one parameter subgroup. Each such algebra is a maximal abelian<br />

subalgebra of A(SU(2)), and its spectrum can be identified with the set of<br />

pairs (m, n) of <strong>in</strong>tegers such that n ≥ 1andm ∈{n−1,n−3,...−n+1}.One<br />

can compute the associated Markov semigroup, us<strong>in</strong>g the Clebsch Gordan<br />

formulas for products of coefficient functions of the group SU(2). However<br />

the images of such an algebra by the morphisms (jn)n≥1 do not commute.<br />

Thus <strong>in</strong> this way we get a noncommutative dilation of a purely commutative<br />

semigroup. We will come back to this Markov cha<strong>in</strong> when we study Pitman’s<br />

theorem and quantum groups.<br />

6 Random Walks on Duals of Compact Groups<br />

It is easy to generalize the construction of the preced<strong>in</strong>g section by replac<strong>in</strong>g<br />

the group SU(2) by an arbitrary compact group G. We will do a construction<br />

parallel to the usual construction of a random walk on a group. Let φ0 and φ<br />

be cont<strong>in</strong>uous positive def<strong>in</strong>ite functions on G, with φ0(e) =φ(e) =1,thus<br />

these functions correspond to normal states ν0 and ν on A(G). The state ν0<br />

will play the role of <strong>in</strong>itial condition of our Markov cha<strong>in</strong>, whereas the state<br />

ν represents the distribution of the <strong>in</strong>crements. To the state ν is associated<br />

a completely positive map<br />

Φν : A(G) →A(G) Φν =(I ⊗ ν) ◦ ∆.<br />

The completely positive map generates a semigroup Φ n ν ; n ≥ 1. We now build<br />

the <strong>in</strong>f<strong>in</strong>ite tensor product N = A(G) ∞ with respect to the state ν0 ⊗ ν ⊗∞ ,<br />

and obta<strong>in</strong> a noncommutative probability space (N ,ω). Let T : N → N<br />

def<strong>in</strong>ed by ∆⊗s where s : A(G) [1,∞[ →A(G) [2,∞[ is the obvious isomorphism.<br />

Let jn : A(G) →N be the morphisms def<strong>in</strong>ed by <strong>in</strong>duction jn = T n ◦ i where<br />

i(a) =a ⊗ I ∞ is the <strong>in</strong>clusion of A(G) <strong>in</strong>to the first factor (strictly speak<strong>in</strong>g<br />

this is an <strong>in</strong>clusion only if the state ν0 is faithful).<br />

Let us translate the above construction <strong>in</strong> the case of an abelian group,<br />

<strong>in</strong> terms of the dual group ˆ G.Thestatesν and ν0 correspond to probability<br />

measures on ˆ G, the probability space is now the product of an <strong>in</strong>f<strong>in</strong>ite number<br />

of copies of ˆ G, with the product probability ν0 ⊗ ν ⊗∞ ,andthemapsjn<br />

correspond to functions Xn : ˆ G ∞ → G given by Xn(g0,g1,...,gk,...)=<br />

g0g1 ...gn−1. We thus recover the usual construction of a random walk.


Introduction to Random Walks on Noncommutative Spaces 81<br />

For n ≥ 0, we let Nn be the algebra generated by the first n +1 factors<br />

<strong>in</strong> the tensor product. There exists a conditional expectation ω(.|Nn) with<br />

respect to the state ω onto such a subalgebra, it is given simply by I ⊗ ν ∞<br />

on the factorization N = A(G) ⊗(n+1) ⊗A(G) ⊗∞ .<br />

Proposition 6.1. The morphisms (jn)n≥0, together with the von Neumann<br />

algebras N , Nn,n ≥ 0 and the state ω, form a dilation of the completely<br />

positive semigroup (Φ n ν )n≥1, with <strong>in</strong>itial distribution ν0.<br />

The proof of this proposition follows exactly the case of SU(2) treated <strong>in</strong> the<br />

preced<strong>in</strong>g section.<br />

Let H be a closed commutative subgroup of G, then its dual group ˆ H<br />

is a countable discrete abelian group. The von Neumann subalgebra A(H)<br />

generated by H <strong>in</strong> A(G) is isomorphic to L ∞ ( ˆ H), and the restriction of<br />

the positive def<strong>in</strong>ite function φ to H is the Fourier transform of a probability<br />

measure µ on ˆ H. The coproduct ∆ restricts to a coproduct on the subalgebra<br />

generated by H, thus the images of A(H) by the morphisms jn generate<br />

commut<strong>in</strong>g subalgebras of N . These restrictions thus give a random walk on<br />

the dual group ˆ H, whose <strong>in</strong>dependent <strong>in</strong>crements are distributed accord<strong>in</strong>g<br />

to µ.<br />

We now consider another commutative algebra, namely the center Z(G) of<br />

A(G). Recall that this center is isomorphic with the space L ∞ ( ˆ G) of bounded<br />

functions on the set of equivalence classes of irreducible representations of G.<br />

As a von Neumann algebra of operators on L 2 (G), it is generated by the<br />

convolution operators by <strong>in</strong>tegrable central functions (recall that a function<br />

f on G is central if it satisfies f(ghg −1 )=f(h) for all h, g ∈ G).<br />

Proposition 6.2. The algebras jn(Z(G)); n ≥ 1 commute.<br />

Proof. Let a, b ∈Z(G) andletk ≤ l we have to prove that jk(a) andjl(b)<br />

commute. Let a ′ = i(a),b ′ = i(b), and note that a ′ and b ′ belong to the center<br />

of N . One has<br />

jk(a)jl(b) =T k (a ′ )T k (T l−k (b ′ ))<br />

= T k (a ′ T l−k (b ′ ))<br />

= T k (T l−k (b ′ )a ′ )<br />

= T l (b ′ )T k (a ′ )<br />

= jl(b)jk(a)<br />

Furthermore one has.<br />

Proposition 6.3. If the function φ is central then Φν(Z(G)) ⊂Z(G).<br />

Proof. Indeed if ψ is central function, belong<strong>in</strong>g to A 0 (G), then Φν(ψ) =φψ<br />

is a central function on the group, and thus def<strong>in</strong>es an element of Z(G). ⊓⊔<br />

We deduce from the preced<strong>in</strong>g propositions that, <strong>in</strong> the case when the <strong>in</strong>crements<br />

correspond to a central state, the restriction of the dilation (jn)n≥0<br />

to the center of the algebra A(G) def<strong>in</strong>es a Markov process on ˆ G,whose<br />

⊓⊔


82 P. Biane<br />

transition operator is given by the restriction of Φν to the center Z(G). We<br />

shall now give a more concrete form of the transition probabilites of this<br />

Markov cha<strong>in</strong>.<br />

Proposition 6.4. For χ an irreducible character of G, let<br />

φχ = �<br />

hφ(χ ′ ,χ)χ ′<br />

χ ′ ∈ ˆ G<br />

be the expansion of the positive def<strong>in</strong>ite central function φχ <strong>in</strong>to a comb<strong>in</strong>ation<br />

characters, then the probability transitions of the Markov cha<strong>in</strong> obta<strong>in</strong>ed from<br />

the restriction of (jn)n≥1 to the center are given by<br />

pφ(χ ′ ,χ)= dim(χ)<br />

dim(χ ′ ) hφ(χ ′ ,χ)<br />

Proof. For χ ∈ ˆ G, the convolution operator associated with the function<br />

dim(χ)χ is the m<strong>in</strong>imal projection of Z(G) associated with χ. Inotherwords<br />

it corresponds to the <strong>in</strong>dicator function 1χ <strong>in</strong> the isomorphism of L ∞ ( ˆ G)<br />

with Z(G). The transition operator for the restriction of the dilation to the<br />

center is the restriction to this center of Φω. On the other hand, the transition<br />

probabilities pφ(χ, χ ′ ) are related to these <strong>in</strong>dicator functions by Φω(1χ ′)=<br />

�<br />

χ pφ(χ ′ ,χ)1χ. The conclusion follows immediately. ⊓⊔<br />

7TheCaseofSU(n)<br />

7.1 Some Facts About the Group SU(n)<br />

We shall <strong>in</strong>vestigate the quantum random walk def<strong>in</strong>ed <strong>in</strong> the preced<strong>in</strong>g<br />

section when the group G is the group SU(n) of unitary matrices with determ<strong>in</strong>ant<br />

1. First we recall some basic facts about this group and its representations.<br />

Let T ⊂ SU(n) be the subgroup of diagonal matrices, which is a<br />

maximal torus. The group of characters of T is an n − 1 dimensional lattice,<br />

generated by the elements (the notation e is here to suggest the exponential<br />

function)<br />

⎛ ⎞<br />

u1 0<br />

⎜<br />

e(ei)( ⎝<br />

. ..<br />

⎟<br />

⎠) =uj<br />

0 un<br />

These elements satisfy the relation (written <strong>in</strong> additive notation) e1 +<br />

e2 + ...+ en = 0, which corresponds to the relation e(e1)e(e2) ...e(en) =1<br />

for the characters. We denote by P this group, and by P+ the subset of<br />

positive weights, i.e.<br />

P+ = {m1e1 + ...+ mn−1en−1|m1 ≥ m2 ≥ ...≥ mn−1}


Introduction to Random Walks on Noncommutative Spaces 83<br />

Fig. 2<br />

We shall also need the set<br />

P++ = {m1e1 + ...+ mn−1en−1|m1 >m2 >...>mn−1}<br />

and note that the two sets are <strong>in</strong> bijection by<br />

P++ = P+ + ρ (7.1)<br />

where ρ =(n−1)e1 +(n− 2)e2 + ...+ en−1 is the half sum of positive roots.<br />

The symmetric group acts on this character group by permutation of the ei.<br />

Below is a picture of P and P+ for the group SU(3). Thus P consists of<br />

the po<strong>in</strong>ts <strong>in</strong> a triangular lattice <strong>in</strong> the plane, and P+ is the <strong>in</strong>tersection of<br />

this triangular lattice with a cone, fundamental doma<strong>in</strong> for the action of the<br />

symmetric group S3. The subset P++ consists <strong>in</strong> po<strong>in</strong>ts of P+ which are <strong>in</strong><br />

the <strong>in</strong>terior of the cone.<br />

Recall from the representation theory of the group SU(n) (seee.g.[BtD],<br />

[GW], or [Z]) that the equivalence classes of irreducible representations of<br />

SU(n) are <strong>in</strong> one to one correspondence with the elements of P+, whichare<br />

called “highest weights”. For each x ∈ P let e(x) be the associated character<br />

of T, then the character of the representation with highest weight x ∈ P+<br />

given, for u ∈ T, by Weyl’s character formula<br />

�<br />

ɛ(σ)e(σ(x + ρ))(u)<br />

σ∈Sn χx(u) = �<br />

σ∈Sn ɛ(σ)e(σ(ρ))(u)<br />

(7.2)<br />

In particular the def<strong>in</strong><strong>in</strong>g representation of SU(n) has character e(e1)+...+<br />

e(en) correspond<strong>in</strong>g to the highest weight e1. The normalized positive def<strong>in</strong>ite


84 P. Biane<br />

Fig. 3 The random walk on the dual of the maximal torus for SU(3).<br />

1/3<br />

1/3<br />

function on SU(n) correspond<strong>in</strong>g to this character is φ(g) = 1<br />

nTr(g). We shall<br />

<strong>in</strong>vestigate the quantum random walk associated with this positive def<strong>in</strong>ite<br />

function.<br />

1/3<br />

7.2 Two <strong>Classical</strong> Markov Cha<strong>in</strong>s<br />

We shall obta<strong>in</strong> two classical Markov cha<strong>in</strong>s by restrict<strong>in</strong>g the Markov cha<strong>in</strong><br />

associated with φ to suitable subalgebras of C∗ (SU(n)). The first subalgebra<br />

is that generated by the maximal torus T. This algebra is isomorphic to the<br />

algebra of functions vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity on the dual group P. Itiseasyto<br />

see that the restriction to the torus of the quantum random walk (jn)n≥0,<br />

constructed us<strong>in</strong>g φ, is a random walk on the lattice P with <strong>in</strong>crements<br />

distributed as 1<br />

n (δe1 + ... + δen). Its one step transition probabilities are<br />

given by<br />

if y ∈{x + e1,...,x+ en}<br />

p1(x, y) =0 ifnot<br />

p1(x, y) = 1<br />

n<br />

We give the picture for the case of the group SU(3).<br />

The second subalgebra is the center of C ∗ (SU(n)). By the preced<strong>in</strong>g section,<br />

it can be identified with the space of functions vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity on<br />

P+. We shall rather use the identification of the set of irreducible representations<br />

with P++ given by (7.1). As we saw <strong>in</strong> section 6 the restriction of the<br />

dilation (jn)n≥0 gives a classical Markov cha<strong>in</strong> on the spectrum of the center,<br />

therefore we obta<strong>in</strong> <strong>in</strong> this way a Markov cha<strong>in</strong> on P++, the generator of the<br />

Markov cha<strong>in</strong> be<strong>in</strong>g given by the restriction of the generator of the quantum<br />

Markov cha<strong>in</strong> on A(SU(2)).


Introduction to Random Walks on Noncommutative Spaces 85<br />

We shall now derive a relation between these two Markov cha<strong>in</strong>s (the one<br />

obta<strong>in</strong>ed from the maximal torus, and the one from the center). The set of<br />

highest weights, P++ is the <strong>in</strong>tersection of P with the Weyl chamber, which<br />

is the cone<br />

C = {x1e1 + ...+ xn−1en−1|x1 >x2 >...>xn−1}<br />

We consider now the random walk on P killed at the exit of this cone. Thus<br />

the transition probabilites of this killed random walk are given by<br />

p 0 (x, y) = 1<br />

n<br />

if y ∈ P++ ∩{x + e1,...,x+ en}<br />

p0 (x, y) =0 ifnot.<br />

The sum �<br />

y p0 (x, y) is< 1 for po<strong>in</strong>ts near the boundary, correspond<strong>in</strong>g to<br />

the fact that the random walk has a nonzero probability of be<strong>in</strong>g killed.<br />

Let x ∈ P++ and consider the irreducible representation ξx of SU(n), with<br />

highest weight x − ρ. We can use Weyl’s character formula (7.2) for decompos<strong>in</strong>g<br />

the representation ξx ⊗ ξn (where ξn is the def<strong>in</strong><strong>in</strong>g representation of<br />

SU(n)).<br />

We remark that 1 �n n j=1 e(ej) = 1 �<br />

n! σ∈Sn e(σ(e1)). One has<br />

χnχx−ρ = 1<br />

n (e(e1)+...+<br />

�<br />

σ∈Sn e(en))<br />

ɛ(σ)e(σ(x))<br />

�<br />

σ∈Sn ɛ(σ)e(σ(ρ))<br />

� �<br />

τ ∈Sn σ∈Sn =<br />

ɛ(σ)e(σ(x)+τ(e1))(u)<br />

n! �<br />

σ∈Sn ɛ(σ)e(σ(ρ))<br />

= 1<br />

�n n<br />

= 1<br />

n<br />

j=1<br />

�<br />

�<br />

σ∈Sn<br />

ɛ(σ)e(σ(x + ej))<br />

σ∈Sn ɛ(σ)e(σ(ρ))<br />

�<br />

y∈{x+e1,x+e2,...,x+en}∩P++<br />

χy−ρ<br />

s<strong>in</strong>ce if y = x + ej belongs to P+ \ P++ then it is fixed by some reflexion <strong>in</strong><br />

Sn and thus the sum �<br />

σ∈Sn ɛ(σ)e(σ(x + ej)) vanishes.<br />

For x ∈ P++, let us denote h(x) the dimension of the representation<br />

with highest weight x − ρ. We conclude from the preced<strong>in</strong>g computation and<br />

Proposition 6.4, that the transition probabilities for the Markov cha<strong>in</strong> on<br />

P++ are<br />

q(x, y) = h(y)<br />

h(x) p0 (x, y) x, y, ∈ P++ (7.3)<br />

S<strong>in</strong>ce the transition operator is unit preserv<strong>in</strong>g, it follows <strong>in</strong> particular that<br />

the function h is a positive harmonic function with respect to the transition<br />

kernel p0 , i.e. satisfies


86 P. Biane<br />

Fig. 4<br />

a<br />

c<br />

h(a)/3h(x)<br />

x<br />

h(c)/3h(x)<br />

h(x) = �<br />

h(y)p 0 (x, y) for all x ∈ P++,<br />

y<br />

h(b)/3h(x)<br />

Aga<strong>in</strong> we draw the picture <strong>in</strong> the case of SU(3).<br />

Let us recall that, given transition (sub)probabilities p(x, y) for a Markov<br />

cha<strong>in</strong>, and a positive harmonic function h for p, i.e. a function such that<br />

h(x) = �<br />

h(y)p(x, y) for all x<br />

y<br />

one calls the Markov cha<strong>in</strong> with transition probabilities h(y)<br />

h(x) p(x, y) the<br />

Doob condition<strong>in</strong>g, or h-transform, of the Markov cha<strong>in</strong> with transition<br />

probabilities p.<br />

We can summarize the preced<strong>in</strong>g discussion <strong>in</strong> the follow<strong>in</strong>g proposition.<br />

Proposition 7.1. The Markov cha<strong>in</strong> obta<strong>in</strong>ed by restriction to the center<br />

is related to the random walk on the dual of the maximal torus by a kill<strong>in</strong>g<br />

at the exit of the Weyl chamber followed by a Doob condition<strong>in</strong>g us<strong>in</strong>g the<br />

dimension function on the set of highest weights.<br />

For more <strong>in</strong>formation on Doob’s condition<strong>in</strong>g, I refer to the books by<br />

Kemeny, Knapp and Snell [KKS], or Revuz [R].<br />

One could ask whether this relation, between the Markov cha<strong>in</strong>s on the<br />

dual of the torus and on the center, holds for more general groups and positive<br />

def<strong>in</strong>ite functions. It turns out that the fundamental concept <strong>in</strong> this direction<br />

is that of a m<strong>in</strong>uscule weight, see [B3] for more details.<br />

b


Introduction to Random Walks on Noncommutative Spaces 87<br />

8 Choquet-Deny Theorem for Duals of Compact Groups<br />

8.1 The Choquet-Deny Theorem <strong>in</strong> an Abelian Group<br />

As we have seen <strong>in</strong> the preced<strong>in</strong>g section, the fact that the dimension function<br />

is a positive harmonic function on the Weyl chamber plays an important role<br />

<strong>in</strong> understand<strong>in</strong>g the quantum random walk on the dual of SU(n). A natural<br />

question arises, is this positive harmonic function unique? We shall answer<br />

this question, which is purely a question of “classical” potential theory us<strong>in</strong>g<br />

the theory of quantum random walks. Actually we shall do this by extend<strong>in</strong>g a<br />

well known result of Choquet and Deny [CD] on harmonic measures for locally<br />

compact abelian goups. The Choquet-Deny theorem describes completely the<br />

solutions of the convolution equation<br />

µ ∗ φ = φ (8.1)<br />

on a locally compact abelian group G,whereµ is a positive f<strong>in</strong>ite measure, and<br />

φ is an unknown positive measure. One assumes that the subgroup generated<br />

by the support of µ is the whole group. In order to state the Choquet-Deny<br />

theorem, one needs to def<strong>in</strong>e the exponentials on the group G. These are the<br />

cont<strong>in</strong>uous functions f on G, withvalues<strong>in</strong>]0, +∞[, which are multiplicative,<br />

i.e. satisfy<br />

f(gh) =f(g)f(h)<br />

for all g, h ∈ G. An exponential e : G →]0, +∞[ is called µ-harmonic if one<br />

has �<br />

G<br />

e(−x)µ(dx) = 1. The set of µ-harmonic exponentials is a Borel subset<br />

of the set of all cont<strong>in</strong>uous functions on G, which we denote by Eµ. Letφbe a positive measure on G of the form φ(dx) =e(x)dx where dx is the Haar<br />

measure on G and e is a µ-harmonic exponential, then one has for all positive<br />

measurable functions f<br />

�<br />

�<br />

G f(x)φ ∗ µ(dx) =�G<br />

G f(x + y)φ(dx)µ(dy)<br />

= � �<br />

G G f(x + y)e(x)dxµ(dy)<br />

= � �<br />

G G f(x)e(x − y)dxµ(dy)<br />

= �<br />

�<br />

f(x)e(x)dx G G e(−y)µ(dy)<br />

= �<br />

G f(x)φ(dx)<br />

therefore the measure φ is µ-harmonic. i.e. satisfies the equation (8.1). The<br />

Choquet-Deny theorem states that every solution is a convex comb<strong>in</strong>ation of<br />

solutions of this k<strong>in</strong>d.<br />

Theorem 8.1 (Choquet-Deny). Assume that the subgroup generated by<br />

the support of the measure µ is G, then every positive measure φ, solution<br />

of the convolution equation (8.1), is absolutely cont<strong>in</strong>uous with respect ot the<br />

Haar measure on G, and its density has a unique representation as an <strong>in</strong>tegral


88 P. Biane<br />

dφ(x)<br />

dx =<br />

�<br />

e(x)dνφ(e)<br />

where νφ is some f<strong>in</strong>ite positive measure on Eµ.<br />

Eµ<br />

This result conta<strong>in</strong>s <strong>in</strong> particular the determ<strong>in</strong>ation of all positive harmonic<br />

functions for the transition operator associated with the measure µ,<br />

<strong>in</strong>deed if h is such a positive µ-harmonic function, then the measure h(−x)dx<br />

is a µ-harmonic measure.<br />

Note that a locally compact abelian group corresponds to a commutative<br />

and cocommutative Hopf C ∗ -algebra. What we shall do next is to extend this<br />

theorem to (a class of) cocommutative Hopf C ∗ -algebras. This will allow us to<br />

answer the question about the uniqueness of the positive harmonic function.<br />

Before we state our analogue of the Choquet-Deny theorem on the dual of a<br />

compact group, we will first clarify some po<strong>in</strong>ts about states and weights.<br />

8.2 Some Further Properties of Duals of Compact<br />

Groups<br />

Let G be a compact group and A(G) be its von Neumann algebra. This<br />

von Neumann algebra has a structure of cocommutative Hopf-von Neumann<br />

algebra for the coproduct ∆, counit ε and antipode i given respectively by<br />

cont<strong>in</strong>uous l<strong>in</strong>ear extension of<br />

∆(λg) =λg ⊗ λg ε(λg) =δe,g i(λg) =λ g −1.<br />

We have seen <strong>in</strong> section 4.3 that the convolution of two f<strong>in</strong>ite normal weights<br />

µ and ν on A(G) is def<strong>in</strong>ed by the formula µ ∗ ν =(µ ⊗ ν) ◦ ∆. Assume<br />

now that µ is a positive f<strong>in</strong>ite weight and φ is positive, then we can write φ<br />

as a sum of f<strong>in</strong>ite weights φ = �<br />

χ φχ with respect to the restrictions of φ<br />

to the subalgebras Mχ. Thesum �<br />

χ ν ∗ φχ is then a sum of positive f<strong>in</strong>ite<br />

weights and thus it def<strong>in</strong>es a (not necessarily normal) weight on A(G). We<br />

can therefore consider the equation (8.1) where this time µ and φ are positive<br />

normal weights, with µ f<strong>in</strong>ite.<br />

Next we def<strong>in</strong>e the notion of exponential. The follow<strong>in</strong>g def<strong>in</strong>ition is a<br />

straighforward extension of the def<strong>in</strong>ition <strong>in</strong> the case of abelian groups.<br />

Def<strong>in</strong>ition 8.2. A group-like element isanonzeroelementfof Â(G), such<br />

that<br />

∆(f) =f ⊗ f.<br />

If this element is positive, then we call it an exponential.<br />

If µ is a f<strong>in</strong>ite weight, f is an exponential and µ(i(f)) = 1 then we call f<br />

a µ-harmonic exponential.<br />

Let f,g,h be positive elements <strong>in</strong> Â(G) then one has the identity<br />

tr(f ⊗ g∆(h)) = tr(i(h) ⊗ g∆(i(f))) (8.2)


Introduction to Random Walks on Noncommutative Spaces 89<br />

which is easily checked on coefficient functions. Let g ∈ Â(G) be associated<br />

to µ, i.e. µ(.) =tr(g.), then it follows from 8.2 that for all h<br />

φ ∗ µ(h) =tr(f ⊗ g∆(h))<br />

= tr(i(h) ⊗ g∆(i(f)))<br />

= tr(i(h)i(f))tr(gi(f))<br />

= φ(h)<br />

Thus if f is a µ-harmonic exponential, and φ is the weight associated with<br />

f, thenφ satisfies the equation (8.1).<br />

8.3 The Analogue of the Choquet-Deny Theorem<br />

We shall assume that the state ν satisfies a non degeneracy condition which<br />

is the analogue, <strong>in</strong> the noncommutative sett<strong>in</strong>g, of the requirement that the<br />

support of the measure generates the whole group. This condition states that<br />

for every f<strong>in</strong>ite weight ρ on A(G) which is supported by some algebra Mχ for<br />

χ ∈ ˆ G, there exists an <strong>in</strong>teger n ≥ 1andaconstantc>0 such that cν∗n ≥ ρ.<br />

Theorem 8.3. Let φ be a ν-harmonic weight, then there exists a unique f<strong>in</strong>ite<br />

positive measure on the set of ν-harmonic exponentials such that<br />

�<br />

φ = edmφ(e)<br />

Eµ<br />

Sketch of proof. We let Sν be the convex cone of normal weights satisfy<strong>in</strong>g<br />

φ ∗ ν ≤ φ<br />

This cone is a closed subset of Â, and we can use Choquet’s <strong>in</strong>tegral representation<br />

theorem to conclude that any element of this cone can be written<br />

as the barycenter of a measure supported on the set of its extremal rays. Now<br />

one can analyze the extremal rays of this set, and see that such an extremal<br />

ray consists <strong>in</strong> multiples of an exponential. The uniqueness argument comes<br />

from the existence of an algebra of functions on the cone, which separates<br />

po<strong>in</strong>ts, see [B4] for details.<br />

8.4 Examples<br />

For a f<strong>in</strong>ite group, the set of exponentials is reduced to the identity.<br />

If the state ν is tracial, then there exists only one ν-harmonic exponential,<br />

namely the identity. In the case of the group SU(n), the exponentials are <strong>in</strong><br />

one to one correspondence with the positive elements <strong>in</strong> SL(n, C).


90 P. Biane<br />

From these results we deduce that the positive harmonic function h of<br />

Proposition 7.1 was unique. Indeed <strong>in</strong> order to prove that a function h is<br />

the unique (up to a multiplicative constant) positive harmonic function for<br />

a transition kernel p, it is enough to check that all the positive harmonic<br />

functions for the relativized kernel h(y)<br />

h(x) p(x, y) are constant, which is what<br />

the Choquet-Deny theorem tells us.<br />

9 The Mart<strong>in</strong> Compactification of the Dual of SU(2)<br />

In the preced<strong>in</strong>g section we have seen that the classical Choquet-Deny theorem<br />

about solutions of convolution equations has an analogue <strong>in</strong> duals of compact<br />

groups. This allows one to give an explicit description of all µ-harmonic<br />

positive functions for a f<strong>in</strong>ite positive measure µ on a commutative group.<br />

By the Choquet-Deny theorem the positive µ-harmonic functions admit a<br />

unique <strong>in</strong>tegral representation <strong>in</strong> terms of m<strong>in</strong>imal µ-harmonic functions,<br />

and these m<strong>in</strong>imal harmonic functions can be identified with exponentials.<br />

The next natural question <strong>in</strong> this l<strong>in</strong>e of ideas is to describe the Mart<strong>in</strong> compactification<br />

associated to a random walk with values <strong>in</strong> Z n .ThisMart<strong>in</strong><br />

compactification provides a way to attach a boundary to the space Z n <strong>in</strong><br />

order to obta<strong>in</strong> a compact space, where the boundary is naturally identified<br />

with the set of m<strong>in</strong>imal µ-harmonic functions. The Mart<strong>in</strong> compactification<br />

of Z n was computed by Ney and Spitzer <strong>in</strong> a classical paper [NS], where<br />

they showed that it consists <strong>in</strong> add<strong>in</strong>g a sphere at <strong>in</strong>f<strong>in</strong>ity, and identify<strong>in</strong>g<br />

this sphere with the set of m<strong>in</strong>imal harmonic functions with the help of the<br />

Gauss map. In this section we will describe the Mart<strong>in</strong> compactification of<br />

some quantum random walks with values <strong>in</strong> the dual of SU(2). We will first<br />

recall some basic facts about classical Mart<strong>in</strong> boundaries, then describe Ney<br />

and Spitzer’s theorem, before go<strong>in</strong>g to the case of the dual of SU(2).<br />

9.1 The Mart<strong>in</strong> Compactification for Markov Cha<strong>in</strong>s<br />

Consider a Markov cha<strong>in</strong> on a countable state space E. TheMarkovcha<strong>in</strong><br />

has transition subprobabilities p(x, y),x,y ∈ E, i.e. we have �<br />

y p(x, y) ≤ 1,<br />

so that the kernel is submarkovian and the process may die <strong>in</strong> a f<strong>in</strong>ite time.<br />

There is an associated transition operator given by<br />

Pf(x) = �<br />

p(x, y)f(y)<br />

y∈E<br />

and the iterated operator is given by n-step transition probabilities<br />

P n f(x) = �<br />

pn(x, y)f(y)<br />

E


Introduction to Random Walks on Noncommutative Spaces 91<br />

We assume that the associated Markov cha<strong>in</strong> is transient, so that the potential<br />

is f<strong>in</strong>ite, i.e. the function<br />

u(x, y) =<br />

U =<br />

∞�<br />

P n<br />

n=0<br />

∞�<br />

pn(x, y) < ∞<br />

n=0<br />

Let us choose an <strong>in</strong>itial distribution r(dx) such that the function rU(y) =<br />

u(x, y)r(dx) iseverywhere> 0. The Mart<strong>in</strong> kernel is def<strong>in</strong>ed by<br />

�<br />

E<br />

k(x, y) =<br />

u(x, y)<br />

rU(y)<br />

It follows from the Harnack <strong>in</strong>equalities that the functions k(x, .) form a uniformly<br />

cont<strong>in</strong>uous family on E. The Mart<strong>in</strong> compactification of the Markov<br />

cha<strong>in</strong> is the smallest compact topological space Ēu, which conta<strong>in</strong>s E as a<br />

dense subset, and such that these functions extend cont<strong>in</strong>uously to the boundary<br />

∂Eu = Ēu \ E. This space exists because the functions k(x, .) separate<br />

the po<strong>in</strong>ts of E and because of the uniform cont<strong>in</strong>uity. For any ξ ∈ ∂Eu<br />

the function x ↦→ k(x, ξ) isap-harmonic function. Recall that a positive pharmonic<br />

function f on E is called m<strong>in</strong>imal if for every positive p-harmonic<br />

function g satisfy<strong>in</strong>g g ≤ Cf for some C>0, one has actually g = cf for<br />

some constant c. One can prove that any m<strong>in</strong>imal p-harmonic function f,<br />

which is r-<strong>in</strong>tegrable, is a multiple of k(., ξ) forsomeξ∈∂Eu. The subset<br />

∂Em of ξ ∈ ∂Eu such that k(., ξ) is m<strong>in</strong>imal is a Borel subset, and one can<br />

prove that any positive p-harmonic function f, whichisr-<strong>in</strong>tegrable, admits<br />

arepresentation<br />

�<br />

f = k(., ξ)dmf (ξ)<br />

∂Eu<br />

with a unique positive measure mf .<br />

9.2 The Mart<strong>in</strong> Compactification of Z d<br />

When the Markov cha<strong>in</strong> is a random walk on Zd , with <strong>in</strong>crements distributed<br />

as µ, a (sub)probability measure on Zd , we have seen that every positive µharmonic<br />

function admits an <strong>in</strong>tegral representation <strong>in</strong> terms of exponentials.<br />

When the <strong>in</strong>crements of the random walk are <strong>in</strong>tegrable, Ney and Spitzer have<br />

determ<strong>in</strong>ed explicitly the Mart<strong>in</strong> compactification of Zd ,whichweshallnow<br />

describe. Let φ : Rd → [0, +∞] be the function<br />

φ(x) = �<br />

e 〈x,y〉 µ(dy)<br />

y∈Z d


92 P. Biane<br />

We assume that it is f<strong>in</strong>ite <strong>in</strong> a neighbourhood of the set<br />

Eµ = {x|φ(x) =1}<br />

Then the set Eµ is the boundary of the convex set {x|φ(x) ≤ 1} and it is<br />

either reduced to a po<strong>in</strong>t or homeomorphic to a sphere. In the latter case the<br />

homeomorphism can be expressed thanks to the Gauss map as<br />

∇φ<br />

�∇φ�<br />

S<strong>in</strong>ce φ is convex this is a homeomorphism from Eµ onto the unit sphere. Ney<br />

and Spitzer proved that the Mart<strong>in</strong> compactification is homeomorphic to the<br />

usual compactification of Z d by a sphere at <strong>in</strong>f<strong>in</strong>ity, where the identification<br />

between the sphere and the set of m<strong>in</strong>imal µ-harmonic functions is provided<br />

by the map above.<br />

9.3 Noncommutative Compactifications<br />

Before we <strong>in</strong>vestigate the problem of f<strong>in</strong>d<strong>in</strong>g an analogue of the Ney-Spitzer<br />

theorem for the dual of SU(2) let us translate <strong>in</strong> the noncommutative language<br />

the notion of a compactification of a topological space. So let X be a<br />

topological space, and ¯ X a compact space such that X ⊂ ¯ X is a dense open<br />

subset, and ∂ ¯ X = ¯ X \ X is the boundary, then C( ¯ X)iscanbeidentified<br />

with a subalgebra of Cb(X), the algebra of all bounded cont<strong>in</strong>uous functions<br />

on X, and one has an exact sequence<br />

0 → C0(X) → C( ¯ X) → C(∂X) → 0<br />

where the first map is the cont<strong>in</strong>uous extension, to ¯ X, by 0 on the boundary,<br />

of a function on X, and the second map is the restriction to a closed subset.<br />

A compactification of C0(X) is thus given by a certa<strong>in</strong> commutative C∗- subalgebra of the multiplier algebra of C0(X), conta<strong>in</strong><strong>in</strong>g C0(X). In the case<br />

of the Mart<strong>in</strong> compactification this subalgebra is just the algebra generated<br />

by C0(X) and by the functions y ↦→ k(x, y). The vector space generated by<br />

functions of the form y ↦→ k(x, y) can be identified with the image of the<br />

Mart<strong>in</strong> kernel, considered as an <strong>in</strong>tegral operator f ↦→ �<br />

x∈E<br />

f(x)k(x, y). It<br />

is this <strong>in</strong>terpretation of the Mart<strong>in</strong> kernel that has a natural noncommutative<br />

analogue.<br />

9.4 The Mart<strong>in</strong> Kernel for the Quantum Random<br />

Walk<br />

We consider a central quantum random walk, therefore we have a positive<br />

def<strong>in</strong>ite central function φ on SU(2) such that φ(e) ≤ 1, and the associated


Introduction to Random Walks on Noncommutative Spaces 93<br />

state ν on A(SU(2)). We have seen that it has an associated transition kernel<br />

given by the completely positive map Φν =(I⊗ ν) ◦ ∆. This map acts as<br />

λg ↦→ φ(g)λg on the generators λg of the von Neumann algebra A(SU(2)). We<br />

know that if φ(e) = 1 then there exists only one φ-harmonic weight, namely<br />

the one given by the function 1. In this case we expect that the Mart<strong>in</strong><br />

boundary will be given by a one po<strong>in</strong>t compactification, which consists just<br />

<strong>in</strong> add<strong>in</strong>g a unit to the algebra, so <strong>in</strong> order to avoid this case, we shall assume<br />

here that φ(e) < 1.<br />

We can consider the associated potential which is equal to U = �∞ n=0 Φnν ,<br />

and which acts by multiplication by the function � ∞<br />

n=0 φn = 1<br />

1−φ .Weshall<br />

consider the action of the potential on the space of coefficients i.e. the direct<br />

sum ⊕χMχ where each element of this space can be identified with a<br />

polynomial function on SU(2) act<strong>in</strong>g by convolution, i.e. by the operator<br />

�<br />

SU(2) p(g)λgdg. Then the Mart<strong>in</strong> kernel will be def<strong>in</strong>ed, by analogy with the<br />

case of classical Markov cha<strong>in</strong>s, by<br />

�<br />

p ↦→ �<br />

SU(2)<br />

SU(2)<br />

p(g)<br />

1−φ(g) λgdg<br />

1<br />

1−φ(g) λgdg<br />

1<br />

1−φ(g) λgdg lies <strong>in</strong> the center of A(SU(2)),<br />

Note that the operator �<br />

SU(2)<br />

therefore there is no ambiguity <strong>in</strong> the quotient of the preced<strong>in</strong>g formula.<br />

9.5 Pseudodifferential Operators of Order Zero<br />

and the Mart<strong>in</strong> Compactification<br />

In order to f<strong>in</strong>d the Mart<strong>in</strong> compactification of the quantum random walk,<br />

we shall identify the C ∗ -algebra generated by C ∗ (SU(2)) and by the image of<br />

the Mart<strong>in</strong> kernel, and show that it gives rise to a three terms exact sequence.<br />

We will first exhibit a certa<strong>in</strong> exact sequence<br />

0 → C ∗ (SU(2)) →M→C(S 2 ) → 0 (9.1)<br />

where M is our sought for Mart<strong>in</strong> compactification, and C(S 2 ) is the algebra<br />

of cont<strong>in</strong>uous functions on the two dimensional sphere S2 . For this we consider<br />

the three operators <strong>in</strong> Â(SU(2)) associated with Pauli matrices, which we call<br />

X, Y, Z, and the Casimir operator C = X 2 +Y 2 +Z 2 +I which acts by (n+1) 2<br />

on the space of coefficients of the n-dimensional representation of SU(2), and<br />

build three operators<br />

x = XC −1/2 ,y = YC −1/2 ,z = ZC −1/2<br />

Clearly these operators are self-adjo<strong>in</strong>t and bounded. Us<strong>in</strong>g the commutation<br />

relations (3.1) one can see that [x, y], [x, z], [y, z] andx 2 + y 2 + z 2 − I are


94 P. Biane<br />

compact operators on L 2 (SU(2)). It follows that there exists a map from<br />

the algebra generated by x, y, z to the algebra of polynomial function on<br />

the sphere, send<strong>in</strong>g x, y, z to the three coord<strong>in</strong>ate functions, and this map<br />

vanishes on the compact operators. Actually this map extends by cont<strong>in</strong>uity<br />

to the C ∗ algebra generated by x, y and z and yields the exact sequence (9.1).<br />

One can <strong>in</strong>terpret also the algebra M as the algebra of right <strong>in</strong>variant pseudo<br />

differential operators of order zero on SU(2), then the map M→C(S 2 )of<br />

(9.1) is the pr<strong>in</strong>cipal symbol map (see [B5]). Once we have <strong>in</strong>troduced the<br />

exact sequence above, we can state the theorem which is the analogue, for<br />

central states on the dual of SU(2), of the Ney-Spitzer theorem.<br />

Theorem 9.1. The C ∗ algebra generated by the image of the Mart<strong>in</strong> kernel<br />

is the algebra M. The Mart<strong>in</strong> kernel yields a section K : C(S 2 ) →M.<br />

The proof of the theorem relies on a detailed analysis of the Clebsch-Gordan<br />

formulas, see [B5].<br />

Recently the problem of the Mart<strong>in</strong> or Poisson boundary have been considered<br />

for quantum groups, see [Col], [I], [INT]<br />

10 Central Limit Theorems for the Bernoulli Random<br />

Walk<br />

Just as <strong>in</strong> the classical case there exists central limit theorems for the<br />

Bernoulli random walk, however the noncommutativity here plays an important<br />

role, and accord<strong>in</strong>g to whether the state we chose is central or not<br />

the limit is quite different.<br />

10.1 The Case of a Central State<br />

We consider the triple of processes (Xn,Yn,Zn)n≥1 constructed <strong>in</strong> section<br />

(3.1). We use the tracial state to contruct the product, M2(C) ⊗∞ which is<br />

thus endowed with the tracial state σ =( 1<br />

2Tr)⊗∞ . We renormalize the three<br />

processes accord<strong>in</strong>g to<br />

X (λ)<br />

t<br />

= X [λt]<br />

√ , Y<br />

λ (λ)<br />

t = Y [λt]<br />

√λ , Z (λ)<br />

t = Z [λt]<br />

√λ<br />

where [x] is the <strong>in</strong>teger part of x. This triple of processes converges when<br />

λ → ∞, <strong>in</strong> the sense of moments, towards a three dimensional Bownian<br />

motion.<br />

Theorem 10.1. Let (Xt,Yt,Zt)t≥0 be a three dimensional Brownian motion,<br />

then for any polynomial <strong>in</strong> 3n noncommut<strong>in</strong>g <strong>in</strong>determ<strong>in</strong>ates P ,andalltimes<br />

t1,...,tn, one has


Introduction to Random Walks on Noncommutative Spaces 95<br />

limλ→∞ σ(P (X (λ)<br />

t1 ,Y(λ) t1 ,Z (λ)<br />

t1 ,...,X(λ) tn ,Y(λ) tn ,Z(λ) tn<br />

E[P (Xt1,Yt1,Zt1,...,Xtn,Ytn,Ztn)]<br />

)) =<br />

Let us sketch the proof of this result. It is enough to prove the result for<br />

monomials. First we see that for any real coefficients x, y, z the process<br />

xX (λ) (λ)<br />

t + yY t + zZ(λ) t converges towards real Brownian motion. By polarization<br />

this implies that for any monomial <strong>in</strong> X, Y, Z the sum over all<br />

monomials with the same total partial degrees <strong>in</strong> X, Y, Z converges towards<br />

the required limit. Consider two monomials of the form M1X (λ)<br />

t Y (λ)<br />

s M2 and<br />

M1Y (λ)<br />

s X (λ)<br />

t M2, their difference is, thanks to the commutation relations,<br />

M1Z (λ)<br />

s∧tM2/ √ λ. This is a monomial of smaller degree, with a factor 1/ √ λ.We<br />

conclude, by <strong>in</strong>duction on degrees of monomials, that the difference betwen<br />

the expectations of the two monomials M1X (λ)<br />

t Y (λ)<br />

s M2 and M1Y (λ)<br />

s X (λ)<br />

t M2<br />

converges to 0 as λ →∞. It follows that the expectations of all monomials<br />

with the same partial degrees <strong>in</strong> variables converge to the same limit. ⊓⊔<br />

We observe that the sp<strong>in</strong> process, normalized by S [λt]/ √ λ converges <strong>in</strong><br />

distribution to a three dimensional Bessel process as λ →∞.<br />

10.2 The Case of a Pure State<br />

Now we consider the quantum Bernoulli random walk with the pure state<br />

given by the vector e0. We shall consider the convergence of the moments of<br />

the triple of processes<br />

(X (λ)<br />

t ,Y (λ)<br />

t<br />

,Z (λ)<br />

t )=( X [λt]<br />

√ ,<br />

λ Y [λt]<br />

√λ , Z [λt]<br />

); t ≥ 0.<br />

λ<br />

We shall prove that there exists operators (Xt,Yt,Zt)t≥0 on some Hilbert<br />

space H with a vector Ω ∈ H, such that for every polynomial <strong>in</strong> noncommut<strong>in</strong>g<br />

<strong>in</strong>determ<strong>in</strong>ates P (Xt1,Yt1,Zt1,...,Xtn,Ytn,Ztn) one has<br />

limn→∞〈P (X (λ)<br />

t1 ,Y(λ) t1 ,Z (λ)<br />

t1 ,...,X(λ) tn ,Y(λ) tn ,Z(λ) tn )e∞0 ,e∞ 0 〉) =<br />

〈P (Xt1,Yt1,Zt1,...,Xtn,Ytn,Ztn)Ω,Ω〉<br />

For this we shall first <strong>in</strong>vestigate the case where n =1,t1 =1,thuswehave<br />

just three operators ( Xn<br />

√ , n Yn<br />

√ , n Zn<br />

n )andletn→∞.LetHbe a Hilbert space<br />

with a countable orthonormal basis ε0,...,εn,..., and let us def<strong>in</strong>e operators,<br />

with doma<strong>in</strong> the algebraic sum ⊕n i=0Cεi, by the formula<br />

a + (εi) = √ j +1εj+1<br />

a − (εj) = √ jεi−1<br />

(10.1)


96 P. Biane<br />

Theorem 10.2. For any polynomial <strong>in</strong> three noncommutative <strong>in</strong>determ<strong>in</strong>ates<br />

one has<br />

Xn<br />

lim 〈P ( √n ,<br />

n→∞ Yn<br />

√n , Zn<br />

n )e⊗n 0 ,e⊗n 0 〉 = 〈P (a+ + a − , 1<br />

i (a+ − a − ),I)ε0,ε0〉<br />

The proof is immediate by <strong>in</strong>spection of the formula (3.2) and comparison<br />

with (10.1). ⊓⊔<br />

Observe that one has the adjo<strong>in</strong>tness relations<br />

〈a + u, v〉 = 〈u, a − v〉<br />

for all u, v <strong>in</strong> the doma<strong>in</strong>. It follows that the operators a + +a − and 1<br />

i (a+ −a − )<br />

are unbounded symmetric operators on H, and thus are closable. We will see<br />

below that they have self-adjo<strong>in</strong>t extensions. They satisfy the commutation<br />

relation<br />

[a + ,a − ]=−I<br />

on their common doma<strong>in</strong>, spanned by the vectors εi.<br />

The operators a + ,a − thus obta<strong>in</strong>ed are well known under the name of<br />

creation and annihilation operators for the quantum harmonic oscillator. One<br />

can give a natural model for these operators us<strong>in</strong>g a gaussian random variable.<br />

For this, remark that the distribution of the operator a + + a − is gaussian.<br />

This follows easily from the fact that each Xn follows a standard b<strong>in</strong>omial<br />

distribution, and the convergence of this b<strong>in</strong>omial distribution to the gaussian<br />

distribution, by the de Moivre-Laplace theorem. This property of the operator<br />

a + + a − , and the fact that ε0 is a cyclic vector for a + + a − , i.e. the vectors<br />

(a + + a − ) n ε0 span a dense subspace of H, allows us to identify the space<br />

H <strong>in</strong> a natural way with the L 2 space of a gaussian random variable. The<br />

vectors εn can be obta<strong>in</strong>ed from the vectors (a + +a − ) n by the Gram-Schmidt<br />

orthogonalization procedure. It is well known that, for a gaussian variable X,<br />

the polynomials obta<strong>in</strong>ed by the Gram-Schmidt orthonormalization process<br />

from the sequence X n are the Hermite polynomials. Thus when we identify<br />

the space H with the L 2 space of the gaussian measure on R the vectors εn<br />

become identified with the Hermite polynomials. Then the operator a − is<br />

identified with d<br />

dx and the operator a+ with x − d<br />

The product a + a − has eigenvalues 0, 1, 2,... correspond<strong>in</strong>g to the the respective<br />

eigenvectors ε0,ε1,.... It is known as the number operator <strong>in</strong> quantum<br />

field theory.<br />

We will <strong>in</strong> the next section systematize the construction above and show<br />

how to deduce the limit of the renormalized quantum random walk to a<br />

quantum Brownian motion.<br />

dx .


Introduction to Random Walks on Noncommutative Spaces 97<br />

10.3 Fock Spaces<br />

Let H be a complex Hilbert space. For each <strong>in</strong>teger n the symmetric group<br />

Sn acts on H ⊗n by permutation of the factors <strong>in</strong> the tensor product.<br />

Def<strong>in</strong>ition 10.3. We denote H ◦n the subspace of H ⊗n formed by vectors<br />

<strong>in</strong>variant under the action of Sn.<br />

If h1,...,hn ∈ H then we let<br />

h1 ◦ ...◦ hn = 1<br />

√ n!<br />

�<br />

σ∈Sn<br />

h σ(1) ⊗ ...⊗ h σ(n)<br />

which is a multiple of the orthogonal projection of h1 ⊗ ...⊗ hn on H◦n .One<br />

has<br />

〈h1 ...◦ hn,h ′ 1 ◦ ...◦ h′ � n�<br />

n 〉 = 〈hi,h ′ σ(i) 〉<br />

σ∈Sn i=1<br />

The Fock space built on H is the Hilbert space direct sum<br />

Γ (H) =⊕ ∞ n=0H ◦n<br />

where H ◦0 is a one dimensional Hilbert space spanned by a unit vector<br />

Ω, called the vacuum vector of the Fock space. The algebraic direct sum<br />

⊕ ∞ alg,n=0 H◦n , denoted by Γ0(H), is a dense subspace of Γ (H).<br />

For every h ∈ H we def<strong>in</strong>e the exponential vector associated with h by<br />

one has<br />

ξ(h) =<br />

∞�<br />

n=0<br />

h ◦n<br />

n!<br />

〈ξ(h),ξ(h ′ )〉 = e 〈h,k〉<br />

furthermore the vectors ξ(h); h ∈ H form a l<strong>in</strong>early free subset, generat<strong>in</strong>g<br />

algebraically a dense subspace of Γ (H).<br />

If the space H is written as the orthogonal direct sum of two Hilbert<br />

subspaces H = H1 ⊕ H2, then there is a canonical isomorphism<br />

Γ (H) ∼ Γ (H1) ⊗ Γ (H2) (10.2)<br />

which can be obta<strong>in</strong>ed, for example, by identify<strong>in</strong>g the exponential vector<br />

ξ(v1+v2) ∈ Γ (H), where v1 ∈ H1 and v2 ∈ H2, with the vector ξ(v1)⊗ξ(v2) ∈<br />

Γ (H1) ⊗ Γ (H2).<br />

Let h ∈ H, we def<strong>in</strong>e two operators on the doma<strong>in</strong> Γ0(H) by<br />

a +<br />

h (h1 ◦ ...◦ hn) =h ◦ h1 ◦ ...◦ hn<br />

a −<br />

h (h1 ◦ ...◦ hn) = � n<br />

i=1 〈hi,h〉h1 ◦ ...◦ ˆ hi ◦ ...◦ hn


98 P. Biane<br />

h are called creation operators while the a−<br />

h are<br />

called annihilation operators. When H is one dimensional, generated by a<br />

unit vector u there is a natural identification of Γ (H) with the Hilbert space<br />

of the preced<strong>in</strong>g section where the vector u◦n is identified with √ n!ɛn. Then<br />

the operators a + u and a−u co<strong>in</strong>cide with a+ and a− .<br />

Another operator of <strong>in</strong>terest on Γ (H) is the number operator Λ which has<br />

eigenvalue n on the subspace H◦n .If(ei)i∈Iis an orthonormal basis <strong>in</strong> H,<br />

then the number operator Λ has the expansion<br />

Operators of the form a +<br />

Λ = �<br />

i<br />

a + e1 a− ei .<br />

One can see that the creation and annihilation operators satisfy the adjo<strong>in</strong>t<br />

relation<br />

〈a +<br />

h u, v〉 = 〈u, a−<br />

h v〉 h, u, v ∈ Γ0(H)<br />

as well as the commutation relation<br />

[a +<br />

h ,a− k ]=−〈h, k〉I<br />

on the doma<strong>in</strong> Γ0(H). In particular they are closable, and it is easy to see<br />

that the exponential vectors belong to the doma<strong>in</strong> of their closure, with<br />

a +<br />

h ξ(h′ )= d<br />

dt ξ(h′ + th)t=0 a −<br />

h ξ(h′ )=〈h ′ ,h〉ξ(h) (10.3)<br />

We shall see that the real part of the creation operator Ph = a +<br />

h +a− h has a selfadjo<strong>in</strong>t<br />

extension, as well as its imag<strong>in</strong>ary part Qh. For this we consider the<br />

follow<strong>in</strong>g Weyl operators, given on the vector space generated by exponential<br />

vectors by the formula<br />

It is easy to check that<br />

1<br />

−〈h,u〉−<br />

Wuξ(h) =ξ(h + u)e 2 〈u,u〉<br />

〈Wuξ(h),Wuξ(h ′ )〉 = 〈ξ(h),ξ(h ′ )〉 (10.4)<br />

for all u, v, h, h ′ ∈ H, therefore the operators Wu extend to unitary operators<br />

on Γ (H), furthermore<br />

WuWv = Wu+ve −iℑ〈u,v〉<br />

(10.5)<br />

and for any u ∈ H the operators (Witu; t ∈ R) form a one parameter group<br />

of unitary operators, whose generator is given by Pu on exponential vectors.<br />

Similarly, the operators (Wtu; t ∈ R) form a one parameter group of unitary<br />

operators, whose generator is given by Qu, and more generally for θ ∈ [0, 2π[<br />

the vectors (W e iθ tu; t ∈ R) form a one parameter group of unitary operators,<br />

whose generator is given by cos θQu +s<strong>in</strong>θPu.


Introduction to Random Walks on Noncommutative Spaces 99<br />

It follows from Stone’s theorem that Pu, Qu, and all their l<strong>in</strong>ear comb<strong>in</strong>ations,<br />

have a self adjo<strong>in</strong>t extension. These operators satisfy the commutation<br />

relation<br />

[Pu,Qv] =−2iℜ〈u, v〉.<br />

We observe that if H splits as an orthogonal direct sum H = H1 ⊕ H2 and<br />

u ∈ H1, then the operator Wu admits a decomposition Wu = Wu ⊗ I <strong>in</strong> the<br />

decomposition (10.2).<br />

Let now K ⊂ H be a real Hilbert subspace such that ℑ〈u, v〉 =0for<br />

all u, v ∈ K (if K is maximal, it is called a Lagrangian subspace), then by<br />

(10.5) the unitary operators Wiu; u ∈ K form a commutative family, and the<br />

generators Pu,u ∈ K of the one parameter subgroups (Witu; t ∈ R) form<br />

a commut<strong>in</strong>g family of self-adjo<strong>in</strong>t operators with common dense doma<strong>in</strong><br />

Γ0(H). We can therefore <strong>in</strong>vestigate the jo<strong>in</strong>t distribution of these operators.<br />

Proposition 10.4. The operators Pu, foru ∈ K form a gaussian family with<br />

covariance 〈Pu,Pv〉 = 〈u, v〉.<br />

The operators Qu, for u ∈ K form a gaussian family with covariance<br />

〈Qu,Qv〉 = 〈u, v〉.<br />

For the proof it is enough to prove that any l<strong>in</strong>ear comb<strong>in</strong>ation of these<br />

operators has a gaussian distribution with the right variance, i.e. that Pu is<br />

a gaussian with variance 〈u, u〉. For this one computes the Fourier transform<br />

〈e iPu Ω,Ω〉 = 〈e iPu 1 −<br />

ξ(0),ξ(0)〉 = 〈ξ(u)e 2 〈u,u〉 1 −<br />

,ξ(0)〉 = e 2 〈u,u〉<br />

The proof for the operators Qu is similar. ⊓⊔<br />

We let now H = L 2 (R+), and take as Lagrangian subspace the subspace of<br />

real valued functions. Then the family (Pt := P1 [0,t] ; t ≥ 0) has the covariance<br />

〈Pt,Ps〉 = s ∧ t, and thus has the distribution of a real brownian motion. The<br />

same is true of the operators Qt := 1<br />

i (a+ 1 [0,t] − a − 1 [0,t] ) which form another<br />

Brownian motion satisfy<strong>in</strong>g the commutation relations<br />

[Pt,Qt] =2it<br />

We shall call the pair (Pt,Qt)t≥0 of cont<strong>in</strong>uous time processes a noncommutative<br />

brownian motion.<br />

We can now state the limit result we had <strong>in</strong> view <strong>in</strong> the beg<strong>in</strong>n<strong>in</strong>g of this<br />

section.<br />

Theorem 10.5. For any polynomial P <strong>in</strong> noncommut<strong>in</strong>g variables, one has<br />

limλ→∞〈(P (X (λ)<br />

t1 ,Y(λ) t1 ,Z (λ)<br />

t1 ,...,X(λ) tn ,Y(λ) tn ,Z(λ) tn )e∞0 ,e∞0 〉)=<br />

〈P (Pt1,Qt1,t1.I,...,Ptn,Qtn,tn.I))Ω,Ω〉]<br />

The proof is an elaboration of the proof we gave for one time.


100 P. Biane<br />

We shall see <strong>in</strong> the next section that, aga<strong>in</strong>, one can <strong>in</strong>terpret this pair of<br />

processes as a noncommutative Markov process with values <strong>in</strong> a noncommutative<br />

space.<br />

The noncommutative Brownian motion is the basis of a theory of noncommutative<br />

stochastic <strong>in</strong>tegration which has been developped by Hudson and<br />

Parthasarathy, see e.g. [Pa].<br />

11 The Heisenberg Group and the Noncommutative<br />

Brownian Motion<br />

The Heisenberg group is the set H = C × R endowed with the group law<br />

(z,w) ⋆ (z ′ ,w ′ )=(z + z ′ ,w+ w ′ + ℑ(z¯z ′ ))<br />

This is a nilpotent group, its center be<strong>in</strong>g {0}×R, and the Lebesgue measure<br />

on C × R is a left and right Haar measure for this group.<br />

The Weyl operators def<strong>in</strong>ed <strong>in</strong> the preced<strong>in</strong>g section on a Fock space Γ (C)<br />

def<strong>in</strong>e unitary representations of the group H, bysett<strong>in</strong>g,forτ ∈ R ∗ ,<br />

if τ>0and<br />

iτ w<br />

ρτ (z,w) =Wzτ1/2e iτ w<br />

ρτ(z,w) =W¯z|τ | 1/2e<br />

if τ


Introduction to Random Walks on Noncommutative Spaces 101<br />

The "quantum plane"<br />

[P,Q]=it<br />

T=t<br />

Fig. 5 The dual of the Heisenberg group.<br />

The plane T=0<br />

As <strong>in</strong> the case of SU(2) one can give a heuristic description of the noncommutative<br />

space dual to H us<strong>in</strong>g the generators of the Lie algebra of H,<br />

which def<strong>in</strong>e three noncommut<strong>in</strong>g unbounded self-adjo<strong>in</strong>t operators P, Q, T .<br />

We th<strong>in</strong>k of these operators as coord<strong>in</strong>ate functions on this dual space, satisfy<strong>in</strong>g<br />

the commutation relations<br />

[P, Q] =−2iT [P, T ]=[Q, T ]=0.<br />

S<strong>in</strong>ce the coord<strong>in</strong>ate T belongs to the center, it allows to decompose the<br />

space <strong>in</strong>to slices accord<strong>in</strong>g to the values of this coord<strong>in</strong>ate. When T =0,<br />

the coord<strong>in</strong>ates P and Q commute, and the correspond<strong>in</strong>g slice is a usual<br />

plane, with two real coord<strong>in</strong>ates. This corresponds to the one dimensional<br />

representations of the group. When T = τ a non zero real number, the<br />

two coord<strong>in</strong>ates P, Q generate a von Neumann algebra isomorphic to B(H),<br />

and correspond<strong>in</strong>g to the irreducible representation send<strong>in</strong>g T to τI.Note<br />

that <strong>in</strong> this representation the operator P 2 + Q 2 has a discrete spectrum<br />

2|τ|, 6|τ|, 10|τ|,....<br />

Let us consider, for t ≥ 0, the functions on H<br />

ϕ ± 1 ±itw−<br />

t (z,w) =〈ρ±t(z,w)Ω,Ω〉 = e 2 tz¯z .<br />

By construction, these functions are positive def<strong>in</strong>ite functions on H, and<br />

form two multiplicative semigroups. To these functions correspond convolution<br />

semigroups of states, and semigroups of completely positive maps.<br />

T


102 P. Biane<br />

The semigroup of noncommutative brownian motion on the dual of H is<br />

the associated semigroup of completely positive maps on C ∗ (H). Recall<br />

that this semigroup is obta<strong>in</strong>ed by compos<strong>in</strong>g the coproduct ∆ : C ∗ (H) →<br />

M(C ∗ (H)) ⊗ M(C ∗ (H)) (recall that M(A) is the multiplier algebra of A)<br />

with the state associated with the function. Thus<br />

or equivalently<br />

Φ ± t =(ν ϕ ±<br />

t<br />

⊗ I) ◦ ∆<br />

Φ ± t (λg) =ϕ ± t (g)λg for g ∈H.<br />

Let ν be a state on C∗ (H), and ρν : C∗ (H) → B(Hν) be the associated<br />

GNS representation. We consider the two families of homomorphisms<br />

j ± t : C ∗ (H) → B(Hν ⊗ Γ (L 2 (R+)))<br />

j ± t (z,w) =ρν(z,w) ⊗ Wz1 [0,t] e ±itw<br />

We shall prove that these homomorphisms constitute dilations of some completely<br />

positive convolution semigroups on C ∗ (H). For each time t ≥ 0we<br />

have a direct sum decomposition L 2 (R+) =L 2 ([0,t]) ⊕ L 2 ([t, +∞[) and a<br />

correspond<strong>in</strong>g factorization<br />

Γ (L 2 (R+)) = Γ (L 2 ([0,t])) ⊗ Γ (L 2 ([t, +∞[))<br />

Accord<strong>in</strong>gly for each t>0 there are subalgebras<br />

and l<strong>in</strong>ear maps<br />

Wt = B(Γ (L 2 ([0,t]))) ⊗ I ⊂ B(Γ (L 2 (R+)))<br />

Et := Id ⊗〈.Ω [t,Ω [t〉 : B(Γ (L 2 (R+))) → Wt<br />

where Ω [t is the vacuum vector of the space Γ (L 2 ([0,t])).<br />

Lemma 11.1. For each t ≥ 0 the map Et is a conditional expectation with<br />

respect to the state 〈.Ω, Ω〉.<br />

Indeed if a ∈ B(L 2 (R+)) has a decomposition a = a t] ⊗ a [t then one has<br />

and for b, c ∈ B(L 2 ([0,t]))<br />

Et(a) =a t] ⊗〈a [tΩ [t,Ω [t〉<br />

〈bacE(u), E(v)〉 = 〈bEt(a)cE(u1 [0,t]), E(v1 1[0,t])〉〈a [tE(u1 [t,+∞[),E(v1 [t,+∞[)〉<br />

= 〈bEt(a)cE(u), E(v)〉<br />

One checks easily that the homomorphism j ± t sends C ∗ (H) toWt, furthermore<br />

we have


Introduction to Random Walks on Noncommutative Spaces 103<br />

Proposition 11.2. The maps (j ± t ,Et,Wt) form a dilation of the completely<br />

positive semigroup Φ ± t , with <strong>in</strong>itial distribution ν.<br />

Proof. This is a bookkeep<strong>in</strong>g exercise us<strong>in</strong>g the def<strong>in</strong>ition of the Weyl operators,<br />

one has to check that, for t, s ≥ 0andu, v ∈ L 2 (R+), one has<br />

〈j ± 1 ±isw−<br />

t+s(z,w)E(u1 [0,t]), E(v1 [0,t])〉 = e 2 s|z|2<br />

〈j ± t (z,w)E(u1 [0,t]), E(v1 [0,t])〉<br />

⊓⊔<br />

For every one parameter subgroup of H, there is a completely positive<br />

semigroup given by restriction of (Φ ± t )t≥0. For subgroups of the form<br />

(xz, 0); u ∈ R, with z ∈ C ∗ isomorphic to R, one sees that the semigroup is<br />

that of Brownian motion on the dual group. Thus we recover the Brownian<br />

motions (Pt,Qt) of the preced<strong>in</strong>g section by look<strong>in</strong>g at jt(P )andjt(Q).<br />

The restriction to the center (0,w); w ∈ R gives a semigroup of uniform<br />

translation on the real l<strong>in</strong>e.<br />

11.1 The Quantum Bessel Process<br />

Bessel processes are radial parts of Brownian motions. Here we shall exploit<br />

the action of the unitary group U(1) on the Heisenberg group <strong>in</strong> order to f<strong>in</strong>d<br />

an abelian algebra which is left <strong>in</strong>variant by the semigroup and study the<br />

Markov process associated with the restriction of Φ ± t<br />

to this subalgebra. Let<br />

e iθ be a complex number with modulus 1, then there exists an automorphism<br />

aθ of the Heisenberg group def<strong>in</strong>ed by<br />

aθ(z,w) =(e iθ z,w)<br />

and this automorphism extends to an automorphism of the C ∗ algebra.<br />

Proposition 11.3. The subalgebra of C ∗ (H) composed of elements <strong>in</strong>variant<br />

under the above action of U(1) is an abelian C ∗ algebra.<br />

The characters of this abelian C∗-algebra have been computed by A. Koranyi,<br />

they are given by the formula<br />

�<br />

χ(f) = ω(g)f(g)dg<br />

H<br />

for f ∈ L 1 (H), <strong>in</strong>variant under the action of U(1), where ω belongs to the<br />

set of functions<br />

{ωτ,m|τ ∈ R ∗ ,m∈ N}∪{ωµ|µ ∈ R+}


104 P. Biane<br />

Fig. 6 The Heisenberg fan.<br />

with<br />

1 iτ w− ωτ,m(z,w) =me 2 |τ ||z|2Lm(|τ||z|<br />

2 )<br />

ωµ(z,w) =j0(µ|z| 2 )<br />

Here Lm are the Laguerre polynomials def<strong>in</strong>ed by the generat<strong>in</strong>g series<br />

∞�<br />

Lm(x)t m =<br />

m=0<br />

e− xt<br />

1−t<br />

1 − t<br />

and j0 is the usual Bessel function.<br />

The spectrum of the algebra C∗ R (H) can be identified with a closed subset<br />

of R2 , which consists <strong>in</strong> the union of the halfl<strong>in</strong>es {(x, kx); x>0} for k ∈ N,<br />

the halfl<strong>in</strong>es (x, kx); x


Introduction to Random Walks on Noncommutative Spaces 105<br />

Fig. 7<br />

If x =(σ, −kσ) with σ0 and τ = σ + t<br />

qt(x, dy) =<br />

�l=k<br />

l=0<br />

(l + k − 1)! τ<br />

(1 −<br />

(k − 1)!l! t )l+k ( τ<br />

t )kδ (τ,lτ)(dy)<br />

t )k−1<br />

t −<br />

e lt δt,lt(dy)<br />

(l − 1)! σ<br />

(1 −<br />

(k − 1)!(l − k)! τ )k−l ( σ<br />

τ )lδ (τ,lτ)(dy)<br />

The computations can be found <strong>in</strong> [B6].<br />

A typical trajectory of the process is depicted <strong>in</strong> the above picture. It<br />

starts from a po<strong>in</strong>t (σ, −σ) withσ


106 P. Biane<br />

12 Dilations for Noncompact Groups<br />

12.1 The General Case<br />

We shall now extend the preced<strong>in</strong>g construction to the case of arbitrary locally<br />

compact groups. Let G be a locally compact group, with right Haar<br />

measure m, and consider its convolution algebra L 1 (G). We endow this algebra<br />

with the norm �f� =sup�ρ(f)� where the supremum is over all unitary<br />

representations ρ of the group G, extended to the algebra L 1 (G). This yields<br />

the full C ∗ -algebra of G, denotedC ∗ (G). When the group G is amenable, for<br />

example if G is compact or for the Heisenberg group which we have met before,<br />

this co<strong>in</strong>cides with the completion of the action of the group on L 2 (G),<br />

which is called the reduced C ∗ -algebra of G. When the group is nonamenable,<br />

C ∗ r (G) is strictly smaller than C∗ (G). A cont<strong>in</strong>uous positive def<strong>in</strong>ite function<br />

on G, such that ϕ(e) = 1 def<strong>in</strong>es a state as well as a completely positive<br />

contraction on C ∗ (G), whose restriction to L 1 (G) isgivenbyf ↦→ fϕ.Let<br />

now ψ be a cont<strong>in</strong>uous, conditionally negative def<strong>in</strong>ite function on G, with<br />

ψ(e) = 0. Recall that this means that for all t ≥ 0 the function e −tψ is a<br />

positive def<strong>in</strong>ite function on G, or equivalently, by Schönberg’s theorem, that<br />

for all z1,...,zn ∈ C with �<br />

i zi =0,andg1,...,gn ∈ G one has<br />

�<br />

ij<br />

zi¯zjψ(g −1<br />

j gi) ≤ 0.<br />

There is an associated semigroup of completely positive contractions on<br />

C ∗ (G). We shall now, follow<strong>in</strong>g Parthasarathy and Schmidt [PS], construct<br />

a dilation of this semigroup. Let ν be a state on C ∗ (G) which will be the<br />

<strong>in</strong>itial state. The GNS construction yields a unitary representation π of G on<br />

a Hilbert space Hπ, andη ∈ Hπ such that ν(x) =〈π(x)η, η〉 for x ∈ C ∗ (G).<br />

A variant of the GNS construction associates to the function ψ a unitary<br />

representation of G <strong>in</strong> a Hilbert space Hψ, andacocyclev : G → Hψ for this<br />

representation. Thus v is a cont<strong>in</strong>uous function which satisfies<br />

and<br />

v(gh) =gv(h)+v(g)<br />

〈v(g),v(h)〉 = −ψ(h −1 g)+ψ(g)+ψ(h −1 ) − ψ(e) (12.1)<br />

for all g, h ∈ G. Conversely, any function ψ satisfy<strong>in</strong>g the above equation for<br />

some representation and cocycle v is conditionally negative def<strong>in</strong>ite. Indeed<br />

one has �<br />

zi¯zjψ(g −1<br />

j gi) =−� �<br />

zjv(gj)� 2 ≤ 0<br />

if �<br />

j zj =0.<br />

ij<br />

j


Introduction to Random Walks on Noncommutative Spaces 107<br />

Let Γ = Γ (L 2 (R+) ⊗ Hψ), and let W = B(Hπ ⊗ Γ ) ∼ B(Hπ) ⊗ B(Γ ).<br />

Let ω be the pure state on W associated with the vector η ⊗ Ω. Onecan<br />

def<strong>in</strong>e the subalgebras Wt = B(Hπ ⊗ Γt) ⊗ Id associated with the orthogonal<br />

decomposition L 2 (R+) ⊗ Hψ =(L 2 ([0,t]) ⊗ Hψ) ⊕ (L 2 ([t, +∞[) ⊗ Hψ), and<br />

the conditional expectations Et : W→Wt with respect to the state ω. One<br />

def<strong>in</strong>es a unitary representation of G on exponential vectors by<br />

V t (g)(E(u)) = e tψ(g)+〈1 [0,t]⊗vt(g),u〉 E(U t (g)(u)+1[0,t] ⊗ vt(g))<br />

One thus gets a representation of G on Hπ ⊗ Γ by tak<strong>in</strong>g the tensor product<br />

of V t with the representation π, and this yields a family of morphisms jt :<br />

C ∗ (G) →W.<br />

Proposition 12.1. The family (jt, W, Wt,Et,ω) forms a dilation of the completely<br />

positive semigroup, with <strong>in</strong>itial distribution ν.<br />

The proof is a bookkeep<strong>in</strong>g exercise. This construction has been extended<br />

by Schürmann to a class of bialgebras, see [Sc], allow<strong>in</strong>g him <strong>in</strong> particular to<br />

give a nice construction of the Azéma mart<strong>in</strong>gales.<br />

12.2 Free Groups<br />

Let now Fn be a free group on n generators g1,...,gn. Each element of Fn<br />

can be written <strong>in</strong> a unique way as a reduced word w = g ε1<br />

i1 ...gεk<br />

ik ,where<br />

one has εj = ±1 for all j and i1 �= i2 �= i3 ...ik−1 �= ik. For such an element<br />

one def<strong>in</strong>es its length l(w) =k. This is the smallest <strong>in</strong>teger k such that w can<br />

−1<br />

,g2,...,gn,gn }.<br />

be expressed as a product of k elements <strong>in</strong> the set {g1,g −1<br />

1<br />

Proposition 12.2. (Haagerup [H]) The function l is conditionally negative<br />

def<strong>in</strong>ite on Fn.<br />

Proof. Consider the Cayley graph of Fn built on the generators. Thus this<br />

graph has as vertices the elements of Fn and its edges are the pairs {g, h}<br />

such that h −1 g is a generator or the <strong>in</strong>verse of a generator. This Cayley graph<br />

is a regular tree <strong>in</strong> which each vertex has 2n neighbours. For any g ∈ Fn one<br />

can consider the unique shortest path <strong>in</strong> the graph between Id and g. LetEn<br />

be the set of edges of the Cayley graph, endowed with the count<strong>in</strong>g measure,<br />

then one def<strong>in</strong>es v(g) ∈ L 2 (En) to be the <strong>in</strong>dicator function of this shortest<br />

path from Id to g <strong>in</strong> the Cayley graph. Thus v(g)(e) = 1 if and only if the<br />

edge e is on the shortest path from Id to g. One can easily check, us<strong>in</strong>g the<br />

properties of trees that for any h, h ∈ Fn one has<br />

l(g)+l(h) − l(h −1 g)=2〈v(g),v(h)〉 L 2 (Fn)


108 P. Biane<br />

<strong>in</strong>deed this scalar product counts the number of common edges <strong>in</strong> the shortest<br />

paths from Id to g and h. This implies, by (12.1) that l is conditionally<br />

negative def<strong>in</strong>ite. ⊓⊔<br />

It follows from the previous proposition that there exists, on the full C ∗<br />

algebra of Fn, a semigroup of unit preserv<strong>in</strong>g completely positive maps, given<br />

by the formula<br />

Φt(λg) =e −tl(g) λg<br />

The theory of the previous section allows us to construct a dilation of this<br />

semigroup. As before we shall be <strong>in</strong>terested <strong>in</strong> the restriction of this completely<br />

positive semigroup to commutative subalgebras. The first one will be<br />

the subalgebra of the subgroup generated by one of the generators. Let gi<br />

be this generator, then this subgroup is isomorphic to Z by k ↦→ gk i , therefore<br />

its dual is isomorphic to the group of complex numbers of modulus 1. The<br />

restriction gives us a Markov semigroup on the group of complex numbers of<br />

modulus 1. This semigroup is easy to characterize, it sends the function e ikθ<br />

on the unit circle to the function e −|k|+ikθ . In other words this is the <strong>in</strong>tegral<br />

operator on the unit circle given by the Poisson kernel<br />

Pt(θ, θ ′ )=<br />

1 − e−2t 1 − 2e−t cos(θ − θ ′ .<br />

)+e−2t This is a convolution semigroup, as expected.<br />

The other commutative algebra of <strong>in</strong>terest is the algebra R(Fn) consist<strong>in</strong>g<br />

of radial elements. It is generated by the elements χl = �<br />

l(g)=l λg for<br />

l =0, 1,..., and it is immediate that the completely positive semigroup associated<br />

with the length function leaves this algebra <strong>in</strong>variant. Actually one<br />

has Φt(χl) =e−tlχl. These elements satisfy the relations<br />

χ0 = I<br />

χ 2 1 = χ2 +2nχ0<br />

χlχl = χl+1 +(2n − 1)χl−1<br />

l ≥ 2<br />

From this we conclude that R(Fn) is the commutative von Neumann algebra<br />

generated by the self-adjo<strong>in</strong>t element χ1, and its spectrum is the spectrum of<br />

χ1. In order to compute the norm of χl we need just to consider the trivial<br />

representation of Fn <strong>in</strong> which all gi are mapped to the identity, and we get<br />

�χl� =2n(2n − 1) l−1 for l ≥ 1,thenumberofelementsoflengthl <strong>in</strong> Fn.<br />

Any character ϕ : R(Fn) → C is determ<strong>in</strong>ed by its values on χ1. Forsucha<br />

character ϕ, withϕ(χ1) =x, one has<br />

ϕ(χ0) =1; ϕ(χ2) =x 2 − 2n; ϕ(χl+1) =xϕ(χl) − (2n − 1)ϕ(χl−1) forl ≥ 2<br />

from which one <strong>in</strong>fers that


Introduction to Random Walks on Noncommutative Spaces 109<br />

ϕ(χl) = λl+1 1<br />

− λl+1 2<br />

λ1 − λ2<br />

− λl−1 1<br />

− λl−1 2<br />

λ1 − λ2<br />

for l ≥ 1 (12.2)<br />

where λ1,λ2 are the two roots of the equation λ 2 − xλ +2n − 1=0.<br />

We verify that the character ϕx def<strong>in</strong>ed by the formula above is real and<br />

bounded if and only if x ∈ [−2n, 2n]. The spectrum of the algebra R(Fn)<br />

thus co<strong>in</strong>cides with the <strong>in</strong>terval [−2n, 2n], and the element χl corresponds to<br />

a polynomial function Pl(χ1) =χl, where the polynomials are determ<strong>in</strong>ed by<br />

the recursion<br />

P0 =1, P1(x) =x, xPl(x) =Pl+1(x)+(2n − 1)Pl−1.<br />

This three term recursion relation is characteristic of a sequence of orthogonal<br />

polynomials. The orthogonaliz<strong>in</strong>g measure is the distribution of χ1 <strong>in</strong> the<br />

noncommutative probability space (A(Fn),δe) whereδeis the pure state,<br />

<strong>in</strong> the left regular representation of Fn, correspond<strong>in</strong>g to the identity. Thus<br />

δe(λg) =1ifg = e and δe(λg) = 0 if not. This measure is known as the<br />

Kesten measure (see [Ke]) and has the density<br />

dmn(x) = 2n<br />

2n − 1 4π<br />

�<br />

4(2n − 1) − x2 4n2 − x2 on the <strong>in</strong>terval [−2 √ 2n − 1, 2 √ 2n − 1]. The discrepancy between the <strong>in</strong>terval<br />

[−2n, 2n] which is the spectrum of χ1 and the support of the measure mn<br />

comes from the fact that Fn is a nonamenable group and therefore some of<br />

its unitary representations are not weakly conta<strong>in</strong>ed <strong>in</strong> the regular representation.<br />

The semigroup of the restriction of (Φt)t≥0 to the subalgebra R(Fn)<br />

sends the polynomial function Pl(x) on the <strong>in</strong>terval [−2n, 2n] to the function<br />

e−tlPl(x). We can compute the transition probabilities pt(x, dy) by f<strong>in</strong>d<strong>in</strong>g<br />

the <strong>in</strong>tegral representation<br />

e −tl �<br />

Pl(x) = pt(x, dy)Pl(y)<br />

[−2n,2n]<br />

for each x ∈ [−2n, 2n].<br />

If x belongs to the support of the Kesten measure, then s<strong>in</strong>ce the polynomials<br />

Pl; l ≥ 0 form an orthogonal basis of the L 2 space of the Kesten<br />

measure, and �Pl� 2 2 =2n(2n − 1)l one obta<strong>in</strong>s pt through the orthogonal<br />

expansion<br />

pt(x, dy) =e −t ∞�<br />

dmn(y)+ e −tl 1<br />

Pl(x)Pl(y)dmn(y)<br />

2n(2n − 1) l−1<br />

l=1<br />

When x is outside this support, then by (12.2) the sequence Pl(x) is unbounded<br />

and has exponential growth of rate ξ = x+√x2−4(2n−1) 2 and the


110 P. Biane<br />

character value e −tl Pl(x) has the asymptotic behaviour e −tl ξ l as l →∞,<br />

and if e −t ξ > 2 √ 2n − 1, then let x(t) = e −t ξ +(2n − 1)e t /ξ. The function<br />

e −tl Pl(x) can be written as is a l<strong>in</strong>ear comb<strong>in</strong>ation of Pl(x(t)), which<br />

picks up the dom<strong>in</strong>ant term as l →∞and the Pl(y) fory <strong>in</strong> the <strong>in</strong>terval<br />

[−2 √ 2n − 1, 2 √ 2n − 1]. More precisely the quantity<br />

Q t l(x) =e −tl Pl(x) − e−t ξ − (2n − 1)e t /ξ<br />

e −t ξ − e t /ξ<br />

decreases exponentially as l →∞, and thus one has<br />

pt(x, dy) =ctδ x(t) +<br />

ξ − (2n − 1)/ξ<br />

Pl(x(t))<br />

ξ − 1/ξ<br />

∞�<br />

e −tl 1<br />

2n(2n − 1) l−1 (Pl(x) − ctPl(x(t)))Pl(y)dmn(y)<br />

l=1<br />

with ct = e−t ξ−(2n−1)e t /ξ<br />

e −t ξ−e t /ξ<br />

ξ−(2n−1)/ξ<br />

ξ−1/ξ .Ifξ(t) < 2 √ 2n − 1 then there is a simi-<br />

lar decomposition, but the term ct is 0.<br />

Thus the process start<strong>in</strong>g form a po<strong>in</strong>t x ∈ [−2n, 2n] \ [−2 √ 2n − 1,<br />

2 √ 2n − 1] performs a translation towards the central <strong>in</strong>terval, and at some<br />

po<strong>in</strong>t jumps <strong>in</strong>to it, and after that performs a certa<strong>in</strong> pure jump process<br />

<strong>in</strong>side the <strong>in</strong>terval [−2 √ 2n − 1, 2 √ 2n − 1] where it rema<strong>in</strong>s forever.<br />

13 Pitman’s Theorem and the Quantum Group SUq(2)<br />

13.1 Pitman’s Theorem<br />

Let (Bt)t≥0 be a real Brownian motion, with B0 =0,andlet<br />

St = sup Bs<br />

0≤s≤t<br />

then Pitman’s theorem states that the stochastic process<br />

Rt =2St − Bt; t ≥ 0<br />

is a three dimensional Bessel process, i.e. is distributed as the norm of a<br />

three dimensional Brownian motion. There is a discrete version of Pitman’s<br />

theorem, actually it is this discrete version that Pitman proved <strong>in</strong> his orig<strong>in</strong>al<br />

paper [Pi]. We start from a symmetric Bernoulli random walk Xn = x1+...+<br />

xn where the xi are i.i.d. with P (xi) =±1 =1/2, and build the processes<br />

Sn =max1≤k≤n Xk, andTn =2Sn − Xn. Pitmanproved<strong>in</strong>[Pi]that(Tn)n≥0<br />

is a Markov cha<strong>in</strong> on the nonegative <strong>in</strong>tegers, with probability transitions


Introduction to Random Walks on Noncommutative Spaces 111<br />

Fig. 8<br />

p(k, k +1)=<br />

k +2<br />

2(k +1)<br />

p(k, k − 1) =<br />

k<br />

2(k +1)<br />

and the theorem about Brownian motion can be obta<strong>in</strong>ed by tak<strong>in</strong>g the<br />

usual approximation of Brownian motion by random walks. We see that this<br />

Markov cha<strong>in</strong> is, up to a shift of 1 <strong>in</strong> the variable, exactly the one that we<br />

obta<strong>in</strong>ed <strong>in</strong> Theorem 3.2 when consider<strong>in</strong>g the sp<strong>in</strong> process. This is not a<br />

co<strong>in</strong>cidence as we shall see, actually we will understand this connection by<br />

<strong>in</strong>troduc<strong>in</strong>g quantum groups <strong>in</strong> the picture. Before that, let us give the proof<br />

of Pitman’s theorem We consider the stochastic process ((Sn,Xn); n ≥ 1),<br />

with values <strong>in</strong> {(s, k) ∈ N × Z | s ≥ k}. It is easy to see that this stochastic<br />

process is a Markov cha<strong>in</strong>, with transition probabilities<br />

p((s, k), (s, k +1))= 1<br />

1<br />

2 , p((s, k), (s, k − 1)) = 2<br />

p((s, s), (s +1,s+1))= 1<br />

2<br />

, p((s, s), (s, s − 1)) = 1<br />

2<br />

for s>k<br />

from which we can deduce the probability transitions of the Markov cha<strong>in</strong><br />

((Tn,Xn); n ≥ 1), with values <strong>in</strong> {(t, k) ∈ N ∗ ×Z | k ∈ (−t, −t+2,...,t−2,t)},<br />

p((t, k), (t − 1,k+1))= 1<br />

1<br />

2 , p((t, k), (t +1,k− 1)) = 2<br />

p((t, t), (t +1,t+1))= 1<br />

2<br />

, p((t, t), (t +1,t− 1)) = 1<br />

2<br />

if t>k (13.1)<br />

The transition probabilities are depicted <strong>in</strong> this picture.<br />

One checks then, by <strong>in</strong>duction on n, that the conditional distribution of<br />

Xn, know<strong>in</strong>g T1,...,Tn, is the uniform distribution on the set {−Tn, −Tn +<br />

2,...,Tn − 2,Tn}. Then it follows that (Tn; n ≥ 0) is a Markov cha<strong>in</strong> with<br />

the right transition probabilities.


112 P. Biane<br />

13.2 A Markov Cha<strong>in</strong> Associated with the Quantum<br />

Bernoulli Random Walk<br />

In section 5.2 we have seen that the quantum Bernoulli random walk gives<br />

rise <strong>in</strong> a natural way to two Markov cha<strong>in</strong>s, one be<strong>in</strong>g the classical Bernoulli<br />

random walk, and the other be<strong>in</strong>g the sp<strong>in</strong> process. These two processes<br />

were obta<strong>in</strong>ed <strong>in</strong> the preced<strong>in</strong>g section as coord<strong>in</strong>ates of a certa<strong>in</strong> twodimensional<br />

Markov cha<strong>in</strong> given by the transition probabilities (13.1). We<br />

can also consider a two-dimensional Markov cha<strong>in</strong> hav<strong>in</strong>g these two processes<br />

as marg<strong>in</strong>als, by consider<strong>in</strong>g the Markov cha<strong>in</strong> of the end of section 5.2.<br />

Recall that this Markov cha<strong>in</strong> was obta<strong>in</strong>ed by restrict<strong>in</strong>g the generator<br />

of the quantum Bernoulli random walk to the commutative subalgebra<br />

P(SU(2)) ⊂A(SU(2)) generated by the center Z(SU(2)) and by a one parameter<br />

subgroup. The spectrum of this algebra can be identified with the set<br />

ˆP = {(r, k) ∈ N × Z | k ∈{−r, −r +2,...,r− 2,r}}<br />

Indeed this algebra is generated by the pair of commut<strong>in</strong>g self-adjo<strong>in</strong>t operators<br />

X, D <strong>in</strong> the sense that it consists <strong>in</strong> bounded functions of the pair<br />

(X, D). The jo<strong>in</strong>t spectrum of these operators can be computed from the<br />

explicit description of the irreducible representations of SU(2), and co<strong>in</strong>cides<br />

with ˆ P . The probability transitions can be obta<strong>in</strong>ed by us<strong>in</strong>g the Clebsch-<br />

Gordan formula, or equivalently by the computation <strong>in</strong> the proof of Lemma<br />

3.3. One f<strong>in</strong>ds<br />

r + k +2<br />

p((r, k), (r +1,k+1))=<br />

2(r +1)<br />

r − k +2<br />

p((r, k), (r +1,k− 1)) =<br />

2(r +1)<br />

r − k<br />

p((r, k), (r − 1,k+1))=<br />

2(r +1)<br />

r + k<br />

p((r, k), (r − 1,k− 1)) =<br />

2(r +1) .<br />

These transition probabilities are on this picture<br />

Thus, although this Markov cha<strong>in</strong> has the same one-dimensional marg<strong>in</strong>al<br />

as the one of the preced<strong>in</strong>g section, they do not co<strong>in</strong>cide. We will see that<br />

<strong>in</strong> order to recover the transitions (13.1) we will have to <strong>in</strong>troduce quantum<br />

groups.<br />

13.3 The Quantum Group SUq(2)<br />

The Hopf algebra A(SU(2)) can be deformed by <strong>in</strong>troduc<strong>in</strong>g a real parameter<br />

q. The algebraic construction proceeds with the <strong>in</strong>troduction of three


Introduction to Random Walks on Noncommutative Spaces 113<br />

Fig. 9<br />

generators t, e, f which are required to satisfy the relations<br />

tet −1 = q 2 e, tft −1 = q −2 f, ef − fe =<br />

and a coproduct which is given by<br />

t − t−1<br />

q − q −1<br />

∆(t) =t ⊗ t, ∆(e) =e ⊗ t −1 +1⊗ e, ∆(f) =f ⊗ 1+t ⊗ f<br />

Formally putt<strong>in</strong>g t = q h and lett<strong>in</strong>g q converge to 1 one f<strong>in</strong>ds <strong>in</strong> the limit the<br />

def<strong>in</strong><strong>in</strong>g relations for the envelopp<strong>in</strong>g algebra of the Lie algebra of SU(2), as<br />

well as the coproduct.<br />

One can prove that the irreducible f<strong>in</strong>ite dimensional representations<br />

of this algebra are deformations of those of SU(2), <strong>in</strong>deed for each <strong>in</strong>teger<br />

r ≥ 0 there exists two representation <strong>in</strong> V +<br />

−<br />

r+1 and Vr+1 , with bases<br />

; k ∈{−r, −r +2,...,r− 2,r}, givenby<br />

v r±<br />

k<br />

tv r±<br />

j<br />

ev r±<br />

j<br />

fv r±<br />

j<br />

= ±qjv r±<br />

j<br />

��r � � �<br />

− j r + j +2<br />

= ±<br />

2 q 2<br />

=<br />

��r � � �<br />

− j +2 r + j<br />

2 2<br />

q<br />

v<br />

q<br />

r±<br />

j+2<br />

v<br />

q<br />

r±<br />

j−2 .<br />

with [n]q = qn −q −n<br />

q−q −1 . Us<strong>in</strong>g the coproduct one can def<strong>in</strong>e the tensor product<br />

of two representations, and this tensor product obeys the same rules as the


114 P. Biane<br />

one for representations of SU(2) i.e. one has<br />

�<br />

V ɛ1 ɛ2<br />

r1+1 ⊗ Vr2+1 =<br />

r=|r2−r1|,|r2−r1|+2,...,r1+r2<br />

V ɛ1ɛ2<br />

r+1 .<br />

We will now restrict our attention to representations of the k<strong>in</strong>d V +<br />

l , and consider<br />

the von Neumann-Hopf algebra A + (SUq(2)) = ⊕r≥0End[V +<br />

r+1 ], which<br />

is isomorphic, as an algebra, to A(SU(2)), but whose coproduct is deformed.<br />

The subalgebra P(SUq(2)) generated by t and by the center rema<strong>in</strong>s unchanged<br />

<strong>in</strong> the deformation. We consider the tracial state 1<br />

2Tr on the twodimensional<br />

component, and consider the associated random walk. As <strong>in</strong> the<br />

case of SU(2), the restriction of the associated Markov transtion operator to<br />

the commutative algebra P(SUq(2)) def<strong>in</strong>es a Markov cha<strong>in</strong> on the spectrum<br />

of this algebra, whose transition probabilities can be computed, us<strong>in</strong>g the<br />

deformed Clebsch Gordan formulas, as <strong>in</strong> Klimyk et Vilenk<strong>in</strong> [KV], formulas<br />

(6) et (9), §14.4.3, to give<br />

p((r, k), (r +1,k+1))=q (r−k)/2<br />

� �<br />

r+k+2<br />

2 q<br />

=<br />

[r +1]q<br />

qr+1 − q−k−1 2(qr+1 − q−r−1 (13.2)<br />

)<br />

�<br />

� r−k+2<br />

2<br />

p((r, k), (r +1,k− 1)) = q −(r+k)/2<br />

2[r +1]q<br />

p((r, k), (r − 1,k+1))=q −(r+k+2)/2<br />

� �<br />

r−k<br />

2 q<br />

2[r +1]q<br />

p((r, k), (r − 1,k− 1)) = q (r−k+2)/2<br />

� �<br />

r+k<br />

2 q<br />

2[r +1]q<br />

q<br />

= q−k+1 − q −r−1<br />

2(q r+1 − q −r−1 )<br />

= q−k−1 − q −r−1<br />

2(q r+1 − q −r−1 )<br />

= qr+1 − q −k+1<br />

2(q r+1 − q −r−1 )<br />

Lett<strong>in</strong>g q tend to 0, one checks that (13.2) converges to (13.1), and thus<br />

we get Pitman’s theorem as an outcome of the q → 0 limit of the quantum<br />

Bernoulli random walk, see [B7] for details.<br />

This observation is at the basis of a vast generalization of Pitman’s theorem,<br />

see e.g. [BBO].<br />

References<br />

[B1] P. Biane, Some properties of quantum Bernoulli random walks. Quantum probability<br />

& related topics, 193–203, QP-PQ, VI, World Sci. Publ., River Edge, NJ,<br />

1991.<br />

[B2] P. Biane, Quantum random walk on the dual of SU(n). Probab. <strong>Theory</strong> Related<br />

Fields 89 (1991), no. 1, 117–129.<br />

[B3] P. Biane, M<strong>in</strong>uscule weights and random walks on lattices. Quantum probability<br />

& related topics, 51–65, QP-PQ, VII, World Sci. Publ., River Edge, NJ, 1992.<br />

[B4] P. Biane, Équation de Choquet-Deny sur le dual d’un groupe compact. Probab.<br />

<strong>Theory</strong> Related Fields 94 (1992), no. 1, 39–51.


Introduction to Random Walks on Noncommutative Spaces 115<br />

[B5] P. Biane, Théorème de Ney-Spitzer sur le dual de SU(2). Trans. Amer. Math. Soc.<br />

345 (1994), no. 1, 179–194.<br />

[B6] P. Biane, Quelques propriétés du mouvement brownien non-commutatif. Hommage<br />

à P. A. Meyer et J. Neveu. Astérisque No. 236 (1996), 73–101.<br />

[B7] P. Biane, Le théorème de Pitman, le groupe quantique SUq(2), et une question de<br />

P. A. Meyer. In memoriam Paul-André Meyer: Sém<strong>in</strong>aire de Probabilités XXXIX,<br />

61–75, Lecture Notes <strong>in</strong> Math., 1874, Spr<strong>in</strong>ger, Berl<strong>in</strong>, 2006.<br />

[BBO] P. Biane, P. Bougerol, N. O’Connell, Littelmann paths and Brownian paths. Duke<br />

Math. J. 130 (2005), no. 1, 127–167.<br />

[BtD] T. Bröcker, T. tom Dieck, Representations of compact Lie groups. Graduate Texts<br />

<strong>in</strong> Mathematics, 98. Spr<strong>in</strong>ger-Verlag, New York, 1985.<br />

[Col] B. Coll<strong>in</strong>s, Mart<strong>in</strong> boundary theory of some quantum random walks. Ann. Inst.<br />

H. Po<strong>in</strong>caré Probab. Statist. 40 (2004), no. 3, 367–384.<br />

[C] A. Connes, Noncommutative geometry. Academic Press, 1995.<br />

[CD] G. Choquet, J. Deny, Sur l’équation de convolution µ = µ ∗ σ. C.R.Acad.Sci.<br />

Paris 250 1960 799–801.<br />

[D1] J. Dixmier Les C ∗ -algèbres et leurs représentations. Gauthier-Villars. Paris, 1969.<br />

[D2] J. Dixmier Les algèbres d’opérateurs dans l’espace hilbertien (algèbres de von<br />

Neumann). Gauthier-Villars, Paris, 1969.<br />

[GW] R. Goodman, N.R. Wallach, Representations and <strong>in</strong>variants of the classical groups.<br />

Encyclopedia of Mathematics and its Applications, 68. Cambridge University<br />

Press, Cambridge, 1998.<br />

[H] U. Haagerup, An example of a nonnuclear C ∗ -algebra, which has the metric approximation<br />

property. Invent. Math. 50 (1978/79), no. 3, 279–293.<br />

[I] M. Izumi, Non-commutative Poisson boundaries and compact quantum group actions.<br />

Adv. Math. 169 (2002), no. 1, 1–57.<br />

[INT] M. Izumi, S. Neshveyev, L. Tuset, Poisson boundary of the dual of SUq(n). Comm.<br />

Math. Phys. 262 (2006), no. 2, 505–531.<br />

[K] M. Kashiwara, Bases cristall<strong>in</strong>es des groupes quantiques. Rédigé par C. Cochet.<br />

Cours Spcialisés, 9. Société Mathmatique de France, Paris (2002).<br />

[Ke] H. Kesten, Symmetric random walks on groups. Trans. Amer. Math. Soc. 92 1959<br />

336–354.<br />

[KKS] J.G. Kemeny, J.L. Snell, A.W. Knapp, Denumerable Markov cha<strong>in</strong>s. Second edition.<br />

With a chapter on Markov random fields, by David Griffeath. Graduate Texts<br />

<strong>in</strong> Mathematics, No. 40. Spr<strong>in</strong>ger-Verlag, New York-Heidelberg-Berl<strong>in</strong>, 1976.<br />

[KV] A.U. Klimyk, N.Ja. Vilenk<strong>in</strong>, Representation of Lie groups and special functions.<br />

Vol. 2. Class I representations, special functions, and <strong>in</strong>tegral transforms. Mathematics<br />

and its Applications (Soviet Series), 74. Kluwer Academic Publishers<br />

Group, Dordrecht (1993).<br />

[NS] P. Ney, F. Spitzer, The Mart<strong>in</strong> boundary for random walk. Trans. Amer. Math.<br />

Soc. 121 1966 116–132.<br />

[Pa] K.R. Parthasarathy, An <strong>in</strong>troduction to quantum stochastic calculus. Monographs<br />

<strong>in</strong> Mathematics, 85. Birkhuser Verlag, Basel, 1992.<br />

[Pi] J.W. Pitman, One-dimensional Brownian motion and the three-dimensional Bessel<br />

process. Advances <strong>in</strong> Appl. <strong>Probability</strong> 7 (1975), no. 3, 511–526.<br />

[PS] K.R. Parthasarathy, K. Schmidt, Positive def<strong>in</strong>ite kernels, cont<strong>in</strong>uous tensor products,<br />

and central limit theorems of probability theory. Lecture Notes <strong>in</strong> Mathematics,<br />

Vol. 272. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>-New York, 1972.<br />

[R] D. Revuz, Markov cha<strong>in</strong>s. Second edition. North-Holland Mathematical Library,<br />

11. North-Holland Publish<strong>in</strong>g Co., Amsterdam, 1984.<br />

[Rie] M.A. Rieffel, Gromov-Hausdorff distance for quantum metric spaces. Matrix algebras<br />

converge to the sphere for quantum Gromov-Hausdorff distance. Mem. Amer.<br />

Math. Soc. 168 (2004), no. 796. American Mathematical Society, Providence, RI,<br />

2004.


116 P. Biane<br />

[S] J.L. Sauvageot, Markov quantum semigroups admit covariant Markov C ∗ -<br />

dilations. Comm. Math. Phys. 106 (1986), no. 1, 91–103.<br />

[Sc] M. Schürmann, White noise on bialgebras. Lecture Notes <strong>in</strong> Mathematics, 1544.<br />

Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 1993.<br />

[T] M. Takesaki, <strong>Theory</strong> of operator algebras. I, II, III. Encyclopaedia of Mathematical<br />

Sciences, 124, 125, 127 Operator Algebras and Non-commutative Geometry, 5, 6,<br />

8. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 2002/2003.<br />

[Z] D.P. ˇ Zelobenko, Compact Lie groups and their representations. Translated from the<br />

Russian by Israel Program for Scientific Translations. Translations of Mathematical<br />

Monographs, Vol. 40. American Mathematical Society, Providence, R.I., 1973.


Interactions between Quantum<br />

<strong>Probability</strong> and Operator<br />

Space <strong>Theory</strong><br />

Quanhua Xu<br />

Abstract We give a brief aspect of <strong>in</strong>teractions between quantum probability<br />

and operator space theory by emphasiz<strong>in</strong>g the usefulness of noncommutative<br />

Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities <strong>in</strong> the latter theory. After a short <strong>in</strong>troduction to<br />

operator spaces, we present various Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities <strong>in</strong> the noncommutative<br />

sett<strong>in</strong>g, <strong>in</strong>clud<strong>in</strong>g those for Rademacher variables, Voiculescu’s<br />

semicircular systems and Shlyakhtenko’s generalized circular systems. As an<br />

illustration of quantum probabilistic methods <strong>in</strong> operator spaces, we prove<br />

Junge’s complete embedd<strong>in</strong>g of Pisier’s OH space <strong>in</strong>to a noncommutative L1,<br />

for which Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equalities for the generalized circular systems are a<br />

key <strong>in</strong>gredient.<br />

1 Introduction<br />

It is well-known that probabilistic methods are important methods <strong>in</strong> Banach<br />

space theory. As operator spaces are quantized Banach spaces, one would naturally<br />

expect that quantum probability should play a non negligible role <strong>in</strong><br />

the young operator space theory. This is <strong>in</strong>deed the case. In fact, quantum<br />

probability and operator space theory are <strong>in</strong>timately related and there exist<br />

many <strong>in</strong>teractions between them. For <strong>in</strong>stance, the recent remarkable development<br />

of noncommutative mart<strong>in</strong>gale <strong>in</strong>equalities is directly <strong>in</strong>fluenced<br />

and motivated by operator space theory. In particular, the establishment<br />

of the noncommutative Doob maximal <strong>in</strong>equality by Junge [J1] is <strong>in</strong>spired<br />

by Pisier’s theory of vector-valued noncommutative Lp-spaces. On the other<br />

hand, the noncommutative Burkholder/Rosenthal <strong>in</strong>equalities can be used<br />

Q. Xu<br />

UFR Sciences et techniques, Université deFranche-Comté, 16, route de Gray,<br />

F-25 030 Besançon Cedex, France<br />

e-mail: quanhua.xu@univ-fcomte.fr<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 117<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


118 Q. Xu<br />

to determ<strong>in</strong>e the l<strong>in</strong>ear structure of symmetric subspaces of noncommutative<br />

Lp-spaces (see [JX]). The recent works of Pisier/Shlyakhtenko [PS] on<br />

the operator space Grothendieck <strong>in</strong>equality and of Junge [J2] on the complete<br />

embedd<strong>in</strong>g of OH <strong>in</strong>to a noncommutative L1 are two more beautiful examples<br />

of illustration of these <strong>in</strong>teractions. Certa<strong>in</strong> Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities<br />

are key <strong>in</strong>gredients <strong>in</strong> both works.<br />

It is clear today that quantum probability is of <strong>in</strong>creas<strong>in</strong>g importance<br />

<strong>in</strong> operator space theory. We will try to conv<strong>in</strong>ce the reader of this <strong>in</strong> this<br />

course by present<strong>in</strong>g a very brief aspect of <strong>in</strong>teractions between the two theories.<br />

Our presentation is around Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities. In consequence,<br />

noncommutative Lp-spaces are at the heart of these lectures.<br />

This course can be divided <strong>in</strong>to three parts. The first one gives a brief<br />

<strong>in</strong>troduction to operator space theory. We start with a short discussion on<br />

completely positive maps between C*-algebras <strong>in</strong> order to prepare the <strong>in</strong>troduction<br />

to operator spaces and completely bounded maps. Our <strong>in</strong>troduction<br />

to operator spaces beg<strong>in</strong>s with concrete operator spaces, i.e., those which<br />

are closed subspaces of B(H) and the Haagerup-Paulsen-Wittstock factorization<br />

theorem for completely bounded maps, which should be understood<br />

together with its predecessor, St<strong>in</strong>espr<strong>in</strong>g’s factorization theorem for completely<br />

positive maps. We then pass to Ruan’s fundamental characterization<br />

of abstract operator spaces. It is Ruan’s theorem that allows us to do basic<br />

operations on operator spaces such that duality, quotient and <strong>in</strong>terpolation.<br />

Meanwhile, some important examples of operator spaces are given, <strong>in</strong>clud<strong>in</strong>g<br />

column space C, rowspaceRandPisier’s operator Hilbert space OH.<br />

This first part ends with an outl<strong>in</strong>e of Pisier’s vector-valued noncommutative<br />

Lp-spaces. We concentrate here only on Schatten classes.<br />

The second part is of quantum probabilistic character. The ma<strong>in</strong> object of<br />

this part is various noncommutative Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equalities. It becomes clear<br />

nowadays that Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities are of paramount importance <strong>in</strong><br />

operator space theory and more generally <strong>in</strong> noncommutative analysis. The<br />

<strong>in</strong>equalities we present are those for Rademacher variables, free group generators,<br />

Voiculescu’s semicircular systems and Shlyakhtenko’s generalized circular<br />

systems. The proofs of these <strong>in</strong>equalities <strong>in</strong> noncommutative Lp-spaces<br />

are often quite technical and tricky. But what will be needed <strong>in</strong> the third<br />

part concerns only the case p = 1, which is often easier. We should po<strong>in</strong>t out<br />

that the von Neumann algebra generated by a generalized circular system<br />

is of type III, which turns our presentation of noncommutative Lp-spaces<br />

somehow more complicated. This is, unfortunately, unavoidable <strong>in</strong> view of<br />

the complete embedd<strong>in</strong>g of OH <strong>in</strong>to a noncommutative L1 presented later.<br />

The short third part is devoted to Junge’s complete embedd<strong>in</strong>g of OH <strong>in</strong>to<br />

a noncommutative L1. We first present OH as a quotient of a subspace of<br />

C ⊕ R (a theorem of Pisier), which is given by the graph of a closed densely<br />

def<strong>in</strong>ed operator on ℓ2. We then prove that any such graph embeds completely<br />

isomorphically <strong>in</strong>to a noncommutative L1-space.


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 119<br />

2 Completely Positive Maps<br />

This section gives a brief discussion on completely positive maps between<br />

C*-algebras. Our references for operator algebras are [KR] and [T]. Let us<br />

fix some notations used throughout this course:<br />

• B(H) denotes the space of all bounded l<strong>in</strong>ear operators on a complex<br />

Hilbert space H.<br />

• A often denotes a C*-algebra; so A can be regarded as a C*-subalgebra of<br />

B(H) forsomeH. A+ denotes the positive cone of A.<br />

• M denotes a von Neumann algebra and M∗ the predual of M.<br />

• Mn denotes the algebra of complex n×n matrices; so Mn can be identified<br />

with B(ℓn 2 ).<br />

• Mn(A) denotes the algebra of n × n matrices with entries <strong>in</strong> A. This<br />

is aga<strong>in</strong> a C*-algebra. If A ⊂ B(H), then Mn(A) is a C*-subalgebra of<br />

Mn(B(H)) � B(ℓn 2 (H)).<br />

• ℓp denotes the ℓp-space of complex sequences (xk) such that ( �<br />

k |xk| p ) 1/p <<br />

∞, 1≤ p ≤∞(with the usual modification for p = ∞). The n-dimensional<br />

version of ℓp is denoted by ℓn p .<br />

Def<strong>in</strong>ition 2.1. Let u : A → B be a l<strong>in</strong>ear map between two C*-algebras.<br />

i) u is called positive if u(A+) ⊂ B+.<br />

ii) u is called completely positive (c.p. for short) if un is positive for every<br />

n ≥ 1, where un =idMn ⊗ u : Mn(A) → Mn(B) is def<strong>in</strong>ed by un((xij)) =<br />

(u(xij)).<br />

Remark 2.2. It is easy to check that a positive map u : A → B is automatically<br />

cont<strong>in</strong>uous. If <strong>in</strong> addition A is unital, then �u� = �u(1)� (see [Pa]).<br />

Examples:<br />

1) Homomorphisms. Every homomorphism π : A → B is c.p.. By homomorphism<br />

we mean a l<strong>in</strong>ear map satisfy<strong>in</strong>g: π(xy) = π(x)π(y) and<br />

π(x ∗ )=π(x) ∗ for all x, y ∈ A. Recall that a homomorphism π is necessarily<br />

contractive, i.e., �u� ≤1. If <strong>in</strong> addition π is <strong>in</strong>jective, then π is<br />

isometric.<br />

2) Multiplications. Let a ∈ A and def<strong>in</strong>e Ca : A → A by Ca(x) =a ∗ xa.<br />

Then Ca is c.p. and �Ca� = �a ∗ a� = �a� 2 . More generally, if H and<br />

K are two Hilbert spaces and a : K → H a bounded operator, then<br />

Ca : B(H) → B(K) def<strong>in</strong>ed by Ca(x) =a ∗ xa, is c.p..<br />

The classical theorem of St<strong>in</strong>espr<strong>in</strong>g states that any c.p. map is the composition<br />

of two maps of the previous types.


120 Q. Xu<br />

Theorem 2.3. (St<strong>in</strong>espr<strong>in</strong>g’s factorization)<br />

Let A be a C*-algebra and B a C*-subalgebra of B(K). Letu : A → B be<br />

c.p.. Then there are a Hilbert space H, a representation π : A → B(H) (i.e.,<br />

a homomorphism) and a bounded operator a : H → K such that<br />

Namely, u = Ca ◦ π.<br />

u(x) =a ∗ π(x)a, ∀ x ∈ A.<br />

We refer to [T] for the proof of this theorem. We deduce immediately the<br />

follow<strong>in</strong>g<br />

Corollary 2.4. If u : A → B is c.p., then<br />

def<br />

�u�cb = sup �un : Mn(A) → Mn(B)� < ∞.<br />

n<br />

Namely, u is completely bounded. Moreover, �u�cb = �u�.<br />

Exercises:<br />

1) Prove that a positive map is automatically cont<strong>in</strong>uous.<br />

2) Let A be a C*-algebra. Prove that x =(xij)i,j ∈ Mn(A) is positive iff x<br />

is a sum of matrices of the form (a∗ i aj)i,j with a1, ..., an ∈ A.<br />

3) Let A and B be C*-algebras with B commutative. Prove that every positive<br />

map ϕ : A → B is automatically c.p..<br />

3 Concrete Operator Spaces and Completely<br />

Bounded Maps<br />

We consider <strong>in</strong> this section closed subspaces of B(H) and completely bounded<br />

maps between them. The analogue for the present sett<strong>in</strong>g of the St<strong>in</strong>espr<strong>in</strong>g<br />

factorization given <strong>in</strong> the previous section is the Haagerup-Paulsen-Wittstock<br />

factorization for completely bounded maps, which is fundamental <strong>in</strong> the theory.<br />

The references for this and next sections are [ER], [Pa] and [P3].<br />

Def<strong>in</strong>ition 3.1. A(concrete) operator space is a closed subspace E of B(H)<br />

for some Hilbert space H.<br />

Let E be a Banach space, and let BE∗ denote the unit ball of the dual E∗<br />

of E. BE∗ becomes a compact topological space when equipped with the w*topology.<br />

Then E ⊂ C(BE ∗) isometrically. More precisely, given x ∈ E def<strong>in</strong>e<br />

�x : BE∗ → C by �x(ξ) =ξ(x). Then x ↦→ �x establishes an isometry from E <strong>in</strong>to<br />

C(BE ∗). But now C(BE ∗) is a commutative C*-algebra, so C(BE ∗) ⊂ B(H)<br />

for some H. In this way, any Banach space is an operator space. However,<br />

accord<strong>in</strong>g to the preced<strong>in</strong>g def<strong>in</strong>ition, an operator space E is given together


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 121<br />

with an embedd<strong>in</strong>g of E <strong>in</strong>to a B(H). More precisely, an operator space is<br />

a pair of a Banach space E and an embedd<strong>in</strong>g of E <strong>in</strong>to some B(H). On<br />

the other hand, B(H) admits a natural matricial structure: Mn(B(H)) =<br />

B(ℓn 2 (H)) for every n ∈ N. This should be reflected <strong>in</strong> E too. In particular,<br />

the “admissible” morphisms <strong>in</strong> the category of operator spaces should respect<br />

this matricial structure.<br />

Matricial Structure. Let E ⊂ B(H) be an operator space. Then E <strong>in</strong>herits<br />

the matricial structure of B(H) by virtue of the embedd<strong>in</strong>g Mn(E) ⊂<br />

Mn(B(H)). More precisely, let Mn(E) bethespaceofn × n matrices with<br />

entries <strong>in</strong> E. ThenMn(E) is equipped with the norm <strong>in</strong>duced by that of<br />

B(ℓn 2 (H)).<br />

Def<strong>in</strong>ition 3.2. Let E ⊂ B(H) andF⊂ B(K) be two operator spaces. Let<br />

u : E → F be a l<strong>in</strong>ear map.<br />

i) u is called completely bounded (c.b. for short) if<br />

�u�cb =sup�un�<br />

< ∞,<br />

n≥1<br />

where un = idMn ⊗ u : Mn(E) → Mn(F ) is def<strong>in</strong>ed by un((xij)) =<br />

(u(xij)). CB(E,F) denotes the space of all c.b. maps from E to F .<br />

ii) u is called a complete isomorphism if u is a c.b. bijection and u −1 is also<br />

c.b..<br />

iii) u is called completely isometric if un is isometric for every n. Ifu is<br />

a bijection and both u and u −1 are completely isometric, u is called a<br />

complete isometry .<br />

Examples:<br />

1) C*-algebras. Let A be a C*-algebra. Then A is a C*-subalgebra of some<br />

B(H); so A is an operator space. The result<strong>in</strong>g matricial structure on<br />

A is called the natural operator space structure of A. In particular, any<br />

von Neumann algebra has its natural operator space structure. Note that<br />

this natural operator space structure of A is <strong>in</strong>dependent, up to complete<br />

isometry, of particular representation of A as C*-subalgebra of B(H) for<br />

a faithful representation is completely isometric.<br />

2) M<strong>in</strong>imal structure. Let E be a Banach space. Then we have the isometric<br />

embedd<strong>in</strong>g E ⊂ C(BE ∗), which turns E an operator space, denoted by<br />

m<strong>in</strong>(E). This operator space structure is called the m<strong>in</strong>imal structure on<br />

E. The adjective “m<strong>in</strong>imal” means that this structure <strong>in</strong>duces the least<br />

matricial norms on Mn(E). Indeed, assume that E itself is an operator<br />

space and consider also the associated m<strong>in</strong>imal operator space m<strong>in</strong>(E).<br />

Then by Proposition 3.3 below, the identity map on E <strong>in</strong>duces a complete<br />

contraction from E to m<strong>in</strong>(E). Namely, for any n and any x ∈ Mn(E)<br />

�x� Mn(m<strong>in</strong>(E)) ≤�x� Mn(E) .


122 Q. Xu<br />

3) Maximal structure. Similarly, we can <strong>in</strong>troduce a maximal operator<br />

space structure on a Banach space E. LetΦbe the family of all pairs<br />

ϕ =(uϕ,Hϕ)withHϕaHilbert space and uϕ : E → B(Hϕ) a contraction.<br />

Let H be the Hilbert space direct sum of the Hϕ:<br />

H = �<br />

Hϕ .<br />

ϕ∈Φ<br />

Then B(H) conta<strong>in</strong>s as C*-subalgebra the C*-algebra direct sum of the<br />

B(Hϕ):<br />

B(H) ⊃ �<br />

B(Hϕ) .<br />

ϕ∈Φ<br />

Def<strong>in</strong>e J : E → B(H) byJ(x) = � uϕ(x) �<br />

ϕ∈Φ ∈⊕ϕ∈ΦB(Hϕ). It is clear<br />

that J is isometric. This yields an isometric embedd<strong>in</strong>g of E <strong>in</strong>to B(H).<br />

Identify<strong>in</strong>g E and J(E) ⊂ B(H), we turn E an operator space. This<br />

operator space structure is called the maximal structure of E and denoted<br />

by max(E). By def<strong>in</strong>ition, one sees that if E itself is an operator space,<br />

then the identity map idE :max(E) → E is completely contractive.<br />

4) Column and row spaces. Let (eij) denote the standard matrix units of<br />

B(ℓ2). Def<strong>in</strong>e<br />

C = span(ei1, i≥ 1) and R = span(e1j, j≥ 1).<br />

Thus we get two operator spaces C ⊂ B(ℓ2) andR⊂B(ℓ2). Note that C<br />

(resp. R) is identified as the first column (resp. row) subspace of B(ℓ2).<br />

The operator space structures of C and R are easily determ<strong>in</strong>ed. Let<br />

x ∈ Mn(C),y ∈ Mn(R) andwrite<br />

x = �<br />

xi ⊗ ei1, y = �<br />

yi ⊗ e1i, xi, yi∈ Mn.<br />

Then<br />

�x� Mn(C) = � � �<br />

i<br />

i<br />

x ∗ �<br />

i xi<br />

i<br />

� 1/2 , �y� Mn(R) = � � �<br />

i<br />

yiy ∗ i<br />

�<br />

�1/2 .<br />

The n-dimensional versions of C and R are denoted by C n and R n ,respectively.<br />

As Banach spaces both C and R are isometric to ℓ2 via the<br />

identification ei1 ∼ e1i ∼ ei, where(ei) denotes the canonical basis of<br />

ℓ2. We will see at the end of this section that they are not completely<br />

isomorphic.<br />

Proposition 3.3. Let E ⊂ B(H) be an operator space and A a commutative<br />

C*-algebra. Then any bounded map u : E → A is automatically c.b. and<br />

�u�cb = �u�.


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 123<br />

Proof. Let us first consider the case where A = C. Letu : E → C be a<br />

cont<strong>in</strong>uous l<strong>in</strong>ear functional with �u� ≤1. For any n ∈ N we must show<br />

that un : Mn(E) → Mn is a contraction. To this end let x ∈ Mn(E), and let<br />

α, β ∈ ℓ n 2 (α and β be<strong>in</strong>g viewed as column matrices). Then<br />

�<br />

� 〈un(x)α, β〉 � � �n<br />

� = �<br />

i,j=1<br />

≤ � � n �<br />

i,j=1<br />

u(xij)αj ¯ � � � �n<br />

βi�<br />

= �u<br />

i,j=1<br />

xijαj ¯ βi<br />

� ��<br />

xijαj ¯ � � �<br />

βi�<br />

= �β∗ xα� ≤�β��x��α�.<br />

Tak<strong>in</strong>g the supremum over all x, α, β <strong>in</strong> the respective unit balls, we deduce<br />

that �un� ≤1. Thus u is a complete contraction.<br />

The general case can be easily reduced to the previous one. Indeed, s<strong>in</strong>ce<br />

A is commutative, we can assume A = C0(Ω) for some locally compact<br />

topological space Ω, whereC0(Ω) denotes the C*-algebra of all cont<strong>in</strong>uous<br />

functions on Ω which tend to zero at <strong>in</strong>f<strong>in</strong>ity. Then Mn(A) =C0(Ω; Mn), the<br />

C*-algebra of cont<strong>in</strong>uous functions from Ω to Mn which vanish at <strong>in</strong>f<strong>in</strong>ity.<br />

The norm of an element y =(yij) ∈ C0(Ω; Mn) isgivenby<br />

�y� =sup<br />

ω∈Ω<br />

Now let u : E → A be bounded. Then<br />

�un� =sup �� � � u(xij)(ω) �� � Mn<br />

�<br />

� � yij(ω) �� � .<br />

Mn<br />

: ω ∈ Ω, x ∈ Mn(E), �x� ≤1 � .<br />

It follows that<br />

�u�cb =sup�δω◦u�cb<br />

,<br />

ω∈Ω<br />

where δω : C0(Ω) → C is the evaluation at ω: δω(f) =f(ω). We are thus<br />

reduced to the one dimensional case. ⊓⊔<br />

Prototypical Examples of c.b. Maps:<br />

1) Homomorphisms between C*-algebras. These maps are c.b. for they<br />

are c.p.. Moreover, they are completely contractive. Note also that an<br />

<strong>in</strong>jective homomorphism is completely isometric.<br />

2) Multiplications by bounded operators. Givena∈B(H) def<strong>in</strong>e<br />

La : B(H) → B(H), x↦→ ax and Ra : B(H) → B(H), x↦→ xa.<br />

Then La and Ra are c.b. and<br />

�La�cb = �Ra�cb = �a�.<br />

� �<br />

Thus if E ⊂ B(H), then La�<br />

and Ra�<br />

are also c.b..<br />

E E


124 Q. Xu<br />

The follow<strong>in</strong>g theorem asserts that any c.b. map is the composition of a<br />

homomorphism, a left multiplication and a right multiplication. This is the<br />

c.b. analogue of St<strong>in</strong>espr<strong>in</strong>g’s factorization. We refer to [ER] and [P3] for the<br />

proof.<br />

Theorem 3.4. (Haagerup-Paulsen-Wittstock factorization)<br />

Let E ⊂ B(H) and F ⊂ B(K) be two operator spaces. Let u : E → F be a c.b.<br />

map. Then there are a Hilbert space � H, a representation π : B(H) → B( � H)<br />

and two bounded operators a, b ∈ B(K, � H) such that<br />

u(x) =b ∗ π(x)a, ∀ x ∈ E.<br />

Namely, u = Lb ∗ ◦ Ra ◦ π � � E . Moreover,<br />

�u�cb =<strong>in</strong>f � �a��b� � ,<br />

where the <strong>in</strong>fimum is taken over all factorizations of u as above.<br />

Corollary 3.5. (Hahn-Banach type extension)<br />

Every c.b. map u : E → B(K) admits a c.b. extension ũ : B(H) → B(K)<br />

such that �ũ�cb = �u�cb.<br />

Proof. Take a factorization u = Lb ∗ ◦ Ra ◦ π � � E such that �u�cb = �a��b�.<br />

Then ũ = Lb ∗ ◦ Ra ◦ π is the desired extension of u. ⊓⊔<br />

Let E ⊂ B(H) be an operator space. Then E <strong>in</strong>herits the order of B(H),<br />

which allows us to def<strong>in</strong>e positive and completely positive maps on E. Thus<br />

amapu : E → B(K) isc.p.ifun sends the positive part of Mn(E) <strong>in</strong>tothat<br />

of Mn(B(K)) for every n.<br />

Corollary 3.6. (Decomposability of c.b. maps)<br />

Every c.b. map u : E → B(K) is decomposable <strong>in</strong> the sense that there are<br />

four c.p. maps uk such that<br />

with �uk�cb ≤�u�cb.<br />

u = u1 − u2 + i(u3 − u4)<br />

Proof. Let u = Lb∗ ◦ Ra ◦ π � � accord<strong>in</strong>g to Theorem 3.4. Then the corollary<br />

E<br />

immediately follows from the polarization identity with<br />

uk = 1<br />

4 Ca+ikb ◦ π � � , E 1 ≤ k ≤ 4,<br />

where Ca is the multiplication by a from the right and by a∗ from the left. ⊓⊔<br />

Usually, we do not dist<strong>in</strong>guish completely isometric operator spaces.<br />

The distance between two operator spaces is measured by the operator space


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 125<br />

analogue of the Banach-Mazur distance <strong>in</strong> the theory of Banach spaces. The<br />

Banach-Mazur distance of two Banach spaces E and F is<br />

d(E,F) =<strong>in</strong>f � �u −1 ��u� : u : E → F is an isomorphism � .<br />

If E and F are isomorphic, d(E,F) < ∞; otherwise,d(E,F)=∞.<br />

Def<strong>in</strong>ition 3.7. Given two operator spaces E and F def<strong>in</strong>e<br />

dcb(E,F) = <strong>in</strong>f � �u −1 �cb �u�cb : u : E → F is a complete isomorphism � .<br />

Pisier [P1] proved that if dim E =dimF = n, thendcb(E,F) ≤ n. We will<br />

see that the upper bound is atta<strong>in</strong>ed for the pair (C n ,R n ). To this end, let<br />

us <strong>in</strong>troduce a general def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 3.8. An operator space E is called homogeneous if every bounded<br />

map on E is c.b. and �u�cb = �u�. E is called Hilbertian if E is isometric to<br />

a Hilbert space.<br />

It is easy to see that m<strong>in</strong>(E) andmax(E) are homogeneous. On the other<br />

hand, C and R are homogeneous Hilbertian operator spaces. We refer to<br />

Chapiter 10 of [P3] for the proofs of the follow<strong>in</strong>g two results.<br />

Proposition 3.9. Let E and F be two n-dimensional Hilbertian homogeneous<br />

operator spaces. Let (e1, ..., en) and (f1, ..., fn) be orthonormal bases of<br />

E and F , respectively. Let u : E → F be the map def<strong>in</strong>ed by u(ei) =fi. Then<br />

dcb(E,F)=�u −1 �cb�u�cb .<br />

Corollary 3.10. dcb(C n ,R n ) = n and dcb(C n , m<strong>in</strong>(ℓ n 2 ))) = √ n. Consequently,<br />

C, R and m<strong>in</strong>(ℓ2) are not completely isomorphic each other.<br />

Exercises:<br />

1) Let E be a Banach space and F an operator space. Prove that any bounded<br />

maps u : F → m<strong>in</strong>(E) andv :max(E) → F are c.b. and �u�cb = �u�,<br />

�v�cb = �v�.<br />

2) Prove that C and R are homogeneous.<br />

3) Prove Corollary 3.10.<br />

4) Let u : C → R. Prove�u�cb = �u�HS, where�u�HS denotes the Hilbert-<br />

Schmidt norm of u regarded as an operator on ℓ2, i.e.,<br />

�u�HS = � ∞ �<br />

�u(ei)� 2�1/2 2 .<br />

i=1


126 Q. Xu<br />

4 Ruan’s Theorem: Abstract Operator Spaces<br />

The def<strong>in</strong>ition of concrete operator spaces presented <strong>in</strong> the previous section<br />

has a major drawback: it does not allow to do basic operations on concrete<br />

operator spaces. For <strong>in</strong>stance, it is not clear at all how to <strong>in</strong>troduce a nice<br />

matricial structure on the Banach dual E∗ of E which reflects the one of E.<br />

This drawback is resorbed <strong>in</strong> Ruan’s def<strong>in</strong>ition of abstract operator spaces.<br />

We have already seen that a concrete operator space E ⊂ B(H) possesses a<br />

natural matricial structure <strong>in</strong>herited from that of B(H): for each n, Mn(E) ⊂<br />

B(ℓn 2 (H)) is aga<strong>in</strong> an operator space, equipped with the norm �·�n <strong>in</strong>duced<br />

by that of B(ℓn 2 (H)). This sequence (�·�n) of matricial norms clearly satisfy<br />

the follow<strong>in</strong>g properties<br />

• (R1): �αxβ�n ≤�α��x�n�β�, ∀ α, β ∈ Mn,x∈ Mn(E),n≥ 1.<br />

• (R2): �x⊕y�n+m ≤ max � �<br />

�x�n, �y�m , ∀ x ∈ Mn(E),y ∈ Mm(E),n,m≥ 1.<br />

Here the product is the usual matrix product. x ⊕ y denotes the (n + m) ×<br />

(n + m)-matrix � �<br />

x 0<br />

.<br />

0 y<br />

(R1) and(R2) above are usually called Ruan’s axioms .<br />

Theorem 4.1. (Ruan’s characterization)<br />

Let E be a vector space. Assume that each Mn(E) is equipped with a norm<br />

�·�n. If these norms �·�n satisfy Ruan’s axioms (R1) and (R2), then there<br />

are a Hilbert space H and a l<strong>in</strong>ear map J : E → B(H) such that<br />

Jn =idMn ⊗ J : Mn(E) → Mn(B(H)) is isometric for every n.<br />

In other words, the sequence (�·�n) comes from the operator space structure<br />

of E given by the embedd<strong>in</strong>g J : E → B(H).<br />

This theorem is proved <strong>in</strong> [R] (see also [ER] for an alternate proof).<br />

Def<strong>in</strong>ition 4.2. An (abstract) operator space is a Banach space E together<br />

with a sequence (�·�n) of norms satisfy<strong>in</strong>g (R1) and(R2) (with�·�1 equal<br />

to the orig<strong>in</strong>al norm of E).<br />

Henceforth, we will drop the adjective “concrete” or “abstract” by say<strong>in</strong>g<br />

only operator spaces. Thus to have an operator space structure on a Banach<br />

space E is to have a sequence of matricial norms verify<strong>in</strong>g Ruan’s axioms. In<br />

the rema<strong>in</strong>der of this section we present some basic operations on operator<br />

spaces. The complex <strong>in</strong>terpolation is postponed, however, to the next one,<br />

where Pisier’s operator Hilbert space will be also <strong>in</strong>troduced.<br />

Spaces of c.b. maps. If E and F are two operator spaces, CB(E,F) denotes<br />

aga<strong>in</strong> the space of all c.b. maps from E to F .ThisisaBanachspace


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 127<br />

equipped with the norm �·�cb. Now we wish to turn CB(E,F) anoperator<br />

space. Let u =(uij) ∈ Mn(CB(E,F)). We view u as a map from E <strong>in</strong>to<br />

Mn(F ) by def<strong>in</strong><strong>in</strong>g u(x) = � uij(x) � for x ∈ E. Then the matricial norm on<br />

Mn(CB(E,F)) is def<strong>in</strong>ed by<br />

Namely, we have the identification<br />

�u�n = � �u : E → Mn(F ) � � cb .<br />

Mn(CB(E,F)) = CB(E,Mn(F )).<br />

It is easy to show that Ruan’s axioms are verified. Thus CB(E,F) becomes<br />

an operator space.<br />

Duality. Specializ<strong>in</strong>g the previous discussion to F = C,weseethatCB(E,C)<br />

is an operator space. However, Proposition 3.3 implies that CB(E,C) =E∗ isometrically. Therefore, E∗ becomes an operator space. The norm of Mn(E ∗ )<br />

is that of CB(E,Mn). This is usually called the standard dual of E. We will<br />

simply say the dual of E s<strong>in</strong>ce only standard duals are used <strong>in</strong> the sequel.<br />

The bidual E∗∗ =(E∗ ) ∗ is an operator space too. Then it is easy to check<br />

that the natural <strong>in</strong>clusion E↩→ E∗∗ is completely isometric. This allows to<br />

view E as a subspace of E∗∗ .<br />

Thus the duals of C*-algebras and the preduals of von Neumann algebras<br />

are operator spaces.<br />

Let u : E → F be a map between two operator spaces. Then u is c.b. iff<br />

its adjo<strong>in</strong>t u∗ : F ∗ → E∗ is c.b.. If this is the case, �u�cb = �u∗�cb. Quotient. Let E be an operator space and F ⊂ E a closed subspace. We<br />

equip Mn(E/F) with the quotient norm of Mn(E)/Mn(F ). Then it is easy to<br />

check that these norms satisfy (R1) and(R2), so E/F becomes an operator<br />

space. The usual duality between subspaces and quotients <strong>in</strong> Banach space<br />

theory rema<strong>in</strong>s available now:<br />

F ∗ = E∗ �E �∗ ⊥<br />

and = F completely isometrically.<br />

F ⊥ F<br />

Direct sum. Let (Ek) be a sequence of operator spaces, Ek ⊂ B(Hk) (the<br />

sequence may be f<strong>in</strong>ite). Let ℓ∞((Ek)) denote the space of all sequences<br />

(xk) withxk ∈ Ek such that sup k �xk� < ∞. This is a Banach space when<br />

equipped with the norm<br />

�(xk)� =sup�xk�.<br />

k<br />

ℓ∞((Ek)) naturally <strong>in</strong>herits a matricial structure from ℓ∞((B(Hk))), the latter<br />

be<strong>in</strong>g a C*-algebra. Thus we have<br />

�<br />

Mn ℓ∞((Ek)) � �<br />

= ℓ∞ (Mn(Ek)) � .


128 Q. Xu<br />

c0((Ek)) denotes the subspace of ℓ∞((Ek)) consist<strong>in</strong>g of all (xk) such that<br />

�xk� →0.<br />

On the other hand, we def<strong>in</strong>e ℓ1((Ek)) to be the space of all sequences<br />

(xk) withxk∈Ek such that �<br />

k �xk� < ∞. This is aga<strong>in</strong> a Banach space<br />

with the natural norm. Recall that<br />

�<br />

ℓ1((Ek)) �∗ = ℓ∞((E ∗ k )) isometrically.<br />

Thus ℓ1((Ek)) is a predual of ℓ∞((Ek)), which allows us to view ℓ1((Ek)) as<br />

an operator space too. We often use the notations<br />

�<br />

∞ Ek and �<br />

1 Ek<br />

k<br />

<strong>in</strong>stead of ℓ∞((Ek)k) andℓ1((Ek)k), respectively. If all Ek are equal, these<br />

spaces are denoted by ℓ∞(E) andℓ1(E), respectively. In particular, if E = C,<br />

we recover ℓ∞ and ℓ1.<br />

Let us <strong>in</strong>troduce the cont<strong>in</strong>uous version of c0(E). Let Ω be a locally compact<br />

topological space and E ⊂ B(H) an operator space. Let C0(Ω; E) denote<br />

the space of cont<strong>in</strong>uous functions from Ω to E which vanish at <strong>in</strong>f<strong>in</strong>ity.<br />

C0(Ω; E) is equipped with the uniform norm<br />

k<br />

�f� =sup�f(ω)�.<br />

ω∈Ω<br />

Then C0(Ω; E) ⊂ C0(Ω; B(H)). The latter space is a C*-algebra. In this way,<br />

we turn C0(Ω; E) an operator space.<br />

Sum and Intersection. Let (E0,E1) be a couple of operator spaces. Assume<br />

(E0,E1) iscompatible <strong>in</strong> the sense that both E0 and E1 cont<strong>in</strong>uously embed<br />

<strong>in</strong>to a common topological vector space V . This allows us to def<strong>in</strong>e<br />

E0 ∩ E1 = � �<br />

x ∈ V : x ∈ E0, x∈ E1<br />

and<br />

E0 + E1 = � x ∈ V<br />

�<br />

: ∃ x0 ∈ E0, x1∈ E1 s.t. x= x0 + x1 ,<br />

equipped respectively with the <strong>in</strong>tersection and sum norms<br />

�x�E0∩E1 =max � �<br />

�x�E0 , �x�E1 ,<br />

�x�E0+E1 =<strong>in</strong>f � �x0�E0 + �x1�E1<br />

�<br />

: x = x0 + x1,x0 ∈ E0, x1∈ E1 .<br />

Note that E0 ∩E1 can be regarded as the diagonal subspace of E0 ⊕∞ E1. On<br />

the other hand, E0 +E1 is identifiable with the quotient space of E0 ⊕1 E1 by<br />

∆, where∆ = {(x0,x1) : x0 + x1 =0}. Equipped with the operator space<br />

structures <strong>in</strong>duced by those of E0 ⊕∞ E1 and E0 ⊕1 E1, respectively, E ∩ E1<br />

and E0 + E1 become operator spaces too.


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 129<br />

Let us consider an important example. Take the column and row spaces<br />

C and R and view them as compatible by identify<strong>in</strong>g both of them with ℓ2<br />

at the Banach space level. Thus if x ∈ C, wewrite<br />

x = �<br />

xiei1 with xi ∈ C.<br />

Recall that<br />

i<br />

�x�C =( �<br />

|xi| 2 ) 1/2 = �(xi)�ℓ2 = � �<br />

i<br />

i<br />

xie1i�R .<br />

Then the compatibility above means that x is identified with the sequence<br />

(xi). Accord<strong>in</strong>gly, we often identify the canonical bases (ei1) ofC and (e1i)<br />

of R with (ei) ofℓ2. Thus as Banach spaces, C ∩ R = C = R (= ℓ2). But this<br />

is no longer true <strong>in</strong> the category of operator spaces.<br />

Let x ∈ Mn(C ∩ R) =Mn(C) ∩ Mn(R). Write<br />

x = �<br />

xi ⊗ ei with xi ∈ Mn.<br />

Then<br />

i<br />

�x�Mn(C∩R) =max � �<br />

�x�Mn(C) , �x�Mn(R) =max � � �<br />

x ∗ i xi� 1/2 , � �<br />

i<br />

i<br />

xix ∗ i �1/2� .<br />

Us<strong>in</strong>g this, we easily check that C ∩ R is not completely isomorphic to C.<br />

Indeed, us<strong>in</strong>g Proposition 3.3, one easily shows that their n-dimensional versions<br />

satisfy the follow<strong>in</strong>g<br />

dcb(C n ,C n ∩ R n )= √ n.<br />

The operator space structure of C + R is a little bit more complicated.<br />

Us<strong>in</strong>g the complete isomorphism between E1⊕∞E2 and E1⊕1E2 (see Exercise<br />

6 below), we deduce that for x = �<br />

i xi ⊗ ei ∈ Mn(C + R)<br />

1<br />

2 �x�Mn(C+R) ≤ <strong>in</strong>f �� � �<br />

i<br />

y ∗ i yi<br />

�<br />

� 1/2 + � � �<br />

i<br />

ziz ∗ i<br />

�<br />

�1/2� ≤ 2�x�Mn(C+R) ,<br />

where the <strong>in</strong>fimum runs over all decompositions xi = yi +zi with yi, zi ∈ Mn.<br />

We will give later a complete isometric description of C + R <strong>in</strong> terms of the<br />

trace class S1.<br />

Exercises:<br />

1) Prove that C ∗ � R and R ∗ � C completely isometrically. More precisely,<br />

the map ξ ∈ C ∗ ↦→ �<br />

i ξ(ei1)e1i establishes a complete isometry between<br />

C ∗ and R.<br />

2) Show that the natural <strong>in</strong>clusion E↩→ E ∗∗ is completely isometric.


130 Q. Xu<br />

3) Let F ⊂ E be operator spaces. Prove<br />

F ∗ = E∗ �E �∗ ⊥<br />

and = F completely isometrically.<br />

F ⊥ F<br />

4) Prove that a map u : E → F is c.b. iff its adjo<strong>in</strong>t u∗ : F ∗ → E∗ is c.b..<br />

Moreover, �u�cb = �u∗�cb. 5) Prove<br />

�<br />

c0((Ek)) �∗ = ℓ1((E ∗ k)) and � ℓ1((Ek)) �∗ = ℓ∞((E ∗ k))<br />

completely isometrically.<br />

6) Prove that the natural <strong>in</strong>clusion from ℓ1((Ek)) <strong>in</strong>to ℓ∞((Ek)) is completely<br />

contractive. Conversely, prove that the c.b. norm of the formal identity<br />

from E1 ⊕∞ ···⊕∞ En to E1 ⊕1 ···⊕1 En is equal to n.<br />

7) Let (E0,E1) be a compatible couple of operator spaces such that E0 ∩ E1<br />

is dense <strong>in</strong> both E0 and E1. Prove<br />

� �∗ ∗<br />

E0∩E1 = E0 +E ∗ 1 ,<br />

� �∗ ∗<br />

E0+E1 = E0 ∩E ∗ 1 completely isometrically.<br />

5 Complex Interpolation and Operator Hilbert Spaces<br />

Hilbert spaces are <strong>in</strong> the center of the family of Banach spaces and play a<br />

crucial role there. It is thus natural to f<strong>in</strong>d their operator space analogues.<br />

One way to def<strong>in</strong>e operator Hilbert spaces is complex <strong>in</strong>terpolation, which is<br />

of great importance for its own right <strong>in</strong> operator space theory. Our references<br />

for this section are [P1] and [P3].<br />

Interpolation. We first recall the def<strong>in</strong>ition of complex <strong>in</strong>terpolation for Banach<br />

spaces. Let (E0,E1) be a compatible couple of complex Banach spaces.<br />

Let S = {z ∈ C : 0 ≤ Re z ≤ 1}, astrip<strong>in</strong>C. LetF(E0,E1) be the family<br />

of all functions f : S → E0 + E1 satisfy<strong>in</strong>g the follow<strong>in</strong>g conditions:<br />

• f is cont<strong>in</strong>uous on S and analytic <strong>in</strong> the <strong>in</strong>terior of S;<br />

• f(k + it) ∈ Ek for all t ∈ R and the function t ↦→ f(k + it) is cont<strong>in</strong>uous<br />

from R to Ek for k =0andk =1;<br />

• lim |t|→∞ �f(k + it)�Ek =0fork =0andk =1.<br />

We equip F(E0,E1) with the norm:<br />

�<br />

�f � � =max<br />

F(E0,E1) � �<br />

sup �f(it)<br />

t∈R<br />

� �<br />

� , sup �f(1 + it)<br />

E0<br />

t∈R<br />

� �<br />

� .<br />

E1<br />

Then it is a rout<strong>in</strong>e exercise to check that F(E0,E1) is a Banach space. For<br />

0


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 131<br />

�x�θ =<strong>in</strong>f �� � f � �F(E0,E1) : f(θ) =x, f ∈F(E0,E1) � ,<br />

Eθ becomes a Banach space. Note that by the maximum pr<strong>in</strong>ciple, the map<br />

f ↦→ f(θ) is a contraction from F(E0,E1) toE0 + E1. Then(E0, E1)θ can<br />

be isometrically identified with the quotient of F(E0,E1) bythekernelof<br />

this map.<br />

Now assume E0 and E1 are operator spaces. Then (Mn(E0), Mn(E1)) is<br />

aga<strong>in</strong> compatible for any n ≥ 1. This allows to def<strong>in</strong>e<br />

Mn(Eθ) = � Mn(E0), Mn(E1) �<br />

θ .<br />

It is easy to show that Ruan’s axioms are satisfied, so Eθ is an operator space.<br />

Let us express Eθ as a quotient of a subspace of C0(R; E0) ⊕∞ C0(R; E1).<br />

Us<strong>in</strong>g Poisson <strong>in</strong>tegral, one sees that C0(R; E0) ⊕∞ C0(R; E1) is just the<br />

space of functions f : S → E0+E1 satisfy<strong>in</strong>g the same conditions as above for<br />

F(E0,E1) but only with “analytic” replaced by “harmonic”. Then F(E0,E1)<br />

is the subspace of C0(R; E0)⊕∞ C0(R; E1) consist<strong>in</strong>g of all analytic functions.<br />

Therefore, Eθ is the quotient space of F(E0,E1) by the subspace of all f such<br />

that f(θ) =0.<br />

Operator Hilbert Spaces. Let H be a complex Hilbert space. Fix<strong>in</strong>g an<br />

orthonormal basis (ei)i∈I <strong>in</strong> H, wecanidentifyHwith ℓ2(I). The classical<br />

Riesz representation theorem asserts that H∗ is isometric to the conjugate<br />

¯H of H. The latter space is H itself with the same norm but with conjugate<br />

multiplication: λ · x = ¯ λx for λ ∈ C and x ∈ H. Viewed as a vector <strong>in</strong><br />

¯H, a vector x ∈ H is often denoted by ¯x. Thusifx =(xi) ∈ ℓ2(I), then<br />

¯x =(¯xi)i∈I. Themapx↦→ ¯x establishes an anti-l<strong>in</strong>ear isometry between H<br />

and ¯ H.Consequently,H∗is isometric to H. In fact, the conjugation can be<br />

def<strong>in</strong>ed for any Banach space X. It is a well-known elementary fact that the<br />

Hilbert spaces are only Banach spaces X such that X∗ is isometric to X.<br />

Now we wish to consider the operator space analogue of this property.<br />

The result<strong>in</strong>g spaces are the so-called operator Hilbert spaces, <strong>in</strong>troduced by<br />

Pisier. Note that if E is an operator space, so is Ē by def<strong>in</strong><strong>in</strong>g Mn( Ē)=<br />

Mn(E).<br />

Theorem 5.1. Let I be an <strong>in</strong>dex set. For any x = �<br />

i xi ⊗ ei ∈ Mn(ℓ2(I))<br />

def<strong>in</strong>e<br />

�x�n = � � � �<br />

xi ⊗ ¯xi �1/2 Mn⊗Mn = � � � �<br />

xi ⊗ ¯xi<br />

i<br />

i<br />

� 1/2<br />

M n 2 .<br />

Then (�·�n) satisfy Ruan’s axioms. The result<strong>in</strong>g operator space is denoted<br />

by OH(I). Moreover, OH(I) is the unique, up to complete isometry, operator<br />

space structure on ℓ2(I) such that OH(I) ∗ is completely isometric to OH(I).<br />

If I = N or I = {1, ..., n}, wedenoteOH(I) simplybyOH or OH n .<br />

The space OH can be also obta<strong>in</strong>ed by <strong>in</strong>terpolat<strong>in</strong>g the column and row<br />

spaces.


132 Q. Xu<br />

Theorem 5.2. OH =(C, R) 1/2 completely isometrically.<br />

Exercises:<br />

1) Let (E0,E1) and(F0,F1) be two compatible couples of operator spaces.<br />

Let T : E0 + E1 → F0 + F1 be a l<strong>in</strong>ear map such that T � � : Ek → Fk<br />

Ek<br />

is c.b. of cb-norm ck for k =0, 1. Then T is c.b. from Eθ to Fθ for any<br />

0


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 133<br />

def<strong>in</strong>ition s<strong>in</strong>ce the usual trace Tr on B(ℓ2) is not f<strong>in</strong>ite. However, it still<br />

has a nice trace <strong>in</strong> the sense that it is normal, semif<strong>in</strong>ite and faithful. The<br />

preced<strong>in</strong>g construction can be done for normal semif<strong>in</strong>ite faithful traces too.<br />

The result<strong>in</strong>g spaces for (B(ℓ2), Tr) are the Schatten classes Sp. Recall that<br />

S1 and S2 are respectively the trace and Hilbert-Schmidt classes. Note that<br />

by def<strong>in</strong>ition S∞ is the whole B(ℓ2).<br />

The previous def<strong>in</strong>ition does not apply to type III von Neumann algebras.<br />

There are several equivalent constructions of noncommutative Lp-spaces <strong>in</strong><br />

the type III case. Here we adopt the one by Kosaki [Ko] via complex <strong>in</strong>terpolation.<br />

Let M be a von Neumann algebra equipped with a normal faithful<br />

state ϕ. Def<strong>in</strong>eL∞(M) =M as before and L1(M) =M∗. Consider the left<br />

<strong>in</strong>jection j of L∞(M) <strong>in</strong>toL1(M) byj(x) =xϕ. The faithfulness of ϕ implies<br />

that j is <strong>in</strong>jective and its range is dense <strong>in</strong> L1(M). This <strong>in</strong>jection makes<br />

(L∞(M), L1(M)) compatible. Thus we can consider the complex <strong>in</strong>terpolation<br />

spaces between them. Now for 1


134 Q. Xu<br />

for x =(xij) ∈ Sp and y =(yij) ∈ Sp ′. This is due to our def<strong>in</strong>ition that S1<br />

is def<strong>in</strong>ed as the predual of B(ℓ2) op .<br />

6.3 Vector-valued Schatten Classes. We wish to def<strong>in</strong>e Schatten classes<br />

with values <strong>in</strong> operator spaces. To this end we first recall the m<strong>in</strong>imal tensor<br />

product <strong>in</strong> the category of operator spaces. Let E ⊂ B(H) andF ⊂ B(K) be<br />

two operator spaces. Let x ∈ B(H) andy ∈ B(K). The tensor x ⊗ y is an<br />

operator on the Hilbert space tensor product H ⊗K def<strong>in</strong>ed by x⊗y(ξ ⊗η) =<br />

x(ξ) ⊗ y(η). It is easy to check that x ⊗ y is bounded and �x ⊗ y� = �x��y�.<br />

Consequently, the algebraic tensor product E ⊗ F is a vector subspace of<br />

B(H ⊗ K). Then the m<strong>in</strong>imal tensor product E ⊗m<strong>in</strong> F is def<strong>in</strong>ed to be the<br />

closure of E ⊗ F <strong>in</strong> B(H ⊗ K).<br />

Now let E be an operator space. Def<strong>in</strong>e S∞[E] tobeS∞ ⊗m<strong>in</strong> E. The<br />

elements <strong>in</strong> S∞[E] are often represented as <strong>in</strong>f<strong>in</strong>ite matrices with entries<br />

<strong>in</strong> E.<br />

To def<strong>in</strong>e S1[E] we use duality. The operator space structure to be put <strong>in</strong><br />

S1[E] will be such that the dual space of S1[E] isS∞[E ∗ ]. More precisely, let<br />

u ∈ S1 ⊗ E. Write<br />

u = �<br />

ak ⊗ xk with ak ∈ S1, xk∈ E.<br />

k<br />

Consider u as a l<strong>in</strong>ear functional on S∞[E∗ ] as follows. For v = �<br />

j bj ⊗ ξj ∈<br />

S∞ ⊗ E∗ def<strong>in</strong>e<br />

〈u, v〉 = �<br />

〈ak,bj〉〈xk,ξj〉 = �<br />

Tr(akb t j) ξj(xk).<br />

k,j<br />

Then the norm of u is def<strong>in</strong>ed to be the l<strong>in</strong>ear functional norm of u on S∞[E∗ ],<br />

which co<strong>in</strong>cides with the norm of u <strong>in</strong> CB(S∞[E∗ ], C) (see Proposition 3.3).<br />

We def<strong>in</strong>e S1[E] to be the closure of S1⊗E with respect to this norm. Next, we<br />

have to <strong>in</strong>troduce a norm �·�n on Mn(S1[E]) for any n. Thisisnoweasy.<br />

As before, every u ∈ Mn(S1 ⊗ E) <strong>in</strong>duces a l<strong>in</strong>ear map from S∞[E∗ ]toMn.<br />

Then def<strong>in</strong>e<br />

�u�n = �u�CB(S∞[E∗ ],Mn).<br />

It is then rout<strong>in</strong>e to check that these norms satisfy Ruan’s axioms. Therefore,<br />

S1[E] becomes an operator space.<br />

Hav<strong>in</strong>g def<strong>in</strong>ed S∞[E] andS1[E], we def<strong>in</strong>e Sp[E] by <strong>in</strong>terpolation for any<br />

1


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 135<br />

Theorem 6.1. Let 1 ≤ p


136 Q. Xu<br />

is a natural algebraic identification of Lp(Mn ⊗ M) withMn(Lp(M)). Then<br />

S n p [Lp(M)] is noth<strong>in</strong>g but the l<strong>in</strong>ear space Mn(Lp(M)) equipped with the<br />

norm of Lp(Mn ⊗M). More generally, if E ⊂ Lp(M) is a closed subspace, the<br />

norm of S n p [E] is <strong>in</strong>duced by that of S n p [Lp(M)]. In the <strong>in</strong>f<strong>in</strong>ite dimensional<br />

case, Sp[Lp(M)] is completely isometrically identified with Lp(B(ℓ2) ¯⊗M) for<br />

all 1 ≤ p < ∞. IfE ⊂ Lp(M), then Sp[E] is the closure of Sp ⊗ E <strong>in</strong><br />

Lp(B(ℓ2) ¯⊗M).<br />

6.4 Column and Row p-spaces. The column and row spaces, C and R,<br />

are pillars of the whole theory of operator spaces. Recall that C and R are<br />

respectively the (first) column and row subspaces of S∞. By analogy, let Cp<br />

(resp. Rp) denote the first column (resp. row) subspace of Sp. S<strong>in</strong>ceS2 is an<br />

OH space, C2 � R2 � OH. Then-dimensional versions of these spaces are<br />

denoted by C n p and Rn p .<br />

Now let E be an operator space. We denote by Cp[E] (resp.Rp[E]) the<br />

closure of Cp ⊗ E (resp. Rp ⊗ E) <strong>in</strong>Sp[E]. If E is a subspace of a noncommutative<br />

Lp(M), the norm of Cp[E] is easy to determ<strong>in</strong>e. We consider only<br />

the case where M is semif<strong>in</strong>ite. For any f<strong>in</strong>ite sequence (xk) ⊂ E<br />

�<br />

� � �<br />

xk ⊗ ek<br />

k<br />

� Cp[E] = � � � �<br />

k<br />

x ∗ �1/2� �Lp(M)<br />

kxk<br />

,<br />

where (ek) denotes the canonical basis of Cp. More generally, if ak ∈ Cp, then<br />

�<br />

� � �<br />

xk ⊗ ak�<br />

= Cp[E] � � � �<br />

〈aj,ak〉x ∗ kxj �1/2� �Lp(M) ,<br />

k<br />

j,k<br />

where 〈 , 〉 denotes the scalar product <strong>in</strong> Cp. (In terms of matrix product,<br />

〈aj,ak〉 = a ∗ k aj.) We also have a similar description for Rp[E].<br />

We end this section by describ<strong>in</strong>g the operator space structure of C + R<br />

with help of Corollary 6.2. To this end we use the identification that C � R1<br />

and R � C1 (see Exercise 2 below). Thus C + R � R1 + C1, whichisthe<br />

quotient of R1 ⊕1 C1 by the subspace {(x, y); x + y =0}. Therefore, for any<br />

f<strong>in</strong>ite sequence (xk) ⊂ S1<br />

�<br />

� (xk) � � =<strong>in</strong>f S1[C+R] �� � (yk) � � + S1[R1] � � (zk) � �<br />

�<br />

S1[C1]<br />

=<strong>in</strong>f �� � �<br />

� ( yky ∗�1/2) �<br />

�S1<br />

k + � � � �<br />

� ∗ 1/2) � �<br />

�S1<br />

( zk zk ,<br />

k<br />

where the <strong>in</strong>fimum runs over all decompositions of xk = yk + zk <strong>in</strong> S1. It<br />

follows that S1[C + R] =C1[S1]+R1[S1] with equal norms (even completely<br />

isometrically).<br />

Exercises:<br />

1) Let ℓ1 have its natural operator space structure. Let x ∈ Mn(ℓ1) with<br />

x = �<br />

i xi ⊗ ei (with (ei) the canonical basis of ℓ1). Prove<br />

k


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 137<br />

�x�Mn(ℓ1) =sup �� � � �<br />

xi ⊗ yi�<br />

Mmn : yi ∈ Mm, �yi� ≤1,m∈ N � .<br />

i<br />

On the other hand, show S1[ℓ1] =ℓ1(S1) completely isometrically.<br />

2) Let 1 ≤ p ≤∞and p ′ be the conjugate <strong>in</strong>dex of p. Prove the follow<strong>in</strong>g<br />

completely isometric identities:<br />

(Cp) ∗ ∼ = Cp ′ ∼ = Rp and (Rp) ∗ ∼ = Rp ′ ∼ = Cp<br />

by identify<strong>in</strong>g the canonical bases <strong>in</strong> question.<br />

3) Prove that Cp and Rp are homogeneous Hilbertian spaces.<br />

4) Compute (or estimate) dcb(C n p , C n q )for1≤ p, q ≤∞.Thenprovethat<br />

Cp and Cq are not completely isomorphic for p �= q.<br />

7 Noncommutative Kh<strong>in</strong>tch<strong>in</strong>e Type Inequalities<br />

This section is devoted to Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities <strong>in</strong> the noncommutative<br />

Lp-spaces. These <strong>in</strong>equalities are of paramount importance <strong>in</strong> operator space<br />

theory and noncommutative analysis.<br />

7.1 The <strong>Classical</strong> Kh<strong>in</strong>tch<strong>in</strong>e Inequalities. Let (εk) beaRademacher<br />

sequence on a probability space (Ω,P), i.e., an <strong>in</strong>dependent sequence of random<br />

variables such that P (εk = 1) = P (εk = −1) = 1/2. The classical<br />

Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality states that for any 1 ≤ p


138 Q. Xu<br />

Now we wish to extend (7.2) to the noncommutative sett<strong>in</strong>g, i.e., replac<strong>in</strong>g<br />

Lp(0, 1) by a noncommutative Lp. We will consider only the case where the<br />

coefficients xk are <strong>in</strong> the Schatten classes Sp. All <strong>in</strong>equalities stated below<br />

are valid for general noncommutative Lp. On the other hand, we will concentrate<br />

ma<strong>in</strong>ly on the case p = 1 (and the case p = ∞ <strong>in</strong> the free case). Thus<br />

our goal is to f<strong>in</strong>d a determ<strong>in</strong>istic expression for<br />

�<br />

� � �<br />

xkεk�<br />

.<br />

Lp(Ω;Sp)<br />

In view of the square function � � |xk| 2�1/2 <strong>in</strong> (7.2), we are naturally led to<br />

conjecture that the determ<strong>in</strong>istic expression to be found should <strong>in</strong>volve the<br />

term �� �<br />

� |xk| 2�1/2� � = p � � � � � ∗ 1/2� xk xk<br />

�p .<br />

This norm is also equal to �(xk)�Cp[Sp]. Here and <strong>in</strong> the sequel, we often<br />

identify a sequence (xk) <strong>in</strong>E with the element �<br />

k xk ⊗ ek <strong>in</strong> Cp[E] orRp[E].<br />

But now because of the noncommutativity, we should also take <strong>in</strong>to account<br />

the right modulus, i.e., the term<br />

�<br />

� � �<br />

∗<br />

|xk | 2�1/2� � = p � � � �<br />

xkx ∗�1/2� �p<br />

k = �(xk)�Rp[Sp] .<br />

Although �a�p = �a ∗ �p for a s<strong>in</strong>gle operator a ∈ Sp, thetwotermsaboveare<br />

not comparable at all if p �= 2. For example, if xk = ek1, then clearly<br />

�<br />

� � n �<br />

x ∗ kxk �1/2� �p = n 1/2<br />

k=1<br />

and � � � n �<br />

k=1<br />

xkx ∗�1/2� �p<br />

k = n 1/p .<br />

7.2 Column and Row Subspaces. In this subsection we collect some<br />

basic properties of the column and row subspaces Cp[Sp] andRp[Sp].<br />

Proposition 7.1. i) The Hölder <strong>in</strong>equality: Let 1 ≤ p, q, r ≤∞such that<br />

1/r =1/p +1/q. Then for any sequences (xk) ∈ Cp[Sp] and (yk) ∈ Cq[Sq]<br />

the series �<br />

k y∗ kxk converges <strong>in</strong> Sr (with respect to the w*-topology if<br />

r = ∞) and<br />

�<br />

� �<br />

k<br />

y ∗ kxk �<br />

� ≤<br />

r � �(xk) � �<br />

� �(yk)<br />

Cp[Sp]<br />

� � .<br />

Cq[Sq]<br />

ii) Complementation: Cp[Sp] and Rp[Sp] are 1-complemented subspaces of<br />

Sp(ℓ2 ⊗ ℓ2) for any 1 ≤ p ≤∞. More precisely, let P : Sp(ℓ2 ⊗ ℓ2) →<br />

Sp(ℓ2 ⊗ ℓ2) be def<strong>in</strong>ed by P (x) =xe, wheree =1⊗ e11. ThenP is a<br />

contractive projection from Sp(ℓ2 ⊗ ℓ2) onto Cp[Sp].<br />

iii) Duality: Let 1 ≤ p


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 139<br />

isometrically with respect to the anti-l<strong>in</strong>ear duality bracket:<br />

〈(xk), (yk)〉 ↦→ �<br />

Proof. i) Given (xk) ⊂ Sp def<strong>in</strong>e<br />

k≥1<br />

Tr(y ∗ k xk).<br />

⎛ ⎞<br />

x1 0 ···<br />

⎜<br />

T ((xk)) = x2 ⎝<br />

0 ··· ⎟<br />

⎠ .<br />

. .<br />

We view T ((xk)) as an element <strong>in</strong> Sp[Sp] =Sp(ℓ2 ⊗ ℓ2). Now let (yk) ⊂ Sq<br />

be another f<strong>in</strong>ite sequence. Then<br />

�<br />

k<br />

y ∗ k xk = T ((yk)) ∗ T ((xk)).<br />

So the desired <strong>in</strong>equality follows from the Hölder <strong>in</strong>equality <strong>in</strong> Sp(ℓ2 ⊗ ℓ2).<br />

ii) This is obvious. Note that P (x) is the matrix whose first column is that<br />

of x and all others are zero.<br />

iii) Us<strong>in</strong>g i) one sees that the duality is well-def<strong>in</strong>ed. The two duality<br />

equalities are immediate consequences of ii) and the duality between Sp(ℓ2 ⊗<br />

ℓ2) andSp ′(ℓ2 ⊗ ℓ2). ⊓⊔<br />

Proposition 7.2. Let 1 ≤ p ≤∞and (xk) ⊂ Sp be a f<strong>in</strong>ite sequence.<br />

i) If 2 ≤ p ≤∞,then<br />

max �� � (xk) � � , Cp[Sp] � � (xk) � � � �<br />

� ≤ �xk� Rp[Sp]<br />

2�1/2 p .<br />

ii) If 1 ≤ p ≤ 2, then<br />

m<strong>in</strong> �� �(xk) � � ,<br />

Cp[Sp] � �(xk) � � � �<br />

� ≥ �xk�<br />

Rp[Sp]<br />

2�1/2 p .<br />

Proof. i) Let p ≥ 2. By the triangle <strong>in</strong>equality <strong>in</strong> Sp/2 we have<br />

�<br />

� (xk) � �2 Cp[Sp] = � � �<br />

|xk| 2�� ≤ p/2 �<br />

�xk� 2 p .<br />

k<br />

Pass<strong>in</strong>g to adjo<strong>in</strong>ts we get the <strong>in</strong>equality on the row norm.<br />

ii) This follows from i) by duality. ⊓⊔<br />

Corollary 7.3. i) Let (Σ,µ) be a measure space, and let f belong to the<br />

algebraic tensor product L2(Σ) ⊗ Sp. Thenif2≤p≤∞, max �� �<br />

�<br />

�<br />

Σ<br />

f(t) ∗ f(t)dµ(t) �1/2� � ,<br />

p � �<br />

�<br />

�<br />

Σ<br />

k<br />

k<br />

k<br />

f(t)f(t) ∗ dµ(t) � 1/2 � � p<br />

� ≤ � �f � �L2(Σ;Sp) .


140 Q. Xu<br />

and if 1 ≤ p ≤ 2,<br />

� �<br />

�f�L2(Σ;Sp) ≤ m<strong>in</strong> �� �<br />

�<br />

�<br />

Σ<br />

f(t) ∗ f(t)dµ(t) �1/2� � ,<br />

p � �<br />

�<br />

�<br />

Σ<br />

f(t)f(t) ∗ dµ(t) �1/2� �<br />

� .<br />

p<br />

ii) In particular, if (ϕk) is an orthonormal sequence <strong>in</strong> L2(Σ) and (xk) a<br />

f<strong>in</strong>ite sequence <strong>in</strong> Sp, then for 2 ≤ p ≤∞<br />

max �� � (xk) � � , Cp[Sp] � � (xk) � � � � �<br />

� ≤ � xkϕk�<br />

;<br />

Rp[Sp]<br />

L2(Σ;Sp)<br />

and for 1 ≤ p ≤ 2<br />

�<br />

� � �<br />

xkϕk�<br />

≤ <strong>in</strong>f L2(Σ;Sp) �� � (yk) � � + Cp[Sp] � � (zk) � �<br />

� , Rp[Sp]<br />

k<br />

where the <strong>in</strong>fimum runs over all decompositions xk = yk + zk with yk and<br />

zk <strong>in</strong> Sp.<br />

Proof. i) Note that s<strong>in</strong>ce f ∈ L2(Σ) ⊗ Sp, thetwo<strong>in</strong>tegrals<br />

�<br />

f(t) ∗ �<br />

f(t)dµ(t) and f(t)f(t) ∗ dµ(t)<br />

Σ<br />

are well-def<strong>in</strong>ed and belong to S p/2. The desired <strong>in</strong>equalities follow from the<br />

correspond<strong>in</strong>g ones <strong>in</strong> the previous proposition.<br />

ii) The first <strong>in</strong>equality immediately follows from i) above. To prove the<br />

second take a decomposition xk = yk + zk. Then aga<strong>in</strong> by the previous<br />

proposition<br />

�<br />

� �<br />

k<br />

�<br />

xkϕk<br />

� L2(Σ;Sp) ≤ � � �<br />

k<br />

�<br />

ykϕk<br />

Σ<br />

k<br />

� L2(Σ;Sp) + � � �<br />

≤ � � (yk) � � Cp[Sp] + � � (zk) � � Rp[Sp] ;<br />

k<br />

�<br />

zkϕk�<br />

L2(Σ;Sp)<br />

whence the desired <strong>in</strong>equality. ⊓⊔<br />

Recall that (C, R) is considered as a compatible pair by identify<strong>in</strong>g their<br />

canonical bases with that of ℓ2. This identification is extended to all spaces<br />

Cp and Rp. Thus any pair (Cp,Rq) is also compatible. Then the maximum<br />

and <strong>in</strong>fimum <strong>in</strong> Corollary 7.3, ii) are respectively �(xk)� Cp[Sp]∩Rp[Sp] and<br />

�(xk)� Cp[Sp]+Rp[Sp]. For notational simplicity, we <strong>in</strong>troduce the follow<strong>in</strong>g<br />

Def<strong>in</strong>ition 7.4. For 1 ≤ p ≤∞def<strong>in</strong>e<br />

CRp[Sp] =Cp[Sp]∩Rp[Sp]ifp ≥ 2 and CRp[Sp] =Cp[Sp]+Rp[Sp]ifp


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 141<br />

with equivalent norms; moreover, the equivalence constants are controlled<br />

by a universal one. By Corollary 6.2, these identities are also completely<br />

isomorphic. Thus if we def<strong>in</strong>e CRp = Cp ∩ Rp for p ≥ 2andCRp = Cp + Rp<br />

for p


142 Q. Xu<br />

Then<br />

and<br />

g 2 = �<br />

x 2 k + �<br />

εjεk(xjxk + xkxj)<br />

k<br />

E(g 4 )= � �<br />

k<br />

x 2 k<br />

≤ 1+2 �<br />

j


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 143<br />

Cont<strong>in</strong>u<strong>in</strong>g this procedure we obta<strong>in</strong> a sequence � f (n)� of functions <strong>in</strong><br />

L∞(Ω; S∞) and a sequence � x (n) �<br />

k <strong>in</strong> S∞ such that<br />

xk = �<br />

2 −n x (n)<br />

k , �f (n) �L∞(Ω;S∞) ≤ 2 √ 3 , x (n)<br />

k = E(f (n) εk) .<br />

Put<br />

n≥0<br />

f = �<br />

n≥0<br />

2 −n f (n) .<br />

Then<br />

�f�L∞(Ω;S∞) ≤ 4 √ 3 and xk = E(fεk) .<br />

Thus the statement (∗) is proved for selfadjo<strong>in</strong>t xk. The general case is easily<br />

reduced to the selfadjo<strong>in</strong>t one by decompos<strong>in</strong>g each xk <strong>in</strong>to its real and imag<strong>in</strong>ary<br />

parts. Note that the f<strong>in</strong>al constant c obta<strong>in</strong>ed <strong>in</strong> this way is 8 √ 3. We<br />

refer to [HM] for a more careful argument which yields √ 3 as constant. ⊓⊔<br />

Remark 7.6. Let Rp be the closed subspace of Lp(Ω) generated by the εk.<br />

Us<strong>in</strong>g Corollary 6.2, we can rephrase Theorem 7.5 as the fact that Rp is<br />

completely isomorphic to CRp. This remark also applies to Theorems 7.8<br />

and 7.9 below.<br />

Remark 7.7. The Rademacher sequence (εk) can be replaced by a standard<br />

Gaussian sequence. More generally, let (ϕk) be an <strong>in</strong>dependent sequence of<br />

random variables on (Ω,P). Assume that (ϕk) is symmetric <strong>in</strong> the sense that<br />

(ϕk) has the same distribution as (±ϕk) for any sequence of signs. Assume<br />

further that<br />

0 < <strong>in</strong>f �ϕk�p, sup �ϕk�2 < ∞ for 1 ≤ p


144 Q. Xu<br />

and zero elsewhere. Let λ : G → B(ℓ2(G)) be the left regular representation.<br />

Namely, for any g ∈ G, λ(g) is the unitary operator on ℓ2(G) def<strong>in</strong>ed by<br />

� λ(g)ϕ � (h) =ϕ(g −1 h), h,g ∈ G, ϕ ∈ ℓ2(G).<br />

Note that λ(g)δh = δgh for all g, h ∈ G. Then the group von Neumann<br />

algebra VN(G) is the von Neumann subalgebra of B(ℓ2(G)) generated by<br />

{λ(g) : g ∈ G}.Namely,VN(G) is the w*-closure of all f<strong>in</strong>ite sums � αgλ(g)<br />

with αg ∈ C. VN(G) also co<strong>in</strong>cides with the left convolution algebra of ℓ2(G).<br />

Recall that if ϕ, ψ ∈ ℓ2(G), their convolution is def<strong>in</strong>ed by<br />

ϕ ∗ ψ(g) = �<br />

ϕ(h)ψ(h −1 g), g ∈ G.<br />

h∈G<br />

Then x ∈ VN(G) iffthereexistsϕ ∈ ℓ2(G) such that xψ = ϕ ∗ ψ for every<br />

ψ ∈ ℓ2(G). Let τG be the vector state on VN(G) determ<strong>in</strong>ed by δe, i.e.,<br />

τG(x) =〈xδe, δe〉 for any x ∈ VN(G). Then it is easy to check that τG is<br />

faithful and tracial.<br />

If we identify an operator x ∈ VN(G) with its symbol xδe <strong>in</strong> ℓ2(G),<br />

then L2(VN(G)) is noth<strong>in</strong>g but ℓ2(G). L1(VN(G)) is traditionally called the<br />

Fourier algebra of G and denoted by A(G). S<strong>in</strong>ce an operator <strong>in</strong> L1(VN(G))<br />

is a product of two operators <strong>in</strong> L2(VN(G)), a function ϕ on G belongs to<br />

A(G) iff there exist two functions ψ, ρ ∈ ℓ2(G) such that ϕ = ψρ. We refer<br />

to [KR] for more <strong>in</strong>formation.<br />

If G is abelian, VN(G) ieequaltoL∞( � G), so is commutative, where � G is<br />

the dual group of G.<br />

An important example of non abelian groups is a free group F on n generators<br />

(gk) with n ∈ N ∪{∞}. The sequence (λ(gk))k of the unitary operators<br />

given by the generators is of particular <strong>in</strong>terest. With a slight abuse of term<strong>in</strong>ology,<br />

we will also call it a sequence of free generators .Notethat(λ(gk))k<br />

is orthonormal <strong>in</strong> L2(VN(F)). Let us determ<strong>in</strong>e its l<strong>in</strong>ear span <strong>in</strong> VN(F).<br />

To this end let Fk be the closed subspace of ℓ2(F) generated by all those<br />

basic vectors δg for which g is a reduced word start<strong>in</strong>g with g −1<br />

k .TheFk,<br />

k =1, 2, ..., are mutually orthogonal. Let Fk be the orthogonal projection<br />

from ℓ2(F) ontoFk. Now, given a f<strong>in</strong>ite sequence (αk) ⊂ C write<br />

�<br />

αkλ(gk) = �<br />

αkλ(gk)Fk + �<br />

k<br />

k<br />

k<br />

αkλ(gk)F ⊥ k .<br />

By the mutual orthogonality of the Fk, wehave<br />

�<br />

� � �<br />

αkλ(gk)Fk �2 ∞ = � �<br />

� ( αkλ(gk)Fk)( �<br />

αkλ(gk)Fk) ∗�� ∞<br />

k<br />

k<br />

= � � �<br />

|αk| 2 λ(gk)Fkλ(gk) ∗�� ≤<br />

∞ �<br />

|αk| 2 .<br />

k<br />

k<br />

k


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 145<br />

To treat another term, observe that for k �= j the range of λ(gk) ∗λ(gj)F ⊥ j is<br />

conta<strong>in</strong>ed <strong>in</strong> Fk, soF⊥ k λ(gk) ∗λ(gj)F ⊥ j = 0. Then it follows that<br />

�<br />

� �<br />

Therefore,<br />

k<br />

αkλ(gk)F ⊥ k<br />

�<br />

�2 ∞ = � �( �<br />

k<br />

= � � �<br />

k<br />

αkλ(gk)F ⊥ k )∗ ( �<br />

|αk| 2 F ⊥ k<br />

k<br />

�<br />

� ≤ ∞ �<br />

|αk| 2 .<br />

�<br />

� �<br />

αkλ(gk) � � ≤ 2 ∞ � �<br />

|αk| 2�1/2 .<br />

k<br />

k<br />

k<br />

αkλ(gk)F ⊥ k )� � ∞<br />

The converse <strong>in</strong>equality with constant 1 is obvious for<br />

�<br />

� �<br />

αkλ(gk) � � ≥ ∞ � � �<br />

αkλ(gk) � � = 2 � �<br />

|αk| 2�1/2 .<br />

k<br />

k<br />

Consequently, for any 1 ≤ p ≤∞the closed subspace generated by the λ(gk)<br />

<strong>in</strong> Lp(VN(F)) is isomorphic to ℓ2. More precisely, for any f<strong>in</strong>ite sequence<br />

(αk) ⊂ C,<br />

�<br />

� �<br />

αkλ(gk) � � �<br />

� ∼c |αk| p 2�1/2 ,<br />

k<br />

where c is a universal constant. This <strong>in</strong>equality rema<strong>in</strong>s true if the scalar coefficients<br />

(αk) are replaced by operator coefficients. The result<strong>in</strong>g <strong>in</strong>equalities<br />

are the Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality for free generators . In the follow<strong>in</strong>g statement<br />

Lp(B(ℓ2) ¯⊗VN(F)) is the noncommutative Lp-space associated with<br />

the von Neumann tensor product B(ℓ2) ¯⊗VN(F), equipped with the tensor<br />

trace Tr ⊗ τF, which is a normal semif<strong>in</strong>ite faithful trace. Note that<br />

Lp(B(ℓ2) ¯⊗VN(F)) = Sp[Lp(VN(F))].<br />

The follow<strong>in</strong>g theorem is due to Haagerup/Pisier [HP] for p = ∞, 1andto<br />

Pisier [P2] for 1


146 Q. Xu<br />

�<br />

� (xk) � � ≤ CR∞[S∞] � � �<br />

xk ⊗ λ(gk) � �<br />

B(ℓ2) ¯⊗VN(F) ≤ 2�� (xk) � �<br />

CR∞[S∞] (7.4)<br />

k<br />

for any f<strong>in</strong>ite sequence (xk) ⊂ S∞. The proof of the second <strong>in</strong>equality above<br />

is the same as <strong>in</strong> the scalar case. Let us show the first one. Take a unit vector<br />

ξ ∈ ℓ2. Then<br />

�<br />

� �<br />

xk ⊗ λ(gk)<br />

k<br />

� �2 ∞ ≥〈�xk<br />

⊗ λ(gk)(ξ ⊗ δe),<br />

k<br />

�<br />

xk ⊗ λ(gk)(ξ ⊗ δe)〉<br />

k<br />

= 〈 �<br />

�<br />

xk(ξ) ⊗ δgk , xk(ξ) ⊗ δgk<br />

k<br />

k<br />

〉<br />

= �<br />

〈xk(ξ), xk(ξ)〉 = 〈 �<br />

x ∗ kxk(ξ), ξ〉;<br />

whence � � � �<br />

Tak<strong>in</strong>g adjo<strong>in</strong>ts, we f<strong>in</strong>d<br />

�<br />

� � �<br />

k<br />

k<br />

k<br />

x ∗ kxk<br />

xkx ∗ k<br />

�1/2� �∞ ≤ � � �<br />

xk ⊗ λ(gk) � � . ∞<br />

k<br />

�1/2� �∞ ≤ � � �<br />

xk ⊗ λ(gk) � � . ∞<br />

Therefore, the lower estimate of (7.4) is proved.<br />

Dualiz<strong>in</strong>g (7.4), we get the case p =1:<br />

1 �<br />

� (xk)<br />

2<br />

� � ≤ CR1[S1] � � �<br />

xk ⊗ λ(gk) � �<br />

L1(B(ℓ2) ¯⊗V N(F)) ≤ � � (xk) � � (7.5)<br />

CR1[S1]<br />

k<br />

for any f<strong>in</strong>ite sequence (xk) ⊂ S1. Indeed, let (yk) ⊂ S∞ be such that<br />

�<br />

� (xk) � � = CR1[S1] �<br />

Tr(y ∗ kxk) and � � (yk) � � ≤ 1<br />

CR∞[S∞]<br />

k<br />

(see Proposition 7.1 iii)). Set<br />

x = �<br />

xk⊗λ(gk) ∈ L1(B(ℓ2) ¯⊗VN(F)), y = �<br />

yk⊗λ(gk) ∈ B(ℓ2) ¯⊗VN(F).<br />

k<br />

Then by (7.4)<br />

�<br />

k<br />

Tr(y ∗ kxk) =Tr⊗ τF(y ∗ x) ≤�y�∞ �x�1 ≤ 2�(yk)� CR∞[S∞] �x�1 ≤ 2�x�1 .<br />

This is the lower estimate of (7.5). To show the upper estimate we consider<br />

the projection P from the algebra of all f<strong>in</strong>ite sums � αgλ(g) withαg ∈ C<br />

onto the l<strong>in</strong>ear span of the λ(gk). Let Q = I−P . It is easy to see that P (x)and<br />

Q(x) are orthogonal relative to the scalar product of L2(VN(F)). Thus the<br />

k<br />

k<br />

k


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 147<br />

extension of P on L2(VN(F)) is the orthogonal projection from L2(VN(F))<br />

onto the closed subspace generated by the λ(gk). Now let x = �<br />

g∈F xg ⊗λ(g)<br />

be a f<strong>in</strong>ite sum with xg ∈ S∞. Write<br />

x =idS∞⊗P (x)+idS∞⊗ Q(x) = �<br />

xk ⊗ λ(gk)+idS∞⊗ Q(x).<br />

Then us<strong>in</strong>g the argument yield<strong>in</strong>g the first <strong>in</strong>equality of (7.4), we get<br />

�<br />

� (xk) � � CR∞[S∞] ≤�x�∞ . (7.6)<br />

Together with a duality argument as above, this <strong>in</strong>equality implies the upper<br />

estimate of (7.5).<br />

Comb<strong>in</strong><strong>in</strong>g (7.4) and (7.6), we get<br />

On the other hand, note that<br />

�P (x)�∞ ≤ 2�x�∞ .<br />

Tr ⊗ τF(y ∗ P (x)) = Tr ⊗ τF(P (y) ∗ x)<br />

for all f<strong>in</strong>ite sums x = �<br />

g∈F xg ⊗ λ(g) andy = �<br />

g∈F yg ⊗ λ(g) with xg ∈ S∞<br />

and yg ∈ S1. We then deduce that P extends to a completely bounded map<br />

P1 on L1(VN(F)) (with �P1�cb ≤ 2), which is a pre-adjo<strong>in</strong>t of P . This implies<br />

that P extends to a completely bounded normal projection on VN(F). Note<br />

that P1 co<strong>in</strong>cides with P on the family of f<strong>in</strong>ite sums as above, which allows<br />

us to denote P1 still by P . F<strong>in</strong>ally, by <strong>in</strong>terpolation P extends to a complete<br />

bounded projection on Lp(VN(F)) for every 1


148 Q. Xu<br />

l<strong>in</strong>ear, while ξ ↦→ a(ξ) is anti-l<strong>in</strong>ear. We have the follow<strong>in</strong>g free commutation<br />

relation:<br />

a(η)c(ξ) =〈ξ, η〉1, ∀ ξ, η ∈ H. (7.7)<br />

Now assume that H is the complexification of a real Hilbert space HR.<br />

The vectors <strong>in</strong> HR are called real. Letξ∈H be real and def<strong>in</strong>e<br />

s(ξ) =c(ξ)+a(ξ).<br />

s(ξ) isasemicircular element <strong>in</strong> Voiculescu’s sense. We will also call it a free<br />

Gaussian variable . Note that the map ξ ↦→ s(ξ) is real l<strong>in</strong>ear from HR <strong>in</strong>to<br />

B(F(H)). Then the free von Neumann algebra Γ (H) associated with H is the<br />

von Neumann subalgebra of B(F(H)) generated by all s(ξ) withrealξ∈H: Γ0(H) = � � ′′<br />

s(ξ) :ξ ∈ HR ⊂ B(F(H)) .<br />

The vector state τ0 def<strong>in</strong>ed by the vacuum, x ↦→ 〈x1l, 1l〉 is faithful and tracial<br />

on Γ0(H). We refer to [VDN] for more <strong>in</strong>formation.<br />

Let us mention some basic properties of free Gaussian variables. Let ξ ∈ H<br />

be a unit real vector. By def<strong>in</strong>ition, s(ξ) is selfadjo<strong>in</strong>t; its spectrum is [−2, 2]<br />

and the correspond<strong>in</strong>g measure <strong>in</strong>duced by τ0 is the so-called Wigner measure<br />

dµ(t) = 1 �<br />

4 − t2 dt.<br />

2π<br />

Thus for any 1 ≤ p


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 149<br />

where ηj = �<br />

k ujk ξk. Theηjform an orthonormal family. This orthogonal<br />

<strong>in</strong>variance implies that if (s(ξk)) is a standard semicircular system, then for<br />

(αk) ⊂ R such that �<br />

k |αk| 2 < ∞ and any 1 ≤ p ≤∞<br />

�<br />

� �<br />

αk s(ξk) � � = p � �<br />

|αk| 2�1/2 � �<br />

�s(ξ) �p .<br />

k≥1<br />

k≥1<br />

In particular, (s(ξk)) is an orthonormal sequence <strong>in</strong> L2(Γ (H)).<br />

For simplicity we now assume that H is <strong>in</strong>f<strong>in</strong>ite dimensional and separable<br />

and put Γ0 = Γ0(H). Let (ek) be an orthonormal basis of H and set sk =<br />

s(ek). Thus (sk) is a semicircular sequence and Γ0 is generated by the sk.<br />

The follow<strong>in</strong>g <strong>in</strong>equality for semicircular systems comes from [P2] .<br />

Theorem 7.9. Let 1 ≤ p ≤∞and (xk) be a f<strong>in</strong>ite sequence <strong>in</strong> Sp. Then<br />

�<br />

� � �<br />

xk ⊗ sk�<br />

Lp(B(ℓ2) ¯⊗Γ0) ∼c<br />

�<br />

� (xk) � � . (7.8)<br />

CRp[Sp]<br />

k<br />

Moreover, the closed subspace of Lp(Γ0) generated by the sk is completely<br />

complemented <strong>in</strong> Lp(Γ0) with constant 2.<br />

Proof. This proof is similar to that of Theorem 7.8. Aga<strong>in</strong>, the case 1


150 Q. Xu<br />

whence � � � �<br />

Similarly,<br />

k<br />

�<br />

� � �<br />

k<br />

x ∗ k xk<br />

xkx ∗ k<br />

� 1/2� �∞ ≤ � � �<br />

k<br />

� 1/2� �∞ ≤ � � �<br />

k<br />

�<br />

xk ⊗ sk�<br />

.<br />

∞<br />

�<br />

xk ⊗ sk�<br />

.<br />

∞<br />

Thus the lower estimate of (7.8) for p = ∞ follows.<br />

(7.8) for p = 1 and the complementation assertion are proved <strong>in</strong> the same<br />

way as the correspond<strong>in</strong>g assertions of Theorem 7.8. The only difference is<br />

the fact that the l<strong>in</strong>ear span of the sk is no longer w*-dense <strong>in</strong> Γ0. Instead,<br />

we have to use the family of polynomials <strong>in</strong> the sk which is w*-dense <strong>in</strong><br />

Γ0. This family is the l<strong>in</strong>ear span of all Wick products associated with the<br />

elementary tensors formed from the basis vectors ek. It then rema<strong>in</strong>s to note<br />

that these Wick products form an orthonormal basis of L2(Γ0). We omit<br />

the details. ⊓⊔<br />

Remark 7.10. The free Gaussian variables <strong>in</strong> the theorem above can be replaced<br />

by Bo˙zejko-Speicher’s q-Gaussians for −1


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 151<br />

for p


152 Q. Xu<br />

We claim that<br />

�<br />

� � �<br />

k<br />

uku ∗ k ⊗ bkb ∗ k<br />

�1/2� �p<br />

′ ≤ sup �bk�p<br />

k<br />

′<br />

�<br />

� � �<br />

k<br />

uku ∗�1/2� �p<br />

k<br />

′ .<br />

Indeed, this is obvious for p ′ =2andp ′ = ∞. Then complex <strong>in</strong>terpolation<br />

yields the case 2


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 153<br />

<strong>in</strong>equality [Ha], we deduce that (E(X)) q ≥ E(X q ) for any positive X ∈<br />

B(ℓ2) ¯⊗M. This clearly implies (7.11).<br />

Return back to our task. Us<strong>in</strong>g (7.11) with q = p/2 (recall<strong>in</strong>g that 1 ≤<br />

p


154 Q. Xu<br />

<strong>in</strong> Remark 7.10. In particular, for q = −1, we get the noncommutative<br />

Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality for a sequence of Fermions.<br />

7.7 Generalized Circular Systems. Rem<strong>in</strong>d that we wish to embed OH<br />

<strong>in</strong>to a noncommutative L1-space by us<strong>in</strong>g a certa<strong>in</strong> noncommutative Kh<strong>in</strong>tch<strong>in</strong>e<br />

type <strong>in</strong>equality. Namely, we have to prove that OH is completely isomorphic<br />

to the closed subspace generated by a sequence of random variables<br />

<strong>in</strong> an L1(M) for a von Neumann algebra M. Such a sequence cannot satisfy<br />

the assumption of Theorem 7.11 for OH is not completely isomorphic<br />

to CR1. In fact, Pisier [P5] showed that OH cannot completely embed <strong>in</strong>to<br />

an L1(M) withM semif<strong>in</strong>ite (i.e., of type I or II). This expla<strong>in</strong>s why we are<br />

forced to seek for Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities for random variables <strong>in</strong> a non<br />

tracial probability space, i.e., the underly<strong>in</strong>g von Neumann algebra is of type<br />

III. We give below only one example of this k<strong>in</strong>d.<br />

Fix an <strong>in</strong>f<strong>in</strong>ite dimensional separable Hilbert space H as <strong>in</strong> subsection 7.5.<br />

Let {e±k}k≥1 be an orthonormal basis of H. We also fix a sequence {λk}k≥1<br />

of positive numbers. Let<br />

gk = c(ek)+ � λk a(e−k).<br />

(gk)k≥1 is a generalized circular system <strong>in</strong> Shlyakhtenko’s sense [S]. Let Γ be<br />

the von Neumann algebra on the full Fock space F(H) generated by the gk.<br />

Let ρ be the vector state on Γ determ<strong>in</strong>ed by the vacuum 1l. Then ρ is faithful<br />

on Γ . By the identification of L1(Γ ) with the predual Γ∗, ρ is a positive unit<br />

element of L1(Γ ), so for any 1 ≤ p ≤∞, ρ 1/p is a positive unit element of<br />

Lp(Γ ), and thus gk ρ 1/p ∈ Lp(Γ ) for any k.<br />

Shlyakhtenko proved that the algebra Γ is a type IIIλ factor (0


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 155<br />

iii) Let Gp be the closed subspace of Lp(Γ ) generated by {gkρ1/p }k≥1. Then<br />

Gp is completely complemented <strong>in</strong> Lp(Γ ) with constant 2.<br />

Proof. We prove only the cases p = ∞, 1 and the complementation assertion.<br />

The proof of i) for p = ∞ is the same as that of (7.4). Let us prove iii), which<br />

will imply ii) for p = 1 by duality. As <strong>in</strong> the semicircular case, consider the<br />

projection P from the algebra of polynomials on the gk onto the l<strong>in</strong>ear span<br />

of the gk. LetQ = I − P . It is easy to see that P (x) andQ(x) are orthogonal<br />

relative to both scalar products (a, b) ↦→ ρ(b∗a)and(a, b) ↦→ ρ(ab∗ ). Now let<br />

x be a polynomial on the gk with coefficients <strong>in</strong> S∞. Writexas x =idS∞⊗P (x)+idS∞⊗ Q(x) = �<br />

ak ⊗ gk +idS∞⊗ Q(x).<br />

Then us<strong>in</strong>g the argument yield<strong>in</strong>g the first <strong>in</strong>equality of (7.4), we get<br />

max �� � � �<br />

a ∗ kak �1/2� �∞ , � � � �<br />

λk aka ∗�1/2� �<br />

�∞<br />

k ≤�x�∞ .<br />

k<br />

Therefore, us<strong>in</strong>g i) <strong>in</strong> the case p = ∞, we deduce that P is completely<br />

bounded. This implies, <strong>in</strong> turn, ii) for p = 1 by duality. Let us also note<br />

that P extends to a completely bounded normal projection on Γ . Indeed, by<br />

the density of {xρ : x polynomial on the gk} <strong>in</strong> L1(Γ ), we f<strong>in</strong>d that ii) <strong>in</strong> the<br />

case p =1showsthatPadmits a pre-adjo<strong>in</strong>t on L1(Γ ) which is completely<br />

bounded. F<strong>in</strong>ally, by <strong>in</strong>terpolation we obta<strong>in</strong> iii) for 1


156 Q. Xu<br />

Theorem 7.8 or Theorem 7.9. This is how Pisier shows that the CRp[Sp] form<br />

a complex <strong>in</strong>terpolation scale (see [P2, Theorem 8.4.8]). In the same way, us<strong>in</strong>g<br />

Theorem 7.13 one deduces that the weighted versions of CRp[Sp] def<strong>in</strong>ed<br />

<strong>in</strong> the previous remark are also an <strong>in</strong>terpolation scale (see [JPX] for more<br />

details).<br />

Remark 7.16. Theorem 7.13 admits a Fermionic analogue. In this case the<br />

correspond<strong>in</strong>g algebra is an Araki-Woods hyperf<strong>in</strong>ite type III factor. The result<strong>in</strong>g<br />

<strong>in</strong>equality holds only for p2<br />

Cp[Sp]<br />

� �<br />

k<br />

�<br />

xkεn�<br />

if p


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 157<br />

sequel are only completely isomorphic, we will forget the <strong>in</strong>dex ∞ or 1 <strong>in</strong> all<br />

direct sums. The preced<strong>in</strong>g theorem is Exercise 7.8 of [P3]. The proof there<br />

gives the follow<strong>in</strong>g more precise statement.<br />

Theorem 8.2. There exists an <strong>in</strong>jective positive selfadjo<strong>in</strong>t (unbounded) operator<br />

∆ : C → R such that OH is completely isometric to a quotient of<br />

G(∆), whereG(∆) is the graph of ∆<br />

G(∆) = � (h, ∆h) : h ∈ D(∆) � ⊂ C ⊕ R,<br />

considered as a subspace of C ⊕ R.<br />

Recall that both C and R are isometric to ℓ2 as Banach spaces. Thus ∆<br />

is an operator on ℓ2 with doma<strong>in</strong> D(∆). S<strong>in</strong>ce ∆ is positive and <strong>in</strong>jective,<br />

its range is also dense. Pass<strong>in</strong>g to duality, we see that OH ∗ is completely<br />

isometric to a subspace of G(∆) ∗ .S<strong>in</strong>ceOH ∗ = OH completely isometrically,<br />

embedd<strong>in</strong>g OH <strong>in</strong>to an L1 is reduced to embedd<strong>in</strong>g G(∆) ∗ <strong>in</strong>to an L1.<br />

To see why Kh<strong>in</strong>tch<strong>in</strong>e type <strong>in</strong>equalities can help us for such a matter, let<br />

us describe the operator space structure of G(∆):<br />

G(∆) ={(h, ∆(h)) : h ∈ D(∆)} ⊂C ⊕∞ R.<br />

Let xk ∈ S∞ and hk ∈ D(∆). Let x = �<br />

k xk ⊗(hk, ∆(hk)) ∈ S∞ ⊗m<strong>in</strong> G(∆).<br />

Then<br />

�x�S∞⊗m<strong>in</strong>G(∆) =max �� � �<br />

k,j<br />

〈hk, hj〉x ∗ j xk�1/2<br />

�<br />

S∞ , � � �<br />

k,j<br />

〈∆(hk), ∆(hj)〉xkx ∗ j<br />

�<br />

� 1/2<br />

S∞<br />

This is a cont<strong>in</strong>uous version of the space (for p = ∞) <strong>in</strong>troduced <strong>in</strong> Remark<br />

7.14. To simplify our discussion and without loss of generality by a simple<br />

argument of approximation, we may assume that ∆ has pure po<strong>in</strong>t spectrum,<br />

i.e., ℓ2 has an orthonormal basis (ek) of eigenvectors of ∆. Let(µk) bethe<br />

associated eigenvalues: ∆ek = µkek. It then follows that G(∆) co<strong>in</strong>cides<br />

(completely isometrically) with the diagonal subspace C ∩ R((λk)) of C ⊕∞<br />

R((λk)), where λk = µ 2 k .<br />

By duality, we deduce that G(∆) ∗ = C1 + R1((λ −1<br />

k<br />

)). More precisely, the<br />

operator space structure of G(∆) ∗ is determ<strong>in</strong>ed as follows: for any f<strong>in</strong>ite<br />

sequence (xk) ⊂ S1:<br />

�x�S1[G(∆) ∗ ] =<strong>in</strong>f �� � � �<br />

y ∗ kyk �1/2� �S1 + � � � �<br />

λ −1<br />

k zkz ∗�1/2� �<br />

�S1<br />

k ,<br />

k<br />

where the <strong>in</strong>fimum runs over all decompositions xk = yk+zk <strong>in</strong> S1. Therefore,<br />

by the Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality <strong>in</strong> Theorem 7.13 with p =1,G(∆) ∗ is completely<br />

isomorphic to G1 there. Thus we have proved the follow<strong>in</strong>g<br />

k<br />

� .


158 Q. Xu<br />

Theorem 8.3. OH is completely isomorphic to a subspace of a noncommutative<br />

L1-space.<br />

Remark 8.4. i) The proof of Theorem 8.3 gives much more. In fact, it shows<br />

that the dual space of any graph <strong>in</strong> C ⊕ R completely embeds <strong>in</strong>to an L1.<br />

From this one can deduce that a quotient of a subspace of C ⊕ R completely<br />

embeds <strong>in</strong>to an L1. We refer to [J2], [P4] and [X1] for more <strong>in</strong>formation.<br />

ii) The von Neumann algebra Γ is not hyperf<strong>in</strong>ite. OH also completely<br />

embeds <strong>in</strong>to the predual of a hyperf<strong>in</strong>ite algebra (see [J3], [HM]).<br />

References<br />

[BKS] M. Bo˙zejko, B. Kümmerer and R. Speicher. q-Gaussian processes: noncommutative<br />

and classical aspects. Comm. Math. Phys., 185:129–154, 1997.<br />

[BS] M. Bo˙zejko and R. Speicher. An example of a generalized Brownian motion.<br />

Comm. Math. Phys., 137:519–531, 1991.<br />

[ER] Ed. Effros and Z-J. Ruan. Operator spaces. The Clarendon Press Oxford University<br />

Press, New York, 2000.<br />

[HM] U. Haagerup and M. Musat. On the best constants <strong>in</strong> noncommutative Kh<strong>in</strong>tch<strong>in</strong>e<br />

type <strong>in</strong>equalities. J. Funct. Anal., 250(2):588–624, 2007.<br />

[HP] U. Haagerup and G. Pisier. Bounded l<strong>in</strong>ear operators between C ∗ -algebras. Duke<br />

Math. J., 71:889–925, 1993.<br />

[Ha] F. Hansen. An operator <strong>in</strong>equality. Math. Ann., 246:249–250, 1979/80.<br />

[J1] M. Junge. Doob’s <strong>in</strong>equality for non-commutative mart<strong>in</strong>gales. J. Re<strong>in</strong>e Angew.<br />

Math., 549:149–190, 2002.<br />

[J2] M. Junge. Embedd<strong>in</strong>g of the operator space OH and the logarithmic ‘little<br />

Grothendieck <strong>in</strong>equality’. Invent. Math., 161:225–286, 2005.<br />

[J3] M. Junge. Operator spaces and Araki-Woods factors: a quantum probabilistic<br />

approach. IMRP Int. Math. Res. Pap., pages Art. ID 76978, 87, 2006.<br />

[JO] M. Junge and T. Oikhberg. Unconditional basic sequences and homogeneous<br />

Hilbertian subspaces of non-commutative Lp spaces. Indiana Univ. Math. J.,<br />

56:733–765, 2007.<br />

[JPX] M. Junge and J. Parcet and Q. Xu. Rosenthal type <strong>in</strong>equalities for free chaos.<br />

Ann. Proba. 35:1374–1437, 2007.<br />

[JX] M. Junge and Q. Xu. Noncommutative Burkholder/Rosenthal <strong>in</strong>equalities II:<br />

Applications. Isreal J. Math., to appear.<br />

[Ka] J-P. Kahane. Some random series of functions. Cambridge University Press,<br />

Cambridge, second edition, 1985.<br />

[Ko] H. Kosaki. Applications of the complex <strong>in</strong>terpolation method to a von Neumann<br />

algebra: noncommutative L p -spaces. J. Funct. Anal., 56:29–78, 1984.<br />

[KR] R. V. Kadison and J. R. R<strong>in</strong>grose. Fundamentals of the theory of operator algebras.<br />

Vol. II. Amer. Math. Soc., Providence, RI, 1997.<br />

[Pa] V. Paulsen. Completely bounded maps and operator algebras, volume 78 of<br />

Cambridge Studies <strong>in</strong> Advanced Mathematics. Cambridge University Press,<br />

Cambridge, 2002.<br />

[LPP] F. Lust-Piquard and G. Pisier. Noncommutative Kh<strong>in</strong>tch<strong>in</strong>e and Paley <strong>in</strong>equalities.<br />

Ark. Mat., 29:241–260, 1991.<br />

[P1] G. Pisier. The operator Hilbert space OH, complex <strong>in</strong>terpolation and tensor norms.<br />

Mem. Amer. Math. Soc., 122(585):viii+103, 1996.<br />

[P2] G. Pisier. Non-commutative vector valued Lp-spaces and completely p-summ<strong>in</strong>g<br />

maps. Astérisque, (247):vi+131, 1998.


Interactions between Quantum <strong>Probability</strong> and Operator Space <strong>Theory</strong> 159<br />

[P3] G. Pisier. Introduction to operator space theory. Cambridge University Press,<br />

Cambridge, 2003.<br />

[P4] G. Pisier. Completely bounded maps <strong>in</strong>to certa<strong>in</strong> Hilbertian operator spaces. Int.<br />

Math. Res. Not., (74):3983–4018, 2004.<br />

[P5] G. Pisier. The operator Hilbert space OH and type III von Neumann algebras.<br />

Bull. London Math. Soc., 36:455–459, 2004.<br />

[PS] G. Pisier and D. Shlyakhtenko. Grothendieck’s theorem for operator spaces. Invent.<br />

Math., 150:185–217, 2002.<br />

[PX] G. Pisier and Q. Xu. Non-commutative L p -spaces. In Handbook of the geometry<br />

of Banach spaces, Vol. 2, pages 1459–1517. North-Holland, Amsterdam, 2003.<br />

[R] Z-J. Ruan. Subspaces of C ∗ -algebras. J. Funct. Anal., 76:217–230, 1988.<br />

[S] D. Shlyakhtenko. Free quasi-free states. Pacific J. Math., 177:329–368, 1997.<br />

[T] M. Takesaki. <strong>Theory</strong> of operator algebras. I. Spr<strong>in</strong>ger-Verlag, New York, 1979.<br />

[VDN] D. V. Voiculescu, K. J. Dykema and A. Nica. Free random variables. Amer.Math.<br />

Soc., Providence, RI, 1992.<br />

[X1] Q. Xu. Embedd<strong>in</strong>g of Cq and Rq <strong>in</strong>to noncommutative Lp-spaces, 1 ≤ p


Dirichlet Forms on Noncommutative<br />

Spaces<br />

Fabio Cipriani<br />

Abstract We show how Dirichlet forms provide an approach to potential theory<br />

of noncommutative spaces based on the notion of energy. The correspondence<br />

with KMS-symmetric Markovian semigroups is expla<strong>in</strong>ed <strong>in</strong> details<br />

and applied to the dynamical approach to equilibria of quantum sp<strong>in</strong> systems.<br />

Second part focuses on the differential calculus underly<strong>in</strong>g a Dirichlet<br />

form. Applications are given <strong>in</strong> Riemannian Geometry to a potential theoretic<br />

characterization of spaces with positive curvature and to the construction of<br />

Fredholm modules <strong>in</strong> Noncommutative Geometry.<br />

1 Introduction<br />

Our purpose <strong>in</strong> these notes is to <strong>in</strong>troduce the reader to those aspects of<br />

<strong>Potential</strong> <strong>Theory</strong> underly<strong>in</strong>g the notion of energy on noncommutative spaces<br />

such as<br />

• fractal spaces whose topology is not underly<strong>in</strong>g any manifold structure<br />

• Riemannian manifolds <strong>in</strong> sp<strong>in</strong> geometry<br />

• space of orbits of dynamical systems<br />

• space of leaves of Riemannian foliations<br />

• space of irreducible unitary representations of a discrete group<br />

• space of observables of sp<strong>in</strong> systems <strong>in</strong> Quantum Statistical Mechanics<br />

• space of random variables <strong>in</strong> Free <strong>Probability</strong>.<br />

To handle the complexity of spaces of this type, the classical tools of analysis<br />

like measure theory, topology and differential calculus, are unfitted or<br />

F. Cipriani<br />

Dipartimento di Matematica, Politecnico di Milano, Piazza Leonardo da V<strong>in</strong>ci, 32, I-20133<br />

Milano, Italy<br />

e-mail: fabio.cipriani@polimi.it<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 161<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


162 F. Cipriani<br />

<strong>in</strong>sufficient as they force to treat<strong>in</strong>g these spaces as s<strong>in</strong>gular. However, as<br />

emphasized by A. Connes <strong>in</strong> Noncommutative Geometry, the properties of<br />

these spaces are naturally encoded <strong>in</strong> algebras which, generaliz<strong>in</strong>g the familiar<br />

abelian ones of measurable, cont<strong>in</strong>uous or smooth functions, are <strong>in</strong> general<br />

noncommutative.<br />

The aim of this <strong>in</strong>troduction is to recall briefly the role of energy <strong>in</strong> Newtonian<br />

and Beurl<strong>in</strong>g-Deny generalized potential theories, then to show the natural<br />

emergence of a potential theory <strong>in</strong> various noncommutative spaces and then<br />

to describe the content of the sections of these notes.<br />

<strong>Classical</strong> potential theory analyzes properties of force field F <strong>in</strong> Euclidean<br />

spaces Rd , like <strong>in</strong>compressibility, irrotationality or conservativity, <strong>in</strong> which<br />

case one is look<strong>in</strong>g, at least locally, for potentials U such that F = ∇U. In<br />

regions Ω where charges are absent, the potential U is harmonic, as it solves<br />

the Laplace equation ∆U = 0. One of the most important achievements of<br />

eighteenth century analysis was the solution of the Dirichlet problem, i.e. the<br />

determ<strong>in</strong>ation of the potential U harmonic <strong>in</strong> Ω once its boundary values are<br />

known, under suitable regularity assumptions on ∂Ω. One of the methods<br />

<strong>in</strong>vented to solve this problem focuses on the construction of the Green’s and<br />

Poisson’s kernels on Ω, by which a solution U admits an <strong>in</strong>tegral representation<br />

<strong>in</strong> terms of its boundary values on ∂Ω.<br />

To recall the multiple connections of classical potential theory with other<br />

fields of mathematics, it suffices to mention those with conformal geometry.<br />

In fact, by methods dat<strong>in</strong>g back to B. Riemann and D. Hilbert (see [Abk]<br />

for a nice account of that long story), the conformal equivalence given by<br />

the Uniformization Theorem, between a simply connected region Ω hav<strong>in</strong>g<br />

smooth enough boundary and the unit disk of the complex plane C, gives<br />

rise to the Green function of Ω.<br />

A second method to solve the Dirichlet problem is variational <strong>in</strong> nature and<br />

it consists of seek<strong>in</strong>g the solution U as the unique m<strong>in</strong>imizer of the Dirichlet<br />

energy <strong>in</strong>tegral<br />

�<br />

E[u] = |∇u(x)| 2 dx (1.1)<br />

Ω<br />

among the cont<strong>in</strong>uous functions hav<strong>in</strong>g the prescribed boundary values. The<br />

characteristic property of this energy functional is that it does not <strong>in</strong>crease<br />

upon modify<strong>in</strong>g the function from u to u ∧ 1:=<strong>in</strong>f(u, 1):<br />

E[u ∧ 1] ≤E[u]. (1.2)<br />

It is easy to recognize here a basic feature of the energy functional of many<br />

physical systems, such as the case where u is meant to describe the values of<br />

the potential at the vertices of an electrical circuit.<br />

In the fall of the ’50’s of the previous century, A. Beurl<strong>in</strong>g and J. Deny<br />

developed <strong>in</strong> two sem<strong>in</strong>al papers [BD1] and [BD2] a generalized potential<br />

theory on locally compact spaces. As their approach relies upon the notion of


Dirichlet Forms on Noncommutative Spaces 163<br />

energy where potentials appear as derived objects, it is a kernel-free po<strong>in</strong>t of<br />

view.<br />

In that theory a Dirichlet space is a locally compact topological Hausdorff<br />

space X equipped with a Dirichlet form E : C0(X) → (−∞, +∞]: this is just a<br />

quadratic, lower semi-cont<strong>in</strong>uous functional, def<strong>in</strong>ed on the algebra C0(X) of<br />

cont<strong>in</strong>uous functions vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity, which is f<strong>in</strong>ite on a dense subspace<br />

and satisfies the characteristic contraction property (1.2).<br />

In study<strong>in</strong>g these functionals, Hilbertian methods are quite natural: by a<br />

result of G. Mokobodzki [Mok], there always exists a positive Radon measure<br />

m on X such that E admits a lower semi-cont<strong>in</strong>uous extension to L 2 (X, m).<br />

In other words, the functional E may be considered, <strong>in</strong> a natural way, as the<br />

closed quadratic form of a self-adjo<strong>in</strong>t operator H on L 2 (X, m), with respect<br />

to a large family of reference measures on X.<br />

The nuance between the theory on the algebra C0(X), and the one on the<br />

Hilbert space L 2 (X, m), relies on the geometrical character of the former and<br />

the dynamical character of the latter.<br />

In this respect, A. Beurl<strong>in</strong>g and J. Deny discovered the fundamental dynamical<br />

characterization of the contraction property of an energy functional E on<br />

L 2 (X, m), <strong>in</strong> terms of the semigroup {e −tH : t ≥ 0} generated by H: the contraction<br />

property of E is equivalent to the so called Markovianity property.<br />

This consists of positivity, <strong>in</strong> the sense that the maps e −tH preserve positivity<br />

of functions, and of contractivity, bywhiche −tH is contractive both<br />

with respect to the Hilbertian norm of L 2 (X, m) and with respect to the uniform<br />

norm of the algebra L ∞ (X, m). In turn, by <strong>in</strong>terpolation and duality,<br />

this implies contractivity with respect to the whole scale of Lebesgue spaces<br />

L p (X, m). In other terms, the solution u(t) =e −tH u(0) of the generalized<br />

heat equation ∂tu = −Hu associated to H satisfies a maximum pr<strong>in</strong>ciple.<br />

The dynamical po<strong>in</strong>t of view is also fundamental for the probabilistic counterpart<br />

of the theory. By the work of M. Fukushima (see [FOT]), a Dirichlet<br />

form E on C0(X) givesrisetoa family of Markov-Hunt stochastic processes<br />

on X. These processes are <strong>in</strong>dexed, essentially, by the Radon measures m<br />

hav<strong>in</strong>g X as support with respect to which E is closable <strong>in</strong> L 2 (X, m). In a<br />

sense, the various processes are essentially the same <strong>in</strong> that they differ one<br />

from another by a random time change only, associated to their speed measures<br />

m. Inthissensethe functional E on the algebra C0(X) is a geometric<br />

object represent<strong>in</strong>g an equivalence class of stochastic processes on X hav<strong>in</strong>g<br />

the same paths.<br />

The transition functions of these processes co<strong>in</strong>cide with the heat kernels of<br />

the Markovian semigroups and one obta<strong>in</strong>s a probabilistic <strong>in</strong>terpretation of<br />

the solution of the generalized heat equation, which extends the well known<br />

relationships among Laplace operator, heat equation and Brownian motion<br />

on R n (see for example [Do]).


164 F. Cipriani<br />

The essential tool needed to associate a Markov process to a Dirichlet form<br />

E and a speed measure m on X, <strong>in</strong> the Beurl<strong>in</strong>g-Deny generalized potential<br />

theory are those functions u, called potentials, such that<br />

E[u + v] ≥E[u]+E[v] (1.3)<br />

for all positive functions v. The class of potentials then def<strong>in</strong>e a notion of<br />

smallness, called polarity, for subsets of X, which is well suited to discuss both<br />

the f<strong>in</strong>e analytic properties of energy and the f<strong>in</strong>e probabilistic properties of<br />

the process.<br />

The need to extend tools like Dirichlet forms and Markovian semigroups,<br />

beyond the above commutative sett<strong>in</strong>g, was recognized a long time ago <strong>in</strong><br />

Quantum Field <strong>Theory</strong>, where, for example, quadratic form techniques were<br />

employed to generate and analyze Hamiltonian operators represent<strong>in</strong>g the<br />

energy of the system.<br />

In the second quantization of Boson systems the symmetric Fock space is<br />

isomorphic to the Lebesgue space L 2 (X,γ) over an <strong>in</strong>f<strong>in</strong>ite dimensional<br />

Gaussian space (X ,γ) and the number operator N <strong>in</strong>to the generator of<br />

a symmetric Markovian semigroup over the commutative von Neumann algebra<br />

L ∞ (X, γ). This allowed the developement of non perturbative techniques,<br />

like hypercontractivity and logarithmic Sobolev <strong>in</strong>equalities, help<strong>in</strong>g the construction<br />

of <strong>in</strong>teract<strong>in</strong>g quantum fields (see [RS Section X.9, Notes]).<br />

An extension of these methods, requir<strong>in</strong>g a noncommutative sett<strong>in</strong>g, wasdeveloped<br />

by L. Gross <strong>in</strong> [G1], to deal with problems of existence and uniqueness<br />

of ground states of Hamiltonians describ<strong>in</strong>g Fermions systems. There, the relevant<br />

algebra is the Clifford algebra Cl(h) of an <strong>in</strong>f<strong>in</strong>ite dimensional Hilbert<br />

space h. Thisisanoncommutative von Neumann algebra, with trivial center<br />

and f<strong>in</strong>ite normal trace τ, hence a II1 factor, which is another example of<br />

those algebras (discovered by F.J. Murray and J. von Neumann [MvN] <strong>in</strong><br />

the early days of the theory of operators algebras) display<strong>in</strong>g a cont<strong>in</strong>uous<br />

geometry.<br />

L. Gross discovered that the Hamiltonian operators H of these systems are<br />

the generators of Markovian semigroups on the Clifford algebra, and then<br />

of a family of contraction semigroups on the noncommutative L p (Cl(h) ,τ)<br />

spaces associated to the trace. The construction of these spaces, which generalize<br />

Lebesgue’s one <strong>in</strong> the commutative case, was achieved by I. Segal [Se]<br />

and reformulated by E. Nelson [Ne]. The method concern<strong>in</strong>g the uniqueness<br />

problem may be seen as a generalization of Perron’s theory of matrices with<br />

positive entries, while the tools needed to approach the existence problem<br />

exploit hypercontractivity aga<strong>in</strong>.<br />

The study of Dirichlet forms <strong>in</strong> the noncommutative sett<strong>in</strong>g of a C ∗ or<br />

von Neumann algebra M and a semif<strong>in</strong>ite trace τ on it was pioneered by<br />

S. Albeverio and R. Hoegh-Krohn <strong>in</strong> [AHK1,2], where, <strong>in</strong> particular, they


Dirichlet Forms on Noncommutative Spaces 165<br />

obta<strong>in</strong>ed the generalization of the Beurl<strong>in</strong>g-Deny characterization of Markovian<br />

semigroups, <strong>in</strong> terms of Dirichlet forms.<br />

This theory was subsequently developed by J.-L. Sauvageot [S2,3],<br />

E. B. Davies-J. M. L<strong>in</strong>dsay [DL1,2]. In particular, D. Guido - T. Isola -<br />

S. Scarlatti [GIS] generalized the Beurl<strong>in</strong>g-Deny correspondence to non<br />

symmetric Dirichlet forms and Markovian semigroups.<br />

This theory met several fields of application. In [S4] and [S7] the author<br />

constructed the transverse Laplacian operator and the correspond<strong>in</strong>g<br />

transverse heat semigroup on the C ∗ -algebra of a Riemannian foliation,<br />

def<strong>in</strong>ed by A. Connes <strong>in</strong> the framework of Noncommutative Geometry [Co2].<br />

This Markovian semigroup was subsequently dilated <strong>in</strong> [S6] to a noncommutative<br />

Markovian stochastic process <strong>in</strong> the sense of Quantum <strong>Probability</strong><br />

[AFLe], [Par].<br />

In the framework of Riemannian Geometry, E. B. Davies - O. Rothaus proved<br />

[DR1] that the Bochner Laplacian of the Levi-Civita connection of a Riemannian<br />

manifold M generates a Markovian semigroup on the Clifford algebra<br />

of M, and used the result to <strong>in</strong>vestigate spectral bounds [DR2]. Markov<br />

structures on the Clifford algebra of a Riemannian manifold were also studied<br />

<strong>in</strong> [SU]. In [D2] E. B. Davies comb<strong>in</strong>ed the technique of Dirichlet forms<br />

on von Neumann semif<strong>in</strong>ite algebras with some ideas of A. Connes <strong>in</strong> Noncommutative<br />

Geometry, to obta<strong>in</strong> sharp heat kernel bounds on graphs, hence<br />

approach<strong>in</strong>g through “noncommutative tools” a purely commutative classical<br />

problem. More recently, D. Goswami and K.B. S<strong>in</strong>ha [GS], us<strong>in</strong>g the first<br />

order differential calculus associated to a Dirichlet form developed <strong>in</strong> [CS1]<br />

and illustrated <strong>in</strong> Chapter below, see also [CS4], succeeded <strong>in</strong> construct<strong>in</strong>g a<br />

noncommutative stochastic process associated to a Dirichlet form, solv<strong>in</strong>g a<br />

quantum stochastic differential equation with unbounded generator.<br />

Despite the fact that traces on operators algebras are quite rare and even<br />

sometimes absent, the need to extend the theory to non tracial states or<br />

weights on C ∗ or von Neumann algebras was suggested by possible applications<br />

<strong>in</strong> at least two fields of Mathematical Physics.<br />

In the theory of Quantum Open Systems [D1] and <strong>in</strong> Quantum Measurement<br />

<strong>Theory</strong> [B], Markovian semigroups, also called dynamical semigroups, appear<br />

naturally as solutions of the master equation: they model dissipative dynamics<br />

of quantum systems coupled to thermal reservoirs [AFLu] or cont<strong>in</strong>uous<br />

measurement processes of a “small” quantum system coupled to a larger one<br />

(which represents the <strong>in</strong>strument).<br />

On the other hand, the progresses made by D. Stroock - B. Zegarl<strong>in</strong>ski at the<br />

end of the eighties <strong>in</strong> the stochastic approach to the equilibrium <strong>in</strong> <strong>Classical</strong><br />

Statistical Mechanics [Lig] suggested the developement of similar tools for<br />

KMS equilibrium states <strong>in</strong> Quantum Statistical Mechanics [MZ1], [MZ2],<br />

[MOZ]. At f<strong>in</strong>ite temperature however, KMS-states are never tracial states<br />

but give rise to type III von Neumann algebras [Del].


166 F. Cipriani<br />

This led to a general formulation of Dirichlet forms and symmetric Markovian<br />

semigroups on von Neumann algebras and to an extension of the Beurl<strong>in</strong>g-<br />

Deny correspondence with respect to faithful and normal states [Cip1],<br />

[GL1,2] and to faithful, normal and semif<strong>in</strong>ite weights [GL3]. This approach<br />

provides <strong>in</strong> particular a maximum pr<strong>in</strong>ciple for solutions of Dirichlet problems<br />

on C∗-algebras [S5], [Cip2], [CS3].<br />

There are two ma<strong>in</strong> differences with respect to the tracial sett<strong>in</strong>g. While <strong>in</strong><br />

the commutative and tracial cases (M ,τ) it is quite clear how to construct<br />

the Hilbert space L2 (M,τ) on which the Markovian semigroups should act,<br />

<strong>in</strong> case of a generic state ω on M the situation offers several possible choices.<br />

Beside the GNS space Hω of the Gelfand-Neimark-Segal representation<br />

(πω, Hω,ξω) ofM, concrete situations may provide more natural candidates,<br />

as for example when M is the group von Neumann algebra λ(G) ′′ of a locally<br />

compact group G.<br />

The second difference concerns the order structure on the Hilbert space<br />

L2 (M ,ω), with respect to which we will consider positivity and Markovianity<br />

of semigroups. In the commutative case, the positive cone of L2 (X, m) is<br />

generated by the positive cone of the von Neumann algebra L∞ (X, m). This<br />

is still true <strong>in</strong> the tracial case as there is a preferred embedd<strong>in</strong>g of the von<br />

Neumann algebra M <strong>in</strong>to the the various choices of L2 (M τ). In case of a non<br />

tracial state ω there are <strong>in</strong>f<strong>in</strong>itely many possible choices.<br />

In Chapter 2 we expla<strong>in</strong> how to overcome these difficulties, adopt<strong>in</strong>g the<br />

po<strong>in</strong>t of view of the Tomita-Takesaki modular theory and, <strong>in</strong> particular, us<strong>in</strong>g<br />

standard forms of von Neumann algebras. This will allow the generalization<br />

of Beurl<strong>in</strong>g-Deny theory to Markovian semigroups on von Neumann algebras<br />

satisfy<strong>in</strong>g the ω-symmetry condition with respect to a fixed, faithful, normal<br />

state ω.<br />

In the same section, however, we <strong>in</strong>troduce a more fundamental symmetry<br />

condition for maps and semigroups, not necessarily positive or Markovian,<br />

on a C∗-algebra A. It is a symmetry relation to be verified with respect to a<br />

fixed (α,β)-KMS state of a dynamical system α = {αt : t ∈ R} on A.<br />

The adopted name of KMS symmetry is <strong>in</strong>tended to suggest that this property<br />

is deeply related to, and <strong>in</strong> fact a deformation of, the KMS condition<br />

characteriz<strong>in</strong>g the equilibrium states of quantum dynamical systems.<br />

By virtue of the KMS-symmetry, it is possible to study semigroups, not necessarily<br />

positive or Markovian, <strong>in</strong> the standard forms of the von Neumann<br />

algebra generated by the KMS state ω. In Chapter 3 we illustrate applications<br />

to the convergence to equilibrium <strong>in</strong> Quantum Statistical Mechanics <strong>in</strong>itially<br />

developed by Y. M. Park [P1,2].<br />

We emphasize the fact that the KMS symmetry condition, despite its similarity<br />

with the detailed balance conditions often considered <strong>in</strong> literature when<br />

deal<strong>in</strong>g with Quantum Open Systems, does not imply the semigroup commutes<br />

with the dynamics. This fact is crucial for applications to the convergence<br />

to equilibrium <strong>in</strong> Quantum Statistical Mechanics.


Dirichlet Forms on Noncommutative Spaces 167<br />

In Chapter 4 we describe the construction of the canonical first order differential<br />

calculus associated to regular Dirichlet forms on a C∗-algebra endowed<br />

with a faithful and semif<strong>in</strong>ite trace, developed <strong>in</strong> [CS1]. When this calculus<br />

is applied to classical potential theory, on a Riemmannian space M say, this<br />

process allows the reconstruction of the Hilbert space of square <strong>in</strong>tegrable<br />

vector fields L2 (TM) and the gradient operator ∇ from the energy <strong>in</strong>tegral<br />

�<br />

E[a] = |∇a| 2 dm .<br />

M<br />

The differential calculus is understood <strong>in</strong> terms of derivations with values<br />

<strong>in</strong> Hilbert bimodules over A, i.e. <strong>in</strong> terms of representations of the maximal<br />

tensor product A ⊗max A◦ . The content of this section is taken from [CS1].<br />

Under the stronger hypothesis that the doma<strong>in</strong> of the self-adjo<strong>in</strong>t generator<br />

associated to the Dirichlet form conta<strong>in</strong>s a dense sub-algebra, it was previously<br />

proved <strong>in</strong> [S2,3]. Apply<strong>in</strong>g decomposition theory of representations<br />

of C∗-algebras, we clarify the mean<strong>in</strong>g of this differential calculus analyz<strong>in</strong>g<br />

the Beurl<strong>in</strong>g-Deny-Le Jan decomposition of Dirichlet forms E on commutative<br />

C∗-algebras. Other noncommutative examples of this differential calculus<br />

are illustrated on groups and Clifford C∗-algebras. In Section 4.4 we describe<br />

certa<strong>in</strong> derivations appear<strong>in</strong>g <strong>in</strong> D. V. Voiculescu’s Free <strong>Probability</strong>.<br />

The last two chapters are devoted to illustrat<strong>in</strong>g two applications of the<br />

canonical differential calculus.<br />

The first one, Chapter 5, concerns the relationships between noncommutative<br />

potential theory and Riemannian geometry. We first illustrate the work of<br />

E. B. Davies and O. S. Rothaus [DR1,2] show<strong>in</strong>g that the Bochner Laplacian<br />

∇ ∗ ∇ is the generator of a Markovian semigroup on the Clifford algebra Cl(M)<br />

of a Riemannian manifold M. Then we characterize Riemannian manifolds<br />

with nonnegative curvature operator as those <strong>in</strong> which the Dirac Laplacian<br />

D 2 , the square of the Dirac operator D on Cl(M), generates a Markovian<br />

semigroup on the Clifford algebra.<br />

The aim of the f<strong>in</strong>al section, Chapter 6, is to conv<strong>in</strong>ce the reader that Dirichlet<br />

forms, commutative or not, share a flavor of geometry, <strong>in</strong> the sense of A.<br />

Connes’ Noncommutative Geometry ([Co2]). In particular we show how a<br />

regular Dirichlet form E on a compact space X gives rise to a Fredholm module,<br />

<strong>in</strong> the sense of M. Atiyah [At], on the function algebra C(X), provided<br />

that the spectrum of the self-adjo<strong>in</strong>t operator associated to E is discrete and<br />

its Green function G is f<strong>in</strong>ite [CS4]. This last condition implies the spectral<br />

dimension of E to be strictly less than two. Although very restrictive at first<br />

sight, it applies however to the regular harmonic structures, constructed by<br />

J. Kigami, on post critically f<strong>in</strong>ite self-similar fractal spaces [Ki]. These are<br />

<strong>in</strong>terest<strong>in</strong>g examples of spaces whose topology does not underlie any differentiable<br />

manifold structure <strong>in</strong> the classical sense. As a consequence of the


168 F. Cipriani<br />

boundedness of G, a f<strong>in</strong>ite energy function a on X has a noncommutative<br />

differential<br />

da := i[F, a]<br />

ly<strong>in</strong>g <strong>in</strong> the Hilbert-Schmidt class of compact operators, F be<strong>in</strong>g the phase<br />

operator of the Fredholm module. This reveals the direct role that the potential<br />

theory of Dirichlet forms may play <strong>in</strong> differential topology <strong>in</strong> the sense<br />

of Nnoncommutative Geometry.<br />

These notes are based on a series of lectures held at the International<br />

Spr<strong>in</strong>g School “Quantum <strong>Potential</strong> <strong>Theory</strong>: Structures and Applications to<br />

Physics”, which took place at Greifswald Germany from February 26 th to<br />

March 9 th 2007. I wish to thank warmly the organizers as well as all the<br />

participants for several valuable discussions.<br />

2 Dirichlet Forms on C ∗ -algebras and KMS-symmetric<br />

Semigroups<br />

In this chapter we show how the Beurl<strong>in</strong>g-Deny theory can be generalized,<br />

<strong>in</strong> a natural way, to situations <strong>in</strong> which the relevant algebra is no longer<br />

the commutative algebra of cont<strong>in</strong>uous functions C0(X) over a locally compact,<br />

metrizable Hausdorff space X or the commutative algebra of essentially<br />

bounded functions L∞ (X, m), but rather a noncommutative C∗-algebra A or<br />

a noncommutative von Neumann algebra M.<br />

The objects of <strong>in</strong>terest will be maps, semigroups and quadratic forms which<br />

behave naturally, from the po<strong>in</strong>t of view of potential theory, with respect to<br />

the order structures of these types of spaces. The difficulties one encounters<br />

<strong>in</strong> deal<strong>in</strong>g with these order structures, which are proportional to the degree of<br />

noncommutativity, are taken care of by the fundamental part of noncommutative<br />

measure theory concerned with the Tomita-Takesaki theory. This theory,<br />

essentially, develops the tools needed to generalize the Radon-Nikodym theorem<br />

at the level of von Neumann algebras. One of these tools, fundamental<br />

for our purposes, will be the existence and properties of the standard form<br />

of a von Neumann algebra.<br />

Sesquil<strong>in</strong>ear scalar product <strong>in</strong> Hilbert spaces will be assumed to be l<strong>in</strong>ear <strong>in</strong><br />

the right-hand entry and conjugate l<strong>in</strong>ear <strong>in</strong> the left-hand one.<br />

2.1 Tomita-Takesaki Modular <strong>Theory</strong> and Standard<br />

Forms of von Neumann Algebras<br />

A detailed exposition of the Tomita-Takesaki modular theory would fall beyond<br />

the scope of these lectures. In this section we content ourselves with


Dirichlet Forms on Noncommutative Spaces 169<br />

collect<strong>in</strong>g some of the tools which will be <strong>in</strong> use throughout the paper. Nice<br />

references for this material are [BR1,2], [T1]. The background material about<br />

C∗-algebras and von Neumann algebras may be found <strong>in</strong> [Arv], [Dix1,2], [Ped].<br />

AC∗-algebra is an <strong>in</strong>volutive Banach algebra A, <strong>in</strong> which the norm, <strong>in</strong>volution<br />

and product are related by<br />

�a ∗ a� = �a� 2<br />

a ∈ A.<br />

The <strong>in</strong>volution determ<strong>in</strong>es the positive cone A+ := {a∗a ∈ A : a ∈ A},<br />

so that A is, <strong>in</strong> particular, an ordered vector space. Assume for simplicity<br />

that A is unital and denote by 1A its unit. A state ω ∈ A ′ + , i.e. a positive<br />

l<strong>in</strong>ear functional of norm one, gives rise to the cyclic or Gelfand-Naimark-<br />

Segal representation (πω, Hω,ξω) ofA, with cyclic and separat<strong>in</strong>g vector ξω ∈<br />

Hω. The double commutant M := πω(A) ′′ is then a von Neumann algebra<br />

of operators <strong>in</strong> Hω whose positive cone M+ co<strong>in</strong>cides both with the set of<br />

elements of M which are positive operators on Hω as well as with the weak<br />

closure of the image πω(A+) of the positive cone of A.<br />

To analyze at a Hilbert space level maps and semigroups on A and M, which<br />

leave globally <strong>in</strong>variant the positive cones, one is lead to consider <strong>in</strong> Hω the<br />

structure of an ordered vector space, and to relate it with those of A and M.<br />

There are, <strong>in</strong> general, several ways to do that, but only one is standard. This<br />

will be constructed and analyzed by the Tomita-Takesaki modular theory.<br />

A natural choice for a positive cone <strong>in</strong> the GNS space Hω could be<br />

as it is justified by the follow<strong>in</strong>g<br />

M+ ξω<br />

(2.1)<br />

Example 2.1. (Self-polarity <strong>in</strong> classical measure theory) When A is commutative<br />

and unital, hence isomorphic to the algebra of cont<strong>in</strong>uous functions C(X)<br />

over a compact Hausdorff space X, thestateω is represented by a positive<br />

Radon measure µω on X, Hω is identified with L 2 (X, µω), the representation<br />

πω is given by the obvious action of cont<strong>in</strong>uous functions on the square<br />

<strong>in</strong>tegrable ones and the von Neumann algebra M co<strong>in</strong>cides with L ∞ (X, µω).<br />

S<strong>in</strong>ce the cyclic vector ξω corresponds to the constant function 1 ∈ L 2 (X, µω),<br />

<strong>in</strong> this sett<strong>in</strong>g positivity has the usual po<strong>in</strong>twise mean<strong>in</strong>g and one realizes<br />

immediately that the positive cone H + ω = L2 + (X, µω) isself-polar <strong>in</strong> the sense<br />

that a function g <strong>in</strong> L 2 (X, µω) is positive if and only if<br />

�<br />

X<br />

fgdµω ≥ 0<br />

for all functions f <strong>in</strong> the positive cone L2 + (X, µω).


170 F. Cipriani<br />

Re-writ<strong>in</strong>g the property above, us<strong>in</strong>g the scalar product of L 2 (X, µω) as<br />

follows<br />

L 2 + (X, µω) ={g ∈ L 2 (X, µω) :(f|g) ≥ 0 ∀f ∈ L 2 + (X, µω)} ,<br />

one realizes that self-polarity can be considered <strong>in</strong> arbitrary Hilbert spaces.<br />

Def<strong>in</strong>ition 2.2. (Self-polar cone) In a complex Hilbert space H, a subset<br />

H + ⊂His self-polar if it co<strong>in</strong>cides with its polar<br />

H + = {ξ ∈H:(ξ| η) ≥ 0 ∀ η ∈H + } . (2.2)<br />

In particular a self-polar set is a closed convex cone.<br />

Together with facial homogeneity and orientability, self-polarity is one of<br />

three fundamental properties characteriz<strong>in</strong>g von Neumann algebras <strong>in</strong> terms<br />

of ordered vector spaces [Co1]. Our <strong>in</strong>terest <strong>in</strong> it relies upon the fact that selfpolarity<br />

is the key property needed to treat Dirichlet forms and Markovian<br />

semigroups on general von Neumann algebras, avoid<strong>in</strong>g be<strong>in</strong>g limited to type<br />

I or type II von Neumann algebras and to trace functionals on them.<br />

A self-polar cone H + endows a Hilbert space H with an order structure which<br />

behaves, <strong>in</strong> some respect, similarly to the one of Example 2.1. In particular<br />

we will make use of the follow<strong>in</strong>g properties, referr<strong>in</strong>g to [Io] for the proof.<br />

Proposition 2.3. (Jordan decomposition <strong>in</strong> self-polar cones) Let H + be a<br />

self-polar cone <strong>in</strong> a complex Hilbert space H. Then<br />

i) H is the complexification H = HJ � iHJ of its real subspace<br />

H J := {ξ ∈H:(ξ| η) ∈ R ∀η ∈H + };<br />

ii) the cone H + gives rise to the structure of an ordered real vector space on<br />

H J and to an anti-unitary <strong>in</strong>volution<br />

J : H→H J(ξ + iη) :=ξ − iη , ξ , η ∈H J ;<br />

iii) any J-real element ξ ∈H J canbewrittenuniquelyasadifferenceξ =<br />

ξ+ − ξ− of two orthogonal, positive elements ξ± ∈H + . This decomposition<br />

characterizes self-dual cones among the closed and convex ones. It is usually<br />

referred as the Jordan decomposition.<br />

The positive part ξ+ of a J-real element ξ ∈ H J is, by def<strong>in</strong>ition, the<br />

Hilbertian projection of ξ onto the closed convex subset H + ⊂H J : ξ+ is<br />

the unique element m<strong>in</strong>imiz<strong>in</strong>g the distance between ξ and the positive cone<br />

H + . The negative part is def<strong>in</strong>ed by difference: ξ− := ξ+ − ξ.<br />

Example 2.4. (Jordan decomposition <strong>in</strong> commutative von Neumann algebras)<br />

In the commutative case of Example 2.1, the Jordan decomposition of a<br />

square summable, complex function f <strong>in</strong>to its real and imag<strong>in</strong>ary parts and


Dirichlet Forms on Noncommutative Spaces 171<br />

the splitt<strong>in</strong>g of these <strong>in</strong>to their positive and negative pieces, has the usual<br />

familiar mean<strong>in</strong>g. In this case, the positive part f+ of a square summable, real<br />

function f can be understood as the composition of f with the real variable<br />

function φ(x) :=x ∨ 0.<br />

Example 2.5. (Jordan decomposition <strong>in</strong> type I and type II von Neumann<br />

algebras) Beyond the commutative case, the choice of the closed convex cone<br />

M+ξω <strong>in</strong> the GNS Hilbert space Hω does not always match self-duality and<br />

it does if and only if the state ω is a trace:<br />

ω(ab) =ω(ba) a, b ∈ A. (2.3)<br />

If this is the case, an element xξω is J-real if and only if x = x ∗ ∈ M and<br />

its Jordan decomposition, given by xξω = x+ξω − x−ξω, essentially reduces<br />

to the Jordan decomposition x = x+ − x− <strong>in</strong> the von Neumann algebra M.<br />

Hence, even <strong>in</strong> noncommutative case, as long as the state ω is a trace, the<br />

order structure of the GNS space given by the cone M+ξω is essentially the<br />

order structure of the von Neumann algebra M.<br />

We will need the construction of a self-dual convex cone H + ω ,<strong>in</strong>thecase<br />

when the state ω is not necessarily a trace. This will be done with the help<br />

of the Tomita-Takesaki modular theory which we now briefly recall. The selfpolar<br />

cone H + ω will be, <strong>in</strong> a precise sense, <strong>in</strong>termediate among M+ ξω and<br />

M ′ + ξω, denot<strong>in</strong>gbyM ′ the commutant von Neumann algebra of M on H:<br />

M ′ = {x ∈ B(H) :xy = yx y ∈ B(H)} .<br />

To work <strong>in</strong> complete generality, we consider a von Neumann algebra of<br />

operators M, act<strong>in</strong>g on a Hilbert space H, andacyclic and separat<strong>in</strong>g vector<br />

ξ for it. The first property means that H = Mξ, while the second means that<br />

for x ∈ M, xξ =0ifandonlyifx =0.<br />

Equivalently, one may consider ξ to be the vector represent<strong>in</strong>g a faithful,<br />

normal state ω of M, <strong>in</strong>theGNSrepresentation,hav<strong>in</strong>g<strong>in</strong>m<strong>in</strong>d,asaprototype,<br />

the case where M is itself generated <strong>in</strong> a GNS representation of a given<br />

C∗-algebra A and a state ω on it.<br />

The isometric <strong>in</strong>volution map x ↦→ x∗ on M gives rise to the anti-l<strong>in</strong>ear map<br />

xξ ↦→ x∗ξ on H, densely def<strong>in</strong>ed on M ξ ⊆H. This map is isometric if and<br />

only if the vector state ω(·) =(ξ|·ξ) isatraceonM. The start<strong>in</strong>g observation<br />

of the Tomita-Takesaki modular theory is that this map is always closable. In<br />

the polar decomposition of its closure Sξ<br />

Sξ = Jξ∆ 1/2<br />

ξ<br />

(2.4)<br />

the anti-unitary part Jξ is called the modular conjugation while ∆ξ = S ∗ ξ Sξ<br />

is called the modular operator associated with the pair (M,ξ). The modular


172 F. Cipriani<br />

operator ∆ξ measures how the state ω differs from be<strong>in</strong>g a trace (<strong>in</strong> which<br />

case ∆ξ reduces to the identity).<br />

Theorem 2.6. (Tomita-Takesaki theorem) Let M be a von Neumann algebra,<br />

act<strong>in</strong>g on the Hilbert space H and ξ be a cyclic and separat<strong>in</strong>g vector.<br />

Let ∆ξ and Jξ be the associated modular operator and modular conjugation,<br />

respectively. It follows that<br />

JξMJξ = M ′<br />

(2.5)<br />

and that<br />

∆ it<br />

ξ M∆ −it<br />

ξ = M t ∈ R. (2.6)<br />

Now let M be a von Neumann algebra, ω a faithful normal state on it<br />

and (πω, Hω,ξω) the associated cyclic representation. Denot<strong>in</strong>g by ∆ω the<br />

modular operator associated to the pair (πω(M),ξω), the Tomita-Takesaki<br />

theorem allows one to construct one of the ma<strong>in</strong> tools of the theory.<br />

Def<strong>in</strong>ition 2.7. (Modular automorphisms group) The modular automorphism<br />

group σ ω : R → Aut(M) of the pair (M,ω) is def<strong>in</strong>ed through<br />

σ ω t<br />

(x) :=π−1<br />

ω (∆it ω<br />

−it<br />

πω(x)∆ω ) x ∈ M , t ∈ R . (2.7)<br />

This construction shows that the pair (M,ω) is a dynamical object. The<br />

celebrated Connes’ theorem of Radon-Nikodym type consists of a cocycle<br />

relation among the modular groups of different states on the same von Neumann<br />

algebra. A fundamental characterization of the modular group is the<br />

follow<strong>in</strong>g.<br />

Theorem 2.8. (Modular condition) The modular group σ ω : R → Aut(M)<br />

of the pair (M,ω) satisfies the follow<strong>in</strong>g properties (referred to as the modular<br />

condition):<br />

i) {σ ω t : t ∈ R} leaves the state ω <strong>in</strong>variant,<br />

ω = ω ◦ σ ω t t ∈ R (2.8)<br />

ii) for all x, y ∈ M there exists a bounded cont<strong>in</strong>uous function Fx,y on the<br />

closed strip D = {z ∈ C :0≤ Imz ≤ 1} which is holomorphic <strong>in</strong> the <strong>in</strong>terior<br />

D of D and satisfies<br />

F (t) =ω(σ ω t (x)y) , F(t + i) =ω(yσω t<br />

(x)) t ∈ R . (2.9)<br />

Moreover, given ω, the modular group is uniquely determ<strong>in</strong>ed by the above<br />

modular condition.<br />

The celebrated KMS-condition on one-parameter automorphism groups and<br />

states of C ∗ -algebras, deeply related to the above relation (2.9), plays a fundamental<br />

role <strong>in</strong> Quantum Statistical Mechanics <strong>in</strong> association with the ideas<br />

of time evolution and equilibrium. See Section 2.3 and Chapter 3.


Dirichlet Forms on Noncommutative Spaces 173<br />

Example 2.9. (Modular automorphisms groups on type I factors) When M =<br />

B(H) is the algebra of all bounded operators on a separable Hilbert space H,<br />

a normal state ω can always be represented as<br />

ω(x) =Trace(ρx) x ∈B(H) (2.10)<br />

for some positive, trace-class operator ρ ∈B(H), so that ω is faithful if and<br />

only if ρ is not s<strong>in</strong>gular. The modular automorphism group then reduces to<br />

σ ω t (x) =ρ it xρ −it for all x ∈B(H) andt ∈ R. In particular, if M = Mn(C),<br />

{ekl; k, l =1,...,n} is the system of matrix units and ρ is diagonal, one has<br />

σt(ekl) =<br />

� λk<br />

λl<br />

� it<br />

ekl<br />

k, l =1,...,n,<br />

where {λk : k =1,...,n} denote the list of eigenvalues of ρ.<br />

Example 2.10. (Modular automorphisms groups on Powers factors) Let Mn<br />

be a copy of M2(C) for all n ∈ N ∗ and consider the sequence of states ωn :<br />

Mn → C, correspond<strong>in</strong>g to a fixed λ ∈ (0, 1) and given by<br />

ωn([xij] 2 i,j=1 )=λx11 +(1− λ)x22<br />

Let A be the (unital) <strong>in</strong>ductive limit of the C ∗ -algebras<br />

An := M1 ⊗ M2 ⊗ ...Mn ,<br />

[xij] 2 i,j=1 ∈ Mn .<br />

obta<strong>in</strong>ed embedd<strong>in</strong>g An <strong>in</strong>to An+1 by a ↦→ a ⊗ 1. On the C ∗ -algebra A, a<br />

state ω is def<strong>in</strong>ed such that<br />

ω(x1 ⊗ ...xn ⊗ 1 ⊗ ...)=ω1(x1)ω2(x2) ...ωn(xn) n ∈ N ∗ ,<br />

so that we may consider the von Neumann algebra M generated by A <strong>in</strong> the<br />

GNS representation of ω. The modular automorphisms group of its normal<br />

extension is then given by<br />

σt(x1 ⊗ ...xn ⊗ 1 ⊗ ...)=σ 1 t (x1)σ 2 t (x2) ...σ n t (xn) n ∈ N ∗ ,<br />

where σ n t is the modular group of ωn on Mn = M2(C),described<strong>in</strong>the<br />

previous example.<br />

Let M be a von Neumann algebra of operators act<strong>in</strong>g on a Hilbert space H<br />

and let ξ ∈Hbe a fixed cyclic and separat<strong>in</strong>g for M. To def<strong>in</strong>e the standard<br />

cone <strong>in</strong> H, denotebyjξ : M → M ′ the anti-l<strong>in</strong>ear, <strong>in</strong>volutive isomorphism<br />

def<strong>in</strong>ed by the Tomita-Takesaki theorem:<br />

jξ(x) :=JξxJξ x ∈ M ⊆B(H) . (2.11)


174 F. Cipriani<br />

Notice that if ξ is a trace vector, <strong>in</strong> the sense that the associated functional<br />

ω(·) =(ξ| ·ξ) is a trace, this map co<strong>in</strong>cides with the <strong>in</strong>volution: jξ(x) =x ∗<br />

for all x ∈ M.<br />

Def<strong>in</strong>ition 2.11. (Standard cone of a von Neumann algebra) The standard<br />

associated with the pair (M,ξ) is def<strong>in</strong>ed as the closure of the set<br />

cone H +<br />

ξ<br />

{xjξ(x)ξ ∈H: x ∈ M} .<br />

Example 2.12. (Standard positive cone of f<strong>in</strong>ite von Neumann algebras) S<strong>in</strong>ce,<br />

by the Spectral Theorem, any positive element y ∈ M+ canbewrittenas<br />

y = xx∗ for some x ∈ M,<strong>in</strong>caseξis the trace vector represent<strong>in</strong>g a f<strong>in</strong>ite trace<br />

on M, as <strong>in</strong> Example 2.5, the cone H +<br />

ξ co<strong>in</strong>cides with M+ ξω. In particular, <strong>in</strong><br />

the commutative case of Example 2.1, the def<strong>in</strong>ition above allows the recovery<br />

of the fact that positive, square <strong>in</strong>tegrable functions can be approximated by<br />

positive, essentially bounded ones.<br />

The above example and the properties listed below suggest that the natural<br />

positive cone H +<br />

ξ should be considered as the analogue of the cone of positive<br />

L2-functions of the commutative case. We omit the proof and refer to<br />

[BR1 2.5.4].<br />

Proposition 2.13. Let M be a von Neumann algebra act<strong>in</strong>g on the Hilbert<br />

space H, withξ∈Ha cyclic and separat<strong>in</strong>g vector for M. The natural positive<br />

has the follow<strong>in</strong>g properties:<br />

cone H +<br />

ξ<br />

i) H +<br />

ξ<br />

is a self-dual set and <strong>in</strong> particular is a closed, convex cone <strong>in</strong> H;<br />

ii) H +<br />

ξ = ∆1/4<br />

ξ<br />

iii) Jξξ = ξ for all ξ ∈H +<br />

ξ ;<br />

ξ M+ξ = ∆ −1/4<br />

M′ + ξ<br />

iv) xjξ(x)H +<br />

ξ ⊆H+ ξ for all x ∈ M;<br />

v) ∆it ξ H+<br />

ξ = H+<br />

ξ for all t ∈ R;<br />

vi) f(log ∆ξ)H +<br />

ξ ⊆H+ ξ for all positive-def<strong>in</strong>ite, cont<strong>in</strong>uous functions f on R.<br />

Property i), together with Proposition 2.3, implies that the Jordan decomposition<br />

can be achieved <strong>in</strong> the ordered vector space (Hξ, H +<br />

ξ ), while property<br />

iii) reveals that Jξ co<strong>in</strong>cides with the anti-unitary <strong>in</strong>volution associated to the<br />

positive cone H +<br />

ξ , as <strong>in</strong> Proposition 2.3. Property iv) is a consequence of the<br />

very def<strong>in</strong>ition of H +<br />

ξ and the fact that for x, y ∈ M, x ∈ M and jξ(y) ∈ M ′<br />

necessarily commute. The identities <strong>in</strong> item ii) show that the natural positive<br />

cone is an <strong>in</strong>termediate deformation of the cones M+ ξω and M ′ + ξω. Property<br />

v) provides a first natural example of a one-parameter family of positive<br />

operators, i.e. operators leav<strong>in</strong>g globally <strong>in</strong>variant the natural positive cone.<br />

Maps and families of this type will be of <strong>in</strong>terest below. Property vi) follows<br />

from v) s<strong>in</strong>ce cont<strong>in</strong>uous, positive def<strong>in</strong>ite functions are, by a classical


Dirichlet Forms on Noncommutative Spaces 175<br />

result S. Bochner on the Fourier transform of f<strong>in</strong>ite, positive measures. Examples<br />

of cont<strong>in</strong>uous, positive def<strong>in</strong>ite functions are f t (x) =eitx for t ∈ R,<br />

f t (x) =e−tx2 for t>0andft (x) =e−t|x| for t>0.<br />

Remark 2.14. (Non self-polar cones) H. Araki studied <strong>in</strong> [Ara] properties<br />

analogous to those above, <strong>in</strong> particular generalizations of the Jordan decomposition,<br />

for the whole family of cones V α<br />

ξ = ∆α ξ M+ξ for α ∈ [0, 1/2].<br />

Despite the lack of all the symmetries the cone V 1/4<br />

ξ = H+<br />

ξ has, the positive<br />

cones correspond<strong>in</strong>g to α �= 1/4, although non self-polar, are important<br />

construct<strong>in</strong>g an <strong>in</strong>terpolat<strong>in</strong>g family of spaces Lp (M), 1 ≤ p ≤ +∞, which<br />

generalizes at the von Neumann algebra level the familiar family of Lebesgue<br />

spaces (see [Ko], [Te]).<br />

In the commutative case, equivalent measures µ, ν on a space X give rise to<br />

the same von Neumann algebra L ∞ (X, µ) =L ∞ (X, ν). We can look at this<br />

as two different representations of the von Neumann algebra of essentially<br />

bounded functions def<strong>in</strong>ed by the σ-algebra of sets of vanish<strong>in</strong>g measure. In<br />

a precise sense this holds true for general von Neumann algebras and to state<br />

it correctly we <strong>in</strong>troduce the follow<strong>in</strong>g def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 2.15. (Standard form of a von Neumann algebra) A standard<br />

form (M, H, H + ,J) of a von Neumann algebra M act<strong>in</strong>g on a Hilbert space<br />

H consists of a self-dual cone H + and an anti-l<strong>in</strong>ear <strong>in</strong>volution J, satisfy<strong>in</strong>g:<br />

i) JMJ = M ′ ;<br />

ii) JxJ = x∗ ∀x ∈ M ∩ M ′ (the center of M);<br />

iii) Jη = η ∀η ∈H + ;<br />

iv) xJxJ(H + ) ⊆H + ∀x ∈ M .<br />

In the commutative case, <strong>in</strong> which the von Neumann algebra is described as<br />

L ∞ (X, µ) for some f<strong>in</strong>ite measure µ on X, all standard forms appear as<br />

(L ∞ (X, µ ),L 2 (X, ν) ,L 2 +(X, ν) ,J)<br />

for some measure ν equivalent to µ. Here the antil<strong>in</strong>ear <strong>in</strong>volution is just<br />

complex conjugation: Ja(x) =a(x), ν-a.e. on X.<br />

Proposition 2.13 tells us that modular theory allows the constructionof standard<br />

forms of von Neumann algebras start<strong>in</strong>g from faithful, normal state on<br />

them. As <strong>in</strong> the commutative case, we are go<strong>in</strong>g to see that all standard forms<br />

look essentially the same.<br />

Proposition 2.16. (Uniqueness of standard forms of von Neumann algebras)<br />

Let (Mi, Hi, H + i ,Ji), i =1, 2, be standard forms of two von Neumann<br />

algebras M1 , M2. Ifα : M1 → M2 is an isomorphism of M1 onto M2 then<br />

there exists a unique unitary U : H1 →H2 such that:


176 F. Cipriani<br />

i) α(x) =UxU∗ x ∈ M1;<br />

ii) J2 = UJ1U∗ ;<br />

iii) H + 2 = UH+ 1 .<br />

iv) Moreover, when H1 = H2, themapU↦→ αU (·) :=U · U ∗ establishes a<br />

topological isomorphism between the subgroup<br />

{U ∈U(H) :UMU ∗ = M, UJU ∗ = J, UH + = H + }<br />

of the group U(H) of all unitary operators on H and the group Aut(M) of all<br />

automorphisms of M.<br />

Hence, <strong>in</strong> the above sense, all standard forms of a von Neumann algebra are<br />

unitarily equivalent; a unified notation for them is (M,L2 (M),L2 + (M),J), as<br />

justified by the properties listed <strong>in</strong> Proposition 2.3, Proposition 2.13 and by<br />

the consideration of the commutative case discussed above.<br />

We already met an example of property iv) <strong>in</strong> the above proposition, by which<br />

any automorphism, or group of them, can always be unitarily implemented<br />

<strong>in</strong> any standard form: namely the modular group of a faithful normal state<br />

(see Def<strong>in</strong>ition 2.7 and Proposition 2.13 v)).<br />

A practical consequence of the uniqueness of a standard form is that <strong>in</strong> work<strong>in</strong>g<br />

with them we have the freedom to choose the most suitable for the problem<br />

at hand. Here are some illustrations of this.<br />

Example 2.17. (Hilbert-Schmidt standard form of type I factors) Let M =<br />

B(H) be the type I factor of all bounded operator on a Hilbert space H.<br />

In the Hilbert space L 2 (H) of all Hilbert-Schmidt operators on H, the cone<br />

L 2 +(H) of the positive ones is self-dual. The related <strong>in</strong>volution J associates<br />

to the Hilbert-Schmidt operator ξ its adjo<strong>in</strong>t ξ ∗ .Moreover<br />

(B(H) , L 2 (H) , L 2 + (H) ,J)<br />

is a standard form of B(H). This standard form may be constructed by<br />

an extension of the Tomita-Takesaki theory which deal with faithful, semif<strong>in</strong>ite,<br />

normal weights on von Neumann algebras, as <strong>in</strong>deed is the trace Tr :<br />

B(H)+ → [0, +∞] on the algebra of all bounded operators. Notice that, under<br />

the identification L 2 (H) =H ⊗ H, the positive cone L 2 +<br />

(H) is generated by<br />

the elementary elements of the form x ⊗ x, x ∈ H while the <strong>in</strong>volution acts<br />

as J(x ⊗ y) =y ⊗ x, x, y ∈ H. HereH denotes the opposite Hilbert space and<br />

x,y its elements. In terms of the Dirac’s bra-ket notation: x ⊗ y = |x〉〈y|.<br />

Example 2.18. (Generalized Hilbert-Schmidt standard form) The construction<br />

above of standard forms of Type I factors, may be generalized to any<br />

von Neumann algebra us<strong>in</strong>g the relative tensor product of Hilbert modules<br />

(see [S1]).


Dirichlet Forms on Noncommutative Spaces 177<br />

Example 2.19. (Modular operators on type I factors) Consider on the above<br />

type I factor B(H), the normal state ω(x) :=Tr(ρx) associated to a positive,<br />

trace-class operator ρ ∈B(H). By Proposition 2.16 the standard form of the<br />

GNS representation of ω is isomorphic the Hilbert-Schmidt standard form<br />

considered <strong>in</strong> Example 2.17. We may then look for the realizations of the<br />

modular operators <strong>in</strong> this representation. Clearly the modular conjugation<br />

co<strong>in</strong>cides with the operation of tak<strong>in</strong>g the adjo<strong>in</strong>t, as it depends upon the<br />

positive cone only, by Proposition 2.3. The state ω can be represented as<br />

ω(x) =(ξρ|xξρ) L2 (H) for a unique positive Hilbert-Schmidt operator ξρ ∈<br />

L 2 +<br />

(H) which then represents the cyclic and separat<strong>in</strong>g vector associated to<br />

ω. S<strong>in</strong>ce this means that Tr(ρx) =Tr(ξρxξρ) for all Hilbert-Schmidt operators<br />

x on H, one identifies ξρ = ρ 1/2 . To recover the action of the modular operator<br />

notice that, by def<strong>in</strong>ition,<br />

J∆ 1/2 (xξρ) =x ∗ ξρ , x ∈B(H) .<br />

Then ∆ 1/2 (xρ 1/2 )=ρ 1/2 x for all x ∈B(H) sothat<br />

∆ 1/2 ξ = ρ 1/2 ξρ −1/2<br />

for all ξ ∈ D(∆ 1/2 ):={η ∈L 2 (H) :ρ 1/2 ηρ −1/2 ∈L 2 (H)}. One may check<br />

that the ideal of f<strong>in</strong>ite rank operators is an operator core for ∆ 1/2 . This result<br />

then makes it easy to identify the action of the modular group:<br />

σ ω t (x) =ρ it xρ −it<br />

x ∈B(H) , t ∈ R .<br />

An important example of standard form <strong>in</strong> harmonic analysis is the<br />

follow<strong>in</strong>g one.<br />

Example 2.20. (Standard forms of group von Neumann algebras and the<br />

Fourier transform) Consider a unimodular, locally compact group G and denote<br />

by s,t... its elements and by ds,dt... a fixed Haar measure on it. On<br />

the Hilbert space L2 (G) theleft regular representation λ of G is def<strong>in</strong>ed by<br />

λ : G →B(L 2 (G)) , (λ(s)f)(t) =f(s −1 t) s, t ∈ G, f ∈ L 2 (G) .<br />

The associated von Neumann algebra is, by def<strong>in</strong>ition, the weak∗-closure <strong>in</strong><br />

B(L2 (G)) of the group of unitaries λ(G). By von Neumann’s density theorem<br />

[Dix2] it co<strong>in</strong>cides both with the double commutant λ(G) ′′ of λ(G) andwith<br />

the weak∗-closure of the algebra L1 (G) act<strong>in</strong>gonL2 (G) byleftconvolution<br />

�<br />

(f ∗ g)(t) :=<br />

f(s)g(s<br />

G<br />

−1 t)ds t ∈ G, f ∈ L 1 (G) , g ∈ L 2 (G) .<br />

A standard form of λ(G) ′′ is def<strong>in</strong>ed <strong>in</strong> the Hilbert space L2 (G) by consider<strong>in</strong>g<br />

the <strong>in</strong>volution<br />

Jf(s) = f(s −1 ) , s ∈ G, f ∈ L 2 (G)


178 F. Cipriani<br />

and the self-polar cone<br />

H +<br />

G = {f ∗ Jf : f ∈ Cc(G)} ,<br />

where Cc(G) denote the space of compactly supported cont<strong>in</strong>uous functions<br />

on G. H +<br />

G co<strong>in</strong>cides with the cone of positive def<strong>in</strong>ite, square summable functions<br />

on G. These properties are proved <strong>in</strong> [Dix2]. Their generalization to a<br />

non unimodular group can be found <strong>in</strong> [Pen].<br />

When G is commutative, the von Neumann algebra λ(G) ′′ is isomorphic to<br />

the abelian von Neumann algebra L∞ ( � G) of essentially bounded functions on<br />

the Pontriagyn dual � G, with respect to any of its Haar measures. The Fourier<br />

transform F : L2 (G) −→ L2 ( � G), is then a unitary operator with the properties<br />

i), ii) and iii) of Proposition 2.16, which implements that isomorphism.<br />

2.2 Order Structures and Symmetric Embedd<strong>in</strong>gs<br />

The positive cones M+, H + and M∗+ endow the von Neumann algebra M,<br />

the standard Hilbert space H and the pre-dual space M∗ with structures of<br />

ordered vector spaces. More precisely,<br />

�<br />

what we get is a complex vector space<br />

E, complexification E = ER ER of a real subspace ER, which <strong>in</strong> turn is<br />

l<strong>in</strong>early generated, over R, by the cone E+: ER = E+ − E+.<br />

The order relation between real elements elements e1,e2 of an ordered vector<br />

space (E,E+) will be <strong>in</strong>dicated as customary: e1 ≤ e2 will mean that e2−e1 ∈<br />

E+. The elements of E+ are exactly those real vectors e <strong>in</strong> E such that 0 ≤ e.<br />

The order <strong>in</strong>terval [e1,e2] ⊂ E, def<strong>in</strong>ed by two real elements e1, e2 of E such<br />

that e1 ≤ e2, is def<strong>in</strong>ed by<br />

[e1,e2] :={e ∈ E : e1 ≤ e ≤ e2} .<br />

For example, the order <strong>in</strong>terval [0, 1] <strong>in</strong> the abelian von Neumann algebra<br />

L ∞ (X, m) (resp. <strong>in</strong> the standard Hilbert space L 2 (X, m)) consists of those<br />

real valued functions a, m-a.e. def<strong>in</strong>ed on X (resp. belong<strong>in</strong>g to L 2 (X, m))<br />

such that 0 ≤ a(x) ≤ 1form-almost all x ∈ X.<br />

Complex l<strong>in</strong>ear maps L : E −→ F between complex ordered vector spaces<br />

(E,E+)and(F, F+), will be called real if L(ER) ⊆ FR and positive if moreover<br />

L(E+) =F+.<br />

The tools we need to compare maps and semigroups on von Neumann algebras,<br />

their standard Hilbert spaces and their pre-dual spaces, are the symmetric<br />

embedd<strong>in</strong>gs. These embedd<strong>in</strong>gs, for example, make it possible to recognize<br />

the elements of the von Neumann algebra M with<strong>in</strong> the standard Hilbert<br />

space H, <strong>in</strong> such a way that an element is positive <strong>in</strong> M if and only if is positive<br />

<strong>in</strong> H. These embedd<strong>in</strong>gs generalize the commutative situation <strong>in</strong> which,


Dirichlet Forms on Noncommutative Spaces 179<br />

for a f<strong>in</strong>ite measure m, one has<br />

L ∞ (X, m) ⊆ L 2 (X, m) ⊆ L 1 (X, m) .<br />

Let (M, H, H + ,J)beastandard form of a von Neumann algebra M, with<br />

J0 ,∆0 the modular operators associated to a cyclic and separat<strong>in</strong>g vector<br />

ξ0 ∈H + and ω0(·) =(ξ0|·ξ0) the correspond<strong>in</strong>g normal state.<br />

When M is commutative, or at least when ξ0 is a trace vector, the embedd<strong>in</strong>g<br />

x ↦→ xξ0 of the von Neumann algebra M <strong>in</strong>to the Hilbert space H is <strong>in</strong>jective,<br />

cont<strong>in</strong>uous with norm dense range and positivity preserv<strong>in</strong>g. Incaseξ0is not<br />

a trace vector, a “distortion” of the above embedd<strong>in</strong>g makes it possible to<br />

keep all these properties throughout.<br />

Def<strong>in</strong>ition 2.21. (Symmetric embedd<strong>in</strong>gs) The symmetric embedd<strong>in</strong>gs associated<br />

to the standard form (M, H, H + ,J) and a cyclic and separat<strong>in</strong>g<br />

vector ξ0 ∈H + are def<strong>in</strong>ed as follows:<br />

i) i0 : M →H i0(x) :=∆ 1/4<br />

0 xξ0 x ∈ M;<br />

ii) i0∗ : H→M∗ 〈i0∗(ξ),y〉 =(i0(y∗ )| ξ) =(∆ 1/4<br />

0 y∗ξ0| ξ) ξ ∈H,y ∈ M;<br />

iii) j0 : M → M∗ 〈j0(x),y〉 =(J0yξ0|xξ0) x, y ∈ M.<br />

These maps are well def<strong>in</strong>ed because Mξ0 ⊆ D(∆ 1/2<br />

0 ), by the def<strong>in</strong>ition of<br />

the modular operator, and because D(∆ 1/2<br />

0 ) ⊆ D(∆ 1/4<br />

0 ) by the Spectral<br />

Theorem.<br />

In the follow<strong>in</strong>g we consider the spaces M , H , M∗ ordered by their positive<br />

cones M+ , H + , M∗+. We will denote by σ(M, M∗) theweak∗-topology of M, byσ(M∗, M) the weak-topology of M∗ and by σ(H, H) theweaktopology<br />

of H.<br />

Proposition 2.22. The above maps are bounded, positivity preserv<strong>in</strong>g <strong>in</strong>jections<br />

with norm dense range. Moreover we have<br />

i) i0 is the dual map of i0∗ and j0 = i0∗ ◦ i0;<br />

ii) i0 is an order isomorphism between the order-<strong>in</strong>tervals [0, 1A] <strong>in</strong> M and<br />

[0,ξ0] <strong>in</strong> H;<br />

iii) i0∗ is an order isomorphism between the order-<strong>in</strong>tervals [0,ξ0] <strong>in</strong> H and<br />

[0,ω0] <strong>in</strong> M∗;<br />

iv) the map i0 is σ(M, M∗) − σ(H, H)-cont<strong>in</strong>uous<br />

v) the map i0∗ is σ(H, H) − σ(M∗, M)-cont<strong>in</strong>uous<br />

vi) the map j0 is σ(M, M∗) − σ(M∗, M)-cont<strong>in</strong>uous.<br />

In particular, item ii) tells us how to identify the algebra M <strong>in</strong> the Hilbert<br />

space H: for example, an element ξ ∈H + is of the form ∆ 1/4<br />

0 xξ0 for some


180 F. Cipriani<br />

x ≥ 0 if and only if 0 ≤ ξ ≤ kξ0 for some k ≥ 0. Moreover, the algebra norm<br />

can be read from the order structure of the positive cone:<br />

�x�M =<strong>in</strong>f{k ∈ [0, +∞) :0≤ ξ ≤ kξ0} .<br />

Example 2.23. (Symmetric embedd<strong>in</strong>g on type I factors) Consider on the type<br />

IfactorB(H), the normal state ω(x) :=Tr(ρx) associated to a positive, traceclass<br />

operator ρ ∈B(H) and the Hilbert-Schmidt standard form considered<br />

<strong>in</strong> Example 2.19. It is easy to verify that ∆ 1/4 ξ = ρ 1/4 ξρ −1/4 for all ξ ∈<br />

D(∆ 1/4 ) so that, s<strong>in</strong>ce the cyclic vector is ξρ = ρ 1/2 ,wehave∆ 1/4 (xξρ) =<br />

∆ 1/4 (xρ 1/2 )=ρ 1/4 xρ 1/4 for all x ∈B(H). It is then clear why the symmetric<br />

embedd<strong>in</strong>g i0 carries a positive element x of the von Neumann algebra <strong>in</strong>to<br />

a positive element i0(x) of the standard Hilbert space. The isomorphism <strong>in</strong><br />

item ii) of the above proposition reduces, <strong>in</strong> this case, to the fact that for<br />

a positive bounded operator x on H and some k ≥ 0, ρ 1/4 xρ 1/4 ≤ kρ 1/2 is<br />

equivalent to x ≤ k · 1 B(H).<br />

S<strong>in</strong>ce Mξ0 ⊆ D(∆ 1/2<br />

0 ) ⊆ D(∆ 1/4<br />

0 ), for any x ∈ M we have<br />

�i0(x)� 2 H =(∆ 1/4<br />

0 xξ0| ∆ 1/4<br />

0 xξ0)H =(∆ 1/2<br />

0 xξ0| xξ0)H =(J0x ∗ ξ0| xξ0)H<br />

= 〈j0(x),x〉≤�x ∗ �M ·�j0(x)�M∗ = �x�M ·�j0(x)�M∗ .<br />

This may suggest that these embedd<strong>in</strong>gs form the skeleton of a symmetric,<br />

complex <strong>in</strong>terpolation system among the spaces M, H, M∗. Thisis<strong>in</strong>fact<br />

the case and it is the start<strong>in</strong>g po<strong>in</strong>t for the construction of noncommutative<br />

L p -spaces (see for example [H2], [Ko], [Te]). We refer to Sections 5 and 6<br />

of Quanhua Xu’s Lectures <strong>in</strong> this volume for a detailed exposition of this<br />

matter. In particular we extrapolate the follow<strong>in</strong>g result from [Te], which we<br />

will need later on.<br />

Proposition 2.24. (Noncommutative <strong>in</strong>terpolation) Let (M, H, H + ,J) be a<br />

standard form of a von Neumann algebra M, ξ0 ∈H + a cyclic and separat<strong>in</strong>g<br />

vector and i0, i0∗ and j0 the associated symmetric embedd<strong>in</strong>gs of M <strong>in</strong>to H,<br />

H <strong>in</strong>to M∗ and M <strong>in</strong>to M∗, respectively.<br />

Let T1 : M∗ → M∗ be a map such that<br />

T1(j0(M)) ⊆ j0(M) , T1(i0∗(H)) ⊆ i0∗(H).<br />

Denote by T∞ : M → M and T2 : H→Hbe the maps determ<strong>in</strong>ed by<br />

T1 ◦ j0 = j0 ◦ T∞ , T1 ◦ i0∗ = i0∗ ◦ T2 .<br />

Then, if T1 and T∞ are bounded with norm M1 and M∞, respectively, then<br />

T2 is bounded with norm less than M 1/2<br />

1<br />

· M 1/2<br />

∞<br />

�T2ξ� 2 H ≤ M1 · M∞�ξ�| 2 H ξ ∈H. (2.12)<br />

In particular, if T1 and T∞ are contractive then T2 is contractive too.


Dirichlet Forms on Noncommutative Spaces 181<br />

2.3 KMS-symmetric Maps and Semigroups<br />

on C ∗ and von Neumann Algebras<br />

In this section we <strong>in</strong>troduce one of the ma<strong>in</strong> object of our <strong>in</strong>vestigation: KMSsymmetric<br />

semigroups on C ∗ and von Neumann algebras and, <strong>in</strong> particular,<br />

KMS-symmetric, positive and Markovian semigroups. In the von Neumann<br />

algebra case and with respect to the modular automorphism group, this class<br />

of semigroups was <strong>in</strong>troduced <strong>in</strong> [Cip1], [GL1,2]. KMS-symmetric positive and<br />

Markovian semigroups were <strong>in</strong>troduced <strong>in</strong> [Cip2]. The consideration of KMSsymmetric<br />

semigroups on C ∗ algebras, which are not necessarily positive or<br />

Markovian, is a novelty.<br />

Markovian semigroups, symmetric with respect to traces, were <strong>in</strong>troduced<br />

by L. Gross [G1],[G2] to study physical Hamiltonians of Boson and Fermion<br />

systems. A general theory of Dirichlet forms on algebras endowed with reasonable<br />

traces was developed by S.Albeverio and R.Hoegh-Krohn [AHK1,2] and<br />

later studied by J.M.L<strong>in</strong>dsay and E.B.Davies [DL1,2] and by J.L.Sauvageot<br />

[S2,3].<br />

In a commutative situation, a map Φ, or a semigroup of them, on the<br />

C ∗ -algebra C0(X) can be extended to the von Neumann algebra L ∞ (X, m),<br />

its predual L 1 (X, m) and the standard Hilbert space L 2 (X, m), for any f<strong>in</strong>ite<br />

Radon measure m on X with respect to which Φ is symmetric:<br />

�<br />

�<br />

Φ(f) · gdm= f · Φ(g) dm f, g ∈ C0(X) .<br />

X<br />

X<br />

These extensions, considered ma<strong>in</strong>ly for positive, or Markovian, maps and<br />

semigroups, are of great value for construct<strong>in</strong>g and study<strong>in</strong>g symmetric<br />

Markov processes (see [FOT]) and <strong>in</strong>teract<strong>in</strong>g particle systems (see [Lig]).<br />

In this section we show how such an extension program can be carried out<br />

on any C∗-algebra, provided a suitable symmetry holds true with respect<br />

to KMS-states (see [BR2]). States <strong>in</strong> this class, <strong>in</strong>troduced by the physicists<br />

R. Kubo [Kub], D.C. Mart<strong>in</strong> and J. Schw<strong>in</strong>ger [MS], and identified mathematically<br />

by R. Haag, N.M. Hugenoltz and M. W<strong>in</strong>n<strong>in</strong>k [HHW], are fundamental<br />

for represent<strong>in</strong>g equilibria of quantum systems of <strong>in</strong>f<strong>in</strong>ite degrees of freedom<br />

and thermal reservoirs (see [BR2], [Sew], [KFGV]). The symmetry condition<br />

we <strong>in</strong>troduce is a generalization of the detailed balance condition (see [Al],<br />

[D1] and [KFGV]).<br />

Unless otherwise stated, maps (Φ,D(Φ)) on a C ∗ -algebra A will always be<br />

considered real, <strong>in</strong> the sense that their doma<strong>in</strong> D(Φ) is an <strong>in</strong>volutive subspace<br />

of A and Φ(a ∗ ) = Φ(a) ∗ for all a <strong>in</strong> D(Φ). For simplicity, we will work<br />

with unital C ∗ -algebras. Otherwise, one may add a unit to A or consider A<br />

embedded <strong>in</strong>to its (unital) envelop<strong>in</strong>g von Neumann algebra A ∗∗ .


182 F. Cipriani<br />

Def<strong>in</strong>ition 2.25. (Positive and Markovian semigroups)AmapΦ : A →<br />

A on a C∗-algebra A is said to be<br />

i) positive if Φ(A+) ⊆ A+;<br />

ii) Markovian if it is positive and Φ(a) ≤ 1A whenever a = a∗ and a ≤ 1A;<br />

iii) completely positive (resp. completely Markovian), if for every n ∈ N∗ the<br />

matricial extension<br />

Φ ⊗ In :Mn(A) → Mn(A) (Φ ⊗ In)[aij] n i,j=1 := [Φ(aij)] n i,j=1<br />

is positive (resp. Markovian) on the C∗-algebra Mn(A) = A ⊗ Mn(C) of<br />

n × n-matrices with entries <strong>in</strong> A.<br />

iv) A one-parameter semigroup {Φt : t ≥ 0} on A is said to be positive<br />

(resp. completely positive, Markovian, completely Markovian) providedΦtis positive (resp completely positive, Markovian, completely Markovian) for all<br />

t ≥ 0.<br />

S<strong>in</strong>ce an�element a of A is <strong>in</strong> the closed unit ball (�a� ≤1) if and only if the<br />

1A a<br />

element<br />

a∗ �<br />

is positive <strong>in</strong> M2(A), one sees that completely Markovian<br />

1A<br />

semigroups are just completely positive contraction semigroups. Thisclassof<br />

semigroups, often called dynamical semigroups <strong>in</strong> the literature, is of special<br />

importance for application to Quantum Open Systems [D1] and Quantum<br />

<strong>Probability</strong>. We refer to Philippe Biane’s Lectures <strong>in</strong> this volume for the important<br />

connection between Markovian semigroups and stochastic processes<br />

<strong>in</strong> noncommutative spaces.<br />

The structure of generators of σ-weakly cont<strong>in</strong>uous, completely Markovian<br />

semigroups on von Neumann algebras, symmetric with respect to traces will<br />

be carefully studied <strong>in</strong> Chapter 4.<br />

Example 2.26. (Structure of norm cont<strong>in</strong>uous, completely positive semigroups)<br />

Uniform cont<strong>in</strong>uity forces the semigroup to have a bounded generator L :<br />

A → A (see [BR1 Proposition 3.1.1]):<br />

Φt = �<br />

n≥0<br />

t n<br />

n! Ln , L n (a) :=L(L(...L(a))) , a ∈ A.<br />

Suppose A is act<strong>in</strong>g on the Hilbert space h and {Φt : t ≥ 0} is norm cont<strong>in</strong>uous<br />

and completely positive. Then, by a theorem due G. L<strong>in</strong>dblad [L<strong>in</strong>]<br />

and V. Gor<strong>in</strong>i-A. Kossakowsi-E.C.G. Sudarsan [GKS], for matrices, and to<br />

E. Christensen-D. E. Evans [CE] for general C ∗ -algebras, the generator L<br />

may be represented as<br />

L(a) =i[H, a]+(Ra + aR) − φ(a) ,


Dirichlet Forms on Noncommutative Spaces 183<br />

where H = H∗ is a self-adjo<strong>in</strong>t operator on h belong<strong>in</strong>g to the σ-weak<br />

closure M ⊆ B(h) ofA, φ : A → M is a completely positive map and<br />

R = L(1A) − φ(1A). Moreover, comb<strong>in</strong><strong>in</strong>g this result with the fundamental<br />

St<strong>in</strong>espr<strong>in</strong>g theorem [Sti] (see Appendix 7.3), one obta<strong>in</strong>s the general form<br />

of the generator of a norm cont<strong>in</strong>uous, σ-weakly cont<strong>in</strong>uous, completely positive,<br />

identity preserv<strong>in</strong>g semigroup on a von Neumann algebra M ⊆ B(h):<br />

L(a) =i[H, a]+ �<br />

k∈I<br />

y ∗ k yka − 2y ∗ k ayk + ay ∗ k yk<br />

where {yk ∈ B(h) :k ∈ I} is a family such that the operator �<br />

k∈I y∗ kyk is bounded and �<br />

k∈I y∗ kayk ∈ M for all a ∈ M. Operators L of the above<br />

type are called L<strong>in</strong>dblad generators. The structure of generators of strongly<br />

cont<strong>in</strong>uous Markovian semigroups which are symmetric with respect to a<br />

semif<strong>in</strong>ite faithful trace will be <strong>in</strong>vestigated <strong>in</strong> Chapter 4.<br />

To <strong>in</strong>troduce the second ma<strong>in</strong> <strong>in</strong>gredient of this section, we recall that, for<br />

a strongly (resp. σ-weakly) cont<strong>in</strong>uous group of automorphisms α := {αt :<br />

t ∈ R} of a C∗ (resp. von Neumann) algebra A, we denote by Aα the norm<br />

(resp. σ-weakly) dense, α-<strong>in</strong>variant ∗-subalgebra of α-analytic elements of A<br />

(see Appendix 7.3).<br />

On a f<strong>in</strong>ite dimensional C∗-algebra, say A = Mn(C), any automorphism<br />

group appears as<br />

a ∈ A<br />

αt(a) =e itH ae −itH<br />

for a self-adjo<strong>in</strong>t matrix H. It is easy to see, us<strong>in</strong>g the cyclicity of the trace,<br />

that the state<br />

ω(a) =Trace(e −βH a)/Trace(e −βH )<br />

satisfies the relation<br />

ω(ba) =ω(aαiβ(b)) a, b ∈ A.<br />

Moreover, ω is the unique state with this property and it may be thought of<br />

as the equilibrium state of a f<strong>in</strong>ite quantum system, whose time evolution is<br />

governed by the automorphism group.<br />

States of the above type still exist <strong>in</strong> <strong>in</strong>f<strong>in</strong>ite dimensional situations, where,<br />

for example, A may be the C ∗ -algebra K(h) of all compact operators on an<br />

<strong>in</strong>f<strong>in</strong>ite dimensional Hilbert space h. This, however, forces the dynamics to be<br />

generated by a Hamiltonian H with discrete spectrum. To avoid this restriction,<br />

one may considers the above relation between ω and α as the start<strong>in</strong>g<br />

po<strong>in</strong>t for isolat<strong>in</strong>g a rich class of states. The only drawback <strong>in</strong>volved <strong>in</strong> this<br />

po<strong>in</strong>t of view is that one has to abandon the idea of represent<strong>in</strong>g states <strong>in</strong> this<br />

class <strong>in</strong> the above simple way, i.e. by the density matrix e −βH /Trace(e −βH ).<br />

Def<strong>in</strong>ition 2.27. ( KMS-states) Letα := {αt : t ∈ R} be a strongly cont<strong>in</strong>uous<br />

group of automorphisms of a C ∗ -algebra A and β ∈ R. Astateω is


184 F. Cipriani<br />

said to be a (α, β)-KMS state if it is α-<strong>in</strong>variant and if the follow<strong>in</strong>g KMScondition<br />

holds true:<br />

ω(aαiβ(b)) = ω(ba) (2.13)<br />

for all a, b <strong>in</strong> a norm dense, α-<strong>in</strong>variant ∗-subalgebra of Aα. IfAis a von<br />

Neumann algebra and α := {αt : t ∈ R} is a σ(M, M∗)-cont<strong>in</strong>uous group of<br />

automorphisms, a state ω is said to be a (α, β)-KMS state if ω is α-<strong>in</strong>variant,<br />

normal and the KMS-condition above holds true for all a, b <strong>in</strong> a σ(M, M∗)dense,<br />

α-<strong>in</strong>variant ∗-subalgebra of Aα. KMS states correspond<strong>in</strong>g to β =0<br />

are just the traces over A.<br />

These states describe equilibria of quantum statistical systems, whose physical<br />

observables are represented by elements of A, correspond<strong>in</strong>g to the time<br />

evolution represented by the automorphisms group {αt : t ∈ R} (see [BR2]).<br />

The parameter β, sometimes called the <strong>in</strong>verse temerature, is proportional<br />

to h/2πT, whereTis the absolute temperature and h Planck’s constant.<br />

The KMS-condition above can be restated <strong>in</strong> a form similar to the modular<br />

condition <strong>in</strong> Theorem 2.8, <strong>in</strong>dependent of the consideration of a norm dense,<br />

α-<strong>in</strong>variant ∗-subalgebras of Aα.<br />

Proposition 2.28. ( KMS condition)([BR2 Proposition 5.3.7]) Let α :=<br />

{αt : t ∈ R} be a strongly (resp. σ-weakly) cont<strong>in</strong>uous group of automorphisms<br />

of a C∗ (resp. von Neumann) algebra A, β ∈ R and ω a state on A (normal<br />

<strong>in</strong> the von Neumann algebra case).<br />

Def<strong>in</strong>e Dβ to be the doma<strong>in</strong> {z ∈ C :0< Imz


Dirichlet Forms on Noncommutative Spaces 185<br />

C ∗ -algebra U(h) generated by a family {a(f) ∈ B(h) :f ∈ h0} of bounded<br />

operators satisfy<strong>in</strong>g<br />

i) h0 ∋ f → a(f) is antil<strong>in</strong>ear;<br />

ii) a(f)a(g)+a(g)a(f) =0, f,g ∈ h0 ;<br />

iii) a(f)a(g) ∗ + a(g) ∗ a(f) =(f|g) · Ih , f,g ∈ h0 .<br />

The C ∗ -algebra U(h) is called the CAR algebra over the Hilbert space h or<br />

the algebra of canonical anti-commutation relations. It represents the algebra<br />

of physical observables of quantum mechanical systems obey<strong>in</strong>g Fermi-Dirac<br />

statistics. Given a self-adjo<strong>in</strong>t operator H on h, themapα : R → Aut(U(h))<br />

def<strong>in</strong>ed by<br />

αt(a(f)) := a(e itH f) , αt(a ∗ (f)) := a ∗ (e itH f) f ∈ h0<br />

def<strong>in</strong>es a strongly cont<strong>in</strong>uous automorphism group of the CAR algebra, called<br />

Bogoliubov transformations (see [BR2]). Clearly, if h0 is a dense set of analytic<br />

vectors for H then the ∗ -algebra generated by the family {a(f) ∈ B(h) :<br />

f ∈ h0} is a dense, α-<strong>in</strong>variant ∗ -sub-algebra of U(h) consist<strong>in</strong>g of analytic<br />

vectors for the automorphisms group. For a fixed β ∈ R, the(α,β)-KMScondition<br />

for a given state ω requires, <strong>in</strong> particular, that<br />

ω(a ∗ (f)a(g)) = ω(a(g)a ∗ (e −βH f)) f,g ∈ h0 .<br />

S<strong>in</strong>ce, by the canonical anti-commutation relations,<br />

a(g)a ∗�<br />

e −βH �<br />

f<br />

we should have<br />

and then<br />

ω(a ∗ (f)a(g)) =<br />

=(g|e −βH f) · Ih − a ∗ (e −βH f)a(g) ,<br />

ω(a ∗ ((Ih + e −βH )f)a(g)) = (g|e −βH f)<br />

�<br />

g| e −βH (I + e −βH ) −1 �<br />

f<br />

f,g ∈ h0 .<br />

This formula determ<strong>in</strong>es, through the CAR relations, a state ωβ over U(h)<br />

which is called the gauge-<strong>in</strong>variant, quasi-free state associated to H and β.<br />

It is the unique (α,β)-KMS state over U(h) (see [BR2 Section 5.3]).<br />

Into this framework falls the “Ideal Fermi Gas” where h := L 2 (R d , dx), h0 is<br />

the space of rapidly decreas<strong>in</strong>g functions, H := −H0 − µ · I, H0 denot<strong>in</strong>g the<br />

Laplacian operator on R d ,andµ ∈ R.<br />

Example 2.30. (KMS-states on CCR C ∗ -algebras) Given a complex pre-<br />

Hilbert space h0 with Hilbert space completion h, there exists a unique up to<br />

∗ -isomorphism, C ∗ -algebra U(h0) generated by a family {W (f) ∈ B(h) :f ∈<br />

h0} of unitary operators satisfy<strong>in</strong>g


186 F. Cipriani<br />

i) W (−f) =W (f) ∗ f ∈ h0 ;<br />

ii) W (f)W (g) =e −iIm(f|g) W (g)W (f) f,g ∈ h0 .<br />

This is Weyl’s form of the canonical commutation relations CCR, describ<strong>in</strong>g<br />

quantum mechanical systems, possibly hav<strong>in</strong>g an unbounded number of particles,<br />

obey<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> statistics. The unitaries W (f) are called Weyl<br />

operators.<br />

Consider a self-adjo<strong>in</strong>t, positive operator H on h, not hav<strong>in</strong>g zero as an<br />

eigenvalue and such that eitHh0 ⊆ h0 for all t ∈ R and h0 ⊆ Dom(e−βH (I −<br />

e−βH ) −1 )forsomeβ>0. The gauge <strong>in</strong>variant, quasi-free state over U(h0)<br />

determ<strong>in</strong>ed by H and β>0isgivenby<br />

� �<br />

�<br />

1 − 4<br />

ω(W (f)) := e<br />

�<br />

f�<br />

(I+e −βH )(I−e −βH ) −1 f<br />

, f ∈ h0 .<br />

The operator H also specifies the group of automorphisms through the Bogoliubov<br />

transformations of U(h0):<br />

αt(W (f)) := W (e itH f) , f ∈ h0 t ∈ R .<br />

Although this is not a strongly cont<strong>in</strong>uous automorphism group of U(h0), it is<br />

easy to check that it leaves the quasi-free state ω <strong>in</strong>variant, it def<strong>in</strong>es a group<br />

of σ-weakly-cont<strong>in</strong>uous automorphisms {�αt : t ∈ R} of the von Neumann<br />

algebra Mω := πω(U(h0)) ′′ <strong>in</strong> the GNS representation (πω , Hω) ofω. This<br />

weak extension is <strong>in</strong>deed a σ-weakly-cont<strong>in</strong>uous group of automorphisms and<br />

one checks that the weak extension �ω is a (�α,β)-KMS state on Mω. Phase<br />

transitions occur for these systems, accommodat<strong>in</strong>g, for example, the phenomena<br />

of Bose-E<strong>in</strong>ste<strong>in</strong> condensation (see [BR2]).<br />

We now def<strong>in</strong>e the class of semigroups, we are <strong>in</strong>terested <strong>in</strong> these lectures.<br />

As already mentioned above, the condition we <strong>in</strong>troduce generalizes the symmetry<br />

condition of the commutative or tracial cases <strong>in</strong>troduced <strong>in</strong> [AHK1].<br />

In the particular case of a von Neumann algebra, and with respect to the<br />

modular automorphisms group, it was <strong>in</strong>troduced <strong>in</strong> [Cip] [GL1,2]. It is a<br />

generalization of the detailed balance condition ofthetypeconsidered<strong>in</strong>the<br />

theory of Quantum Open Systems (see [D1], [GKS]). We notice, <strong>in</strong> particular,<br />

that it applies to dynamical semigroups and KMS states on C ∗ -algebras.<br />

Def<strong>in</strong>ition 2.31. (KMS-symmetry and KMS-duality) Letα := {αt :<br />

t ∈ R} be a strongly cont<strong>in</strong>uous group of automorphisms of a C∗-algebra A<br />

and ω be a fixed (α, β)-KMS state for some β ∈ R. A bounded map Φ : A → A<br />

is said to be (α, β)-KMS symmetric with respect to ω if<br />

� � �<br />

ω bΦ(a) = ω<br />

α − iβ<br />

2<br />

(a)Φ(α iβ<br />

+ (b))<br />

2<br />

�<br />

(2.15)<br />

for all a, b <strong>in</strong> a norm dense, α-<strong>in</strong>variant ∗ -subalgebra B of Aα. Astrongly<br />

cont<strong>in</strong>uous semigroup {Φt : t ≥ 0} on A is said to be (α, β)-KMS symmetric


Dirichlet Forms on Noncommutative Spaces 187<br />

with respect to ω if Φt is (α, β)-KMS symmetric with respect to ω for all<br />

t ≥ 0. In the von Neumann algebra case, ω is assumed to be normal, maps<br />

and semigroups to be σ-weakly cont<strong>in</strong>uous and the subalgebra B has to be<br />

σ-weakly dense. Two maps Φ1 ,Φ2 : A → A satisfy<strong>in</strong>g the condition<br />

� � �<br />

ω bΦ1(a) = ω<br />

α − iβ<br />

2<br />

(a)Φ2(α iβ<br />

+ (b))<br />

2<br />

for all a, b <strong>in</strong> a norm dense, α-<strong>in</strong>variant ∗ -subalgebra B of Aα will be said to<br />

be <strong>in</strong> (α, β)-KMS duality with respect to ω.<br />

In the von Neumann algebra case the duality condition above was <strong>in</strong>troduced<br />

<strong>in</strong> [AC] to fully generalize the notion of conditional expectation and it was<br />

generalized <strong>in</strong> [Petz] for a weight <strong>in</strong> place of a state. In the von Neumann algebra<br />

case this condition plays a dist<strong>in</strong>ctive role <strong>in</strong> Quantum Communication<br />

Processes as described <strong>in</strong> [OP].<br />

Notice that if <strong>in</strong> the above def<strong>in</strong>ition β =0,thenωis necessarily a tracial<br />

state and (2.15) simplifies to<br />

ω(Φ(a)b) =ω(aΦ(b)) a, b ∈ A. (2.16)<br />

Maps satisfy<strong>in</strong>g the above symmetry condition with respect to traces on C∗ or von Neumann algebras, were considered <strong>in</strong> [G1], [AHK1], [S2], [DL1].<br />

Notice also that, for a general (α,β)-KMS state ω, if the maps Φ commutes<br />

with the automorphism group α then, us<strong>in</strong>g the KMS-condition (2.14), one<br />

easily checks that the KMS-symmetry (2.15) is equivalent to the condition<br />

(2.16). This symmetry condition, <strong>in</strong>troduced <strong>in</strong> [KFGV] as the quantum detailed<br />

balance condition (but also referred to as GNS-symmetry) playsarole<br />

<strong>in</strong> the theory of Quantum Open Systems.<br />

The essential difference between KMS-symmetry and the quantum detailed<br />

balance condition is that, as opposed to the former, only the latter forces the<br />

map Φ to commute with the automorphism group. More explicitly, if a map<br />

Φ satisfies (2.16) with respect to the (α, β)-KMS state ω then it commutes<br />

with α and satisfies (2.15) too (see [Cip1] for the proof and [FU] for several<br />

examples of this fact).<br />

In the framework of the dynamical approach to equilibrium, given a dynamical<br />

system α := {αt : t ∈ R} on A, one would like to f<strong>in</strong>d, for each fixed phase<br />

of the system represented by the (α, β)-KMS state ω, an ergodic, identity<br />

preserv<strong>in</strong>g, positive semigroup Φ = {Φt : t ≥ 0} on A, hav<strong>in</strong>gωasaunique, <strong>in</strong>variant state. In other words, one requires that<br />

and also that, for all states ω ′ on A<br />

ω(Φt(a)) = ω(a) a ∈ A, t≥ 0 ,<br />

lim<br />

t→0 ω′ ◦ Φt = ω,<br />


188 F. Cipriani<br />

the convergence tak<strong>in</strong>g place <strong>in</strong> weak ∗ -topology of A ∗ . If a phase transition<br />

occurs, these requirements are <strong>in</strong>compatible with the commutativity between<br />

the semigroup and the automorphism group, i.e. between the dissipative dynamics<br />

and the time evolution of the system. In that case, <strong>in</strong> fact, all equilibria<br />

of the system, be<strong>in</strong>g <strong>in</strong> particular α-<strong>in</strong>variant states, are Φ-<strong>in</strong>variant too, <strong>in</strong><br />

contrast to the required ergodicity.<br />

Notice also that, when Φ is the identity map then, us<strong>in</strong>g the α-<strong>in</strong>variance of<br />

ω only, one may check that the identity (2.15) reduce to the KMS condition<br />

(2.14):<br />

�<br />

ω(ba) =ω<br />

α − iβ<br />

2<br />

� � �<br />

(a)α iβ<br />

+ (b)<br />

2<br />

= ω aα+iβ(b) .<br />

In other words, an (α, β)- KMS symmetric cont<strong>in</strong>uous semigroup {Φt : t ≥ 0}<br />

on A represents a one parameter deformation of the (α,β)-KMS condition<br />

for ω.<br />

Example 2.32. (KMS-symmetric, norm cont<strong>in</strong>uous, completely Markovian<br />

semigroups) We have seen, <strong>in</strong> Example 2.26, the general form of the L<strong>in</strong>dblad<br />

generator L of a completely Markovian, identity preserv<strong>in</strong>g, norm cont<strong>in</strong>uous,<br />

σ-weakly cont<strong>in</strong>uous semigroup on a von Neumann algebra M:<br />

L(a) =i[H, a]+ �<br />

y ∗ i yia − 2y ∗ i ayi + ay ∗ i yi ,<br />

i∈I<br />

where �<br />

k∈I y∗ i yi has to converge <strong>in</strong> M and �<br />

i∈I y∗ i ayi ∈ M for all a ∈ M.<br />

Suppose that (M , H , H + , J ) is a standard form of M and ξ0 ∈H + is the<br />

vector represent<strong>in</strong>g a fixed normal state ω0 on it. Assume I to be a f<strong>in</strong>ite<br />

<strong>in</strong>dex and the operator coefficients {yi : i ∈ I} as well as H to be entire<br />

analytic with respect to the modular group {σt : t ∈ R} of ω0. Noticethat,<br />

s<strong>in</strong>ce L is bounded, the series<br />

Φt = �<br />

n≥0<br />

t n<br />

n! Ln<br />

is norm convergent and the ω0-KMS symmetry of L would imply the ω0-<br />

KMS symmetry of Φt, which <strong>in</strong> turn would be a ω0-KMS symmetric, norm<br />

cont<strong>in</strong>uous, completely Markovian semigroup on M. In this respect it has<br />

been proved <strong>in</strong> [P3] that under the assumption<br />

�<br />

xiJ0xiJ0 = �<br />

i∈I<br />

i∈I<br />

x ∗ i J0x ∗ i J0 ,<br />

L is ω0-KMS symmetric if and only if the follow<strong>in</strong>g identity holds true<br />

H = �<br />

�<br />

i∈I<br />

σt(x<br />

R<br />

∗ kσ−i/2(xk) − σi/2(x ∗ k )xk)f0(t) dt ,


Dirichlet Forms on Noncommutative Spaces 189<br />

for xi := σ i/4(yi) i ∈ I and f0(t) :=1/ cosh(2πt) t ∈ R. Moreover, a complete<br />

characterization has been recently achieved <strong>in</strong> [FU1 Theorem 7.3] <strong>in</strong> the case<br />

when M is a type I factor B(h). Us<strong>in</strong>g the Hilbert-Schmidt standard form<br />

of Example 2.19 where the state ω0 is represented by a positive, trace class<br />

operator ρ ∈ B(h), ω0(x) =Trace(ρx), sett<strong>in</strong>g<br />

G := − 1<br />

2<br />

�<br />

k∈I<br />

y ∗ k yk − iH ,<br />

one has that the ω0-KMS symmetry of L is equivalent to the follow<strong>in</strong>g conditions<br />

ρ 1/2 G ∗ = Gρ 1/2 + icρ 1/2<br />

ρ 1/2 y ∗ k<br />

= �<br />

l∈I<br />

uklylρ 1/2<br />

k ∈ I<br />

for some c ∈ R and some unitary matrix [ukl] n k,l=1 ∈ Mn(C), n be<strong>in</strong>g the<br />

card<strong>in</strong>ality of the <strong>in</strong>dex set I.<br />

Us<strong>in</strong>g the notation of Proposition 2.28, we are go<strong>in</strong>g to see that KMSsymmetry<br />

may be expressed, alternatively <strong>in</strong> a way which does not refer to a<br />

specific dense, <strong>in</strong>variant ∗ -subalgebra of analytic elements. For simplicity, we<br />

state and prove this result for C ∗ -algebras only, leav<strong>in</strong>g the reader to adapt<br />

the details to the von Neumann algebra situation.<br />

Proposition 2.33. Let α := {αt : t ∈ R} be a strongly cont<strong>in</strong>uous group of<br />

automorphisms of a C∗-algebra A and ω a (α,β)-KMS state for some β ∈ R.<br />

Then, for a bounded map Φ : A → A, the follow<strong>in</strong>g conditions are equivalent:<br />

i) Φ is (α, β)-KMS symmetric with respect to ω;<br />

ii) for any pair a, b ∈ A, there exists a bounded, cont<strong>in</strong>uous function Fa,b :<br />

Dβ −→ A which is analytic <strong>in</strong> Dβ and such that, for all t ∈ R,<br />

Fa,b(t) =ω � α − t<br />

2 (a)Φ(α + t<br />

2 (b))� , Fa,b(t + iβ) =ω � α + t<br />

2 (b)Φ(α − t<br />

2 (a))� .<br />

(2.17)<br />

Furthermore, if the above conditions are satisfied, then the function Fa,b is<br />

bounded on Dβ and<br />

|Fa,b(z)| ≤�Φ�·�a�·�b� , z ∈ Dβ .<br />

F<strong>in</strong>ally, for a ∈ A,b ∈ Aα, the function Fa,b is the restriction to Dβ of the<br />

entire function<br />

Ga,b(z) :=ω � α− z (a)Φ(α+ z<br />

2 2 (b))� , z ∈ C .<br />

Proof. Assum<strong>in</strong>g i) and denot<strong>in</strong>g by B ⊆ Aα the subalgebra occurr<strong>in</strong>g <strong>in</strong><br />

Def<strong>in</strong>ition 2.31, consider, for a fixed pair a, b ∈ B, the function


190 F. Cipriani<br />

Fa,b(z) :=ω � α− z (a)Φ(α+ z<br />

2 2 (b))� , z ∈ C .<br />

Then Fa,b is an entire function and, us<strong>in</strong>g the (α, β)-KMS symmetry, we have,<br />

for any t ∈ R,<br />

Fa,b(t) =ω � α − t<br />

2 (a)Φ(α + t<br />

2 (b))� ,<br />

Fa,b(t + iβ) =ω � α − t+iβ<br />

2<br />

(b)Φ(α t+iβ<br />

+ (a))<br />

2<br />

�<br />

= ω � α iβ<br />

− (α t −<br />

2 2 (a))Φ(α + iβ (α t +<br />

2 2 (b)))�<br />

= ω � α t + 2 (b)Φ(α− t<br />

2 (a))� .<br />

As the map z ↦→ αz(b) is analytic, it follows that the functions [0,β] ∋ s ↦→<br />

�α ± is<br />

2 (b)� are cont<strong>in</strong>uous, hence they are bounded and M± := sup{�α ± is<br />

2 (b)�:<br />

s ∈ [0,β]} are f<strong>in</strong>ite. We then have, for all t + is ∈ Dβ,<br />

|Fa,b(t + is)| = � � ω � α− t<br />

2 (α − is<br />

2 (a))Φ(α + t<br />

2 (α + is<br />

2 (b)))�� � ≤�Φ�·M− · M+ ,<br />

|Fa,b(t)| ≤�Φ�·�a�·�b�<br />

|Fa,b(t + iβ)| ≤�Φ�·�a�·�b� .<br />

Apply<strong>in</strong>g the Three-L<strong>in</strong>e Theorem [BR2 Proposition 5.3.5], we f<strong>in</strong>ally get the<br />

bound<br />

sup<br />

z∈Dβ<br />

|Fa,b(z)| ≤max � sup |Fa,b(t)| ; sup |Fa,b(t + iβ)|<br />

t∈R<br />

t∈R<br />

� ≤�Φ�·�a�·�b� .<br />

For general a, b ∈ A one proceeds by approximation, choos<strong>in</strong>g sequences<br />

{an ∈ B : n ≥ 1} and {bn ∈ B : n ≥ 1}, norm convergent to a and<br />

b, respectively, and uniformly bounded with �an� ≤�a�, �bn� ≤�b� for<br />

n, m ≥ 1. For z belong<strong>in</strong>g to the boundary l<strong>in</strong>es of Dβ, wehave<br />

|Fan,bn(z) − Fam,bm(z)|<br />

≤ max � sup |Fan,bn(t) − Fam,bm(t)| ; sup |Fan,bn((t + iβ))<br />

t∈R<br />

t∈R<br />

− Fam,bm(t)(t + iβ)| �<br />

=max � sup |ω<br />

t∈R<br />

� α t − 2 (an)Φ(α t + 2 (bn)) � − ω � α t − 2 (am)Φ(α t + 2 (bm)) � | ;<br />

|ω � α + t<br />

2 (bn)Φ(α − t<br />

2 (an)) � − ω � α + t<br />

2 (bm)Φ(α − t<br />

sup<br />

t∈R<br />

�<br />

�<br />

≤�Φ� �an − am�·�b� + �a�·�bn − bm� .<br />

2 (am)) � | �<br />

This shows that {Fan,bn : n ≥ 1} is a uniformly Cauchy sequence of bounded,<br />

cont<strong>in</strong>uous functions on the boundary of the closed strip Dβ, hence, by the<br />

Three-L<strong>in</strong>e Theorem aga<strong>in</strong>, it is a Cauchy sequence of bounded, cont<strong>in</strong>uous<br />

functions on the whole closed strip. Its limit function Fa,b is then a bounded,<br />

cont<strong>in</strong>uous function on Dβ, which is analytic on Dβ and satisfies


Dirichlet Forms on Noncommutative Spaces 191<br />

Fa,b(t) = lim<br />

n→∞ ω�α t − 2 (an)Φ(α<br />

�<br />

t + (bn))<br />

2<br />

= ω � α − t<br />

2 (a)Φ(α + t<br />

2 (b))�<br />

Fa,b(t + iβ) = lim<br />

n→∞ ω� α + t<br />

2 (b)Φ(α− t<br />

2<br />

= ω � α t + 2 (bn)Φ(α t − 2 (a))�<br />

sup |Fa,b(z)| ≤�Φ�·�a�·�b� .<br />

z∈Dβ<br />

Conversely, assume condition ii) and def<strong>in</strong>e the function<br />

Ga,b(z) :=ω � α− z (a)Φ(α+ z<br />

2 2 (b))� ,<br />

t ∈ R ,<br />

�<br />

(an))<br />

t ∈ R ,<br />

for a fixed pair a, b ∈ B. ThenGa,b is an entire function such that Ga,b(t) =<br />

Fa,b(t) for any t ∈ R. Be<strong>in</strong>g analytic with<strong>in</strong> the open strip, the functions Ga,b<br />

and Fa,b have to co<strong>in</strong>cide there and then on the whole closed strip Dβ, so<br />

that<br />

Fa,b(z) =ω � α− z (a)Φ(α+ z<br />

2 2 (b))� z ∈ Dβ .<br />

In particular, by (2.18), we have the desired identity<br />

�<br />

�<br />

ω<br />

= Ga,b(+iβ) =Fa,b(+iβ) =ω(bΦ(a)) .<br />

α − iβ<br />

2<br />

(a)Φ(α iβ<br />

+ (b))<br />

2<br />

Corollary 2.34. The (α, β)-KMS symmetry of a map Φ : A → A with respect<br />

to a (α, β)-KMS state ω holds true with respect to a specific dense, α-<strong>in</strong>variant<br />

∗-subalgebra of α-analytic elements if and only if it holds true on the whole<br />

of Aα.<br />

Remark 2.35. The above result allows the derivation of alternative forms of<br />

the KMS symmetry, for example the follow<strong>in</strong>g one:<br />

ω � Φ(a)α + iβ<br />

2<br />

(b) � = ω � α iβ<br />

− (a)Φ(b)<br />

2<br />

�<br />

a, b ∈ Aα . (2.18)<br />

In fact, by Corollary 2.34 we may consider the KMS symmetry (2.15) to hold<br />

on Aα. Apply<strong>in</strong>g the KMS condition for ω to both sides of (2.15) we get<br />

(2.18) because α iβ<br />

+ is bijective on Aα, the <strong>in</strong>verse be<strong>in</strong>g the map α iβ<br />

2<br />

− (see<br />

2<br />

Appendix 7.3). Us<strong>in</strong>g the KMS condition on the right hand side of (2.18), we<br />

get our last form of the KMS symmetry:<br />

ω(Φ(a)α + iβ<br />

2<br />

(b)) = ω(Φ(b)α iβ<br />

+ (a)) a, b ∈ Aα . (2.19)<br />

2<br />

Hence, the KMS-symmetry reduces to the symmetry of the bil<strong>in</strong>ear form<br />

(a, b) ↦→ ω(Φ(a)α iβ<br />

+ (b)) on Aα. There is a difference among these formu-<br />

2<br />

lations of the KMS-symmetry which forces one to prefer (2.15) as the true<br />

def<strong>in</strong>ition. While, as a consequence of Proposition 2.33, (2.15) holds true for a<br />

specific dense, α-<strong>in</strong>variant ∗-subalgebra B of Aα if and only if it holds true for


192 F. Cipriani<br />

all of them, the last two forms of the KMS property are equivalent to (2.15)<br />

when they are given, as above, on the whole algebra Aα. Thisisbecause<br />

not only αt(Aα) =Aα for all t ∈ R, butalsoαz(Aα) =Aα for all z ∈ C.<br />

This property is not shared by a generic dense, α-<strong>in</strong>variant ∗ -subalgebra B<br />

of analytic elements.<br />

The first relevant consequence of the KMS-symmetry is that it allows to<br />

study maps and semigroups <strong>in</strong> GNS-representation (πω, Hω,ξω) of the KMSstate<br />

ω. This is based on the follow<strong>in</strong>g result.<br />

Lemma 2.36. Let {αt : t ∈ R} be a strongly cont<strong>in</strong>uous group of automorphisms<br />

of a C ∗ -algebra A and ω be a fixed (α, β)-KMS state for some β ∈ R.<br />

Then a map Φ : A → A which is (α, β)-KMS symmetric with respect to ω<br />

leaves globally <strong>in</strong>variant the kernel ker(πω) of the GNS-representation of ω.<br />

With the obvious assumptions, the same occurs <strong>in</strong> the von Neumann algebra<br />

case too.<br />

Proof. As ker πω = {a ∈ A : ω(a ∗ a) = 0}, kerπω is clearly α-<strong>in</strong>variant.<br />

Assum<strong>in</strong>g a ∈ ker πω, wethenhaveαt(a) ∈ ker πω for all t ∈ R. By condition<br />

(2.18), the α-<strong>in</strong>variance of ω and by the Cauchy-Schwarz <strong>in</strong>equality, we have,<br />

for a fixed b ∈ A and all t ∈ R,<br />

|Fa,b(t)| 2 ≤ ω � α t − 2 (a)∗α t − 2 (a)� · ω � Φ(α t + 2 (b))∗Φ(α t + 2 (b))�<br />

= ω(a ∗ a) · ω � Φ(α t + 2 (b))∗Φ(α t + 2 (b))� =0.<br />

Be<strong>in</strong>g analytic, Fa,b vanishes identically, so that, aga<strong>in</strong> by condition (2.18),<br />

ω � bΦ(a) � = Fa,b(+iβ) =0.<br />

Choos<strong>in</strong>g b = Φ(a) ∗ we f<strong>in</strong>ally have Φ(a) ∈ ker πω.<br />

Proposition 2.37. (Weak extension of the KMS-symmetry maps) Let α :=<br />

{αt : t ∈ R} be a strongly cont<strong>in</strong>uous group of automorphisms of a C ∗ -<br />

algebra A and ω be a fixed (α, β)-KMS state for some β ∈ R. Let(πω, Hω,ξω)<br />

be the correspond<strong>in</strong>g GNS-representation, ˆω the normal extension of ω to<br />

M := πω(A) ′′ and ˆα := {ˆαt : t ∈ R} the unique σ-weakly cont<strong>in</strong>uous group of<br />

automorphisms of M such that<br />

ˆαt(πω(a)) = πω(αt(a)) , a ∈ A t ∈ R , (2.20)<br />

with respect to which ˆω is a (ˆα, β)-KMS state (see [BR2 Corollary 5.3.4]).<br />

Suppose that Φ : A → A is a (α,β)-KMS symmetric map. Then there exists<br />

auniqueσ-weakly cont<strong>in</strong>uous map ˆ Φ : M → M determ<strong>in</strong>ed by<br />

ˆΦ(πω(a)) = πω(Φ(a)) a ∈ A. (2.21)


Dirichlet Forms on Noncommutative Spaces 193<br />

This extension is ˆω-KMS symmetric <strong>in</strong> the sense that<br />

�<br />

ˆω ˆΦ(x)ˆσ− i<br />

2 (y)<br />

� �<br />

=ˆω ˆσ i + 2 (x) ˆ �<br />

Φ(y)<br />

(2.22)<br />

for all x, y <strong>in</strong> a σ-weakly dense, ˆσ-<strong>in</strong>variant ∗-subalgebra of M. Moreover,<br />

� ˆ Φ�≤�Φ� and if Φ is positive then ˆ Φ is positive too.<br />

Proof. By Lemma 2.36, formula (2.21) validly determ<strong>in</strong>es a unique, σ-weakly<br />

densely def<strong>in</strong>ed map ( ˆ Φ, D( ˆ Φ)) on M, its doma<strong>in</strong> be<strong>in</strong>g D( ˆ Φ):=πω(A) ⊂ M.<br />

By Corollary 2.34, we may suppose the KMS symmetry of ω satisfied on<br />

the whole algebra of analytic elements Aα. As the modular group �σt of the<br />

normal extension �ω of the (α,β)-KMS state ω co<strong>in</strong>cides with the group �α−βt<br />

(see [BR2 Corollary 5.3.4] and the discussion after [BR2 Def<strong>in</strong>ition 5.3.1]),<br />

the �ω-KMS symmetry condition (2.22) follows directly from (2.18) and it is<br />

satisfied on the σ-weakly dense, �σ-<strong>in</strong>variant ∗-subalgebra πω(Aα) ⊆ M.<br />

For the sake of simplicity, we shall assume from now on that πω is faithful, so<br />

that ker πω = {0}. In particular, we shall identify elements a of A with their<br />

images πω(a) <strong>in</strong>M so that A ⊆ M and D( � Φ)=A.<br />

To extend the map � Φ to the whole von Neumann algebra M, westartby<br />

notic<strong>in</strong>g that it is bounded on its doma<strong>in</strong>:<br />

� � Φ(a)� ≤�Φ�·�a� ∀a ∈ A ⊆ M .<br />

We now proceed to show that the �ω-KMS symmetry condition (2.22) implies<br />

the map ( ˆ Φ, D( ˆ Φ)) is σ-weakly closable on M. Infact,let{xi : i ∈ I} ⊂A be<br />

a net <strong>in</strong> its doma<strong>in</strong>, σ-weakly converg<strong>in</strong>g to zero, for which the net { ˆ Φ(xi) :<br />

i ∈ I} ⊂M is σ-weakly convergent to some z ∈ M. Then by (2.22)<br />

�<br />

�ω<br />

z�σ − i<br />

2 (y)<br />

�<br />

= lim<br />

�<br />

�ω �Φ(xi) �σ i −<br />

i∈I 2 (y)<br />

�<br />

�<br />

�ω �σ i + (xi)<br />

i∈I 2<br />

� �<br />

Φ(y)<br />

� �<br />

ξ�ω<br />

� −<br />

∆<br />

i∈I<br />

1<br />

1<br />

2 + 2<br />

�ω xi∆ �ω � Φ(y)ξ�ω<br />

� � 1<br />

ξ�ω<br />

� + �<br />

2 xi∆<br />

= lim<br />

= lim<br />

= lim<br />

i∈I<br />

=0<br />

�ω � Φ(y)ξ�ω<br />

�<br />

(2.23)<br />

for any y ∈ Aα. Choos<strong>in</strong>g y = �σ + i<br />

2<br />

(x), we have �ω(zx) = 0 for all x ∈ Aα.<br />

S<strong>in</strong>ce Aα is σ-weakly dense <strong>in</strong> M we get z =0,thusprov<strong>in</strong>g � Φ to be σweakly<br />

closable. To show that (2.21) determ<strong>in</strong>es a σ-weakly cont<strong>in</strong>uous map,<br />

def<strong>in</strong>ed everywhere on M, we are left to prove that the σ-weak closure of<br />

( ˆ Φ, D( ˆ Φ)) is σ-weakly cont<strong>in</strong>uous. This will be a consequence of the Closed<br />

Graph Theorem, once we show that the closure of ( ˆ Φ, D( ˆ Φ)) is everywhere<br />

def<strong>in</strong>ed. Let x ∈ M be a generic, fixed vector and choose a net {xi : i ∈ I}<br />

⊂ Aσ-weakly converg<strong>in</strong>g to it. The boundedness of the map � Φ implies that<br />

{ � Φ(xi) :i ∈ I} is bounded <strong>in</strong> M so that it is, by the Banach-Alaoglu Theorem,


194 F. Cipriani<br />

a σ-weakly relatively compact set. Possibly consider<strong>in</strong>g a suitable subnet, we<br />

deduce that { � Φ(xi) :i ∈ I} is σ-weakly convergent <strong>in</strong> M, sothatx is <strong>in</strong> the<br />

doma<strong>in</strong> of the closure of the map ( � Φ, D( � Φ)).<br />

To extend the bound � � Φ(x)� ≤ �Φ� ·�x� from x <strong>in</strong> A to the whole von<br />

Neumann algebra, consider x <strong>in</strong> the unit ball of M. By Kaplanski’s density<br />

theorem [BR1 Theorem 2.4.16], there exists a net {xi : i ∈ I} <strong>in</strong> the unit<br />

ball of A, whichisσ-strongly ∗ , hence σ-weakly, convergent to x. S<strong>in</strong>ce � Φ is<br />

σ-weakly cont<strong>in</strong>uous, for every normal functional τ ∈ M∗ we then have<br />

|〈τ , � Φ(x)〉| = | lim<br />

i∈I |〈τ , � Φ(xi)〉 ≤�τ�M∗| lim <strong>in</strong>f<br />

i∈I �Φ�·�xi�M = �τ�M∗ ·�Φ� ,<br />

thus � � Φ(x)� ≤�Φ�. S<strong>in</strong>ce this holds true for all x <strong>in</strong> the unit ball of M, we<br />

get the desired bound.<br />

We now proceed to show that if Φ : A → A is a positive map then � Φ : M → M<br />

is positive too. Fix an element x ∈ M+ and consider, by Kaplansky’s Density<br />

Theorem, a net {xi : i ∈ I} <strong>in</strong> A, σ-strongly∗ convergent to it <strong>in</strong> M. S<strong>in</strong>ce<br />

convergent nets are bounded, by [T2 Chapter II Lemma 2.5] we have xi → x<br />

strongly∗ and hence strongly. Apply<strong>in</strong>g [T2 Chapter II Theorem 4.7], we<br />

have |xi| →x strongly and, by [T2 Chapter Proposition 4.1], |xi| →x <strong>in</strong> the<br />

strong∗ topology too. Us<strong>in</strong>g aga<strong>in</strong> [T2 Chapter II Lemma 2.5] we have that<br />

|xi| →x <strong>in</strong> the σ-strong∗ topology and then σ-weakly too. By the σ-weak<br />

cont<strong>in</strong>uity of the map � Φ,weget� Φ(|xi|) → � Φ(x) σ-weakly <strong>in</strong> M+. S<strong>in</strong>ce the<br />

positive cone M+ is σ-weakly closed, we get � Φ(x) ∈ M+.<br />

With the hypotheses of the previous theorem we have the follow<strong>in</strong>g<br />

Corollary 2.38. (Duality relation of KMS-symmetric maps) The map � Φ∗ :<br />

M∗ → M∗ determ<strong>in</strong>ed by restrict<strong>in</strong>g to the pre-dual space M∗ the adjo<strong>in</strong>t<br />

map Φ ∗ : M ∗ → M ∗ satisfies the duality relation<br />

ˆΦ∗ ◦ jˆω = jˆω ◦ ˆ Φ, (2.24)<br />

where j�ω : M → M∗ is the symmetric embedd<strong>in</strong>g determ<strong>in</strong>ed by �ω. Moreover,<br />

if Φ is positive then � Φ∗ is positive and if Φ is Markovian then � Φ∗ is Markovian<br />

<strong>in</strong> the follow<strong>in</strong>g sense � Φ∗(�ω) ≤ �ω, or, more explicitly,<br />

�ω( � Φ(a)) ≤ �ω(a) a ∈ M+.<br />

Proof. Denot<strong>in</strong>g by ξ�ω the cyclic vector represent<strong>in</strong>g the state �ω and by<br />

∆�ω ,J�ω the associated modular operators, we have J�ωξ�ω = ξ�ω and ∆ z �ω ξ�ω = ξ�ω<br />

for all z ∈ C.<br />

To prove the duality relation (2.24), recall that �σz(x) =∆iz �ω x∆−iz<br />

�ω for all<br />

complex numbers z and all analytic vectors x (see for <strong>in</strong>stance the discussion<br />

follow<strong>in</strong>g [BR1 Corollary 2.5.24]). From the �ω-KMS symmetry (2.22), the<br />

def<strong>in</strong>ition of j�ω (see Def<strong>in</strong>ition 2.21 iii)) and the properties of the modular


Dirichlet Forms on Noncommutative Spaces 195<br />

operators, it follows that, for all x and y <strong>in</strong> the subalgebra M�σ ⊆ M of<br />

analytic elements of the modular group, we have<br />

� �<br />

� � �<br />

ξ�ω �Φ(x)�σ− � i (y)ξ�ω = ξ�ω ��σ i<br />

2<br />

+ (x)<br />

2<br />

� �<br />

Φ(y)ξ�ω<br />

� �<br />

1 1<br />

ξ�ω<br />

�� 2 2<br />

Φ(x)∆�ω y∆− �ω ξ�ω<br />

� � �<br />

= ξ�ω<br />

� −<br />

∆ 1 1<br />

2 2<br />

�ω x∆�ω � �<br />

Φ(y)ξ�ω<br />

� �<br />

1<br />

ξ�ω<br />

�� 2 Φ(x)∆�ω yξ�ω<br />

� � � 1<br />

= ξ�ω<br />

� 2 x∆�ω � �<br />

Φ(y)ξ�ω<br />

� �<br />

ξ�ω<br />

�� Φ(x)J�ωy ∗ � � �<br />

ξ�ω = ξ�ω<br />

�xJ�ω � Φ(y) ∗ �<br />

ξ�ω<br />

� �<br />

ξ�ω �Φ(x)J�ωy � ∗ � � �<br />

J�ωξ�ω = ξ�ω �xJ�ω � Φ(y) ∗ �<br />

J�ωξ�ω<br />

� �<br />

ξ�ω<br />

�J�ωy ∗ J�ω � � � �<br />

Φ(x)ξ�ω = ξ�ω<br />

�J�ω � Φ(y) ∗ �<br />

J�ωxξ�ω<br />

� � � �<br />

J�ωyJ�ωξ�ω<br />

�� Φ(x)ξ�ω = J�ω � � �<br />

Φ(y)J�ωξ�ω<br />

�xξ�ω � � � �<br />

J�ωyξ�ω �Φ(x)ξ�ω � = J�ω � � �<br />

Φ(y)ξ�ω �xξ�ω<br />

〈j�ω( � Φ(x)) ,y〉 = 〈j�ω(x) , � Φ(y)〉 .<br />

S<strong>in</strong>ce M�σ is σ-weakly dense <strong>in</strong> M, we get the duality relation on the whole<br />

of M. The statements concern<strong>in</strong>g positivity and Markovianity follow directly<br />

from the order properties of the symmetric embedd<strong>in</strong>g j�ω stated <strong>in</strong><br />

Proposition 2.23.<br />

The next result shows that not only s<strong>in</strong>gle maps but also KMS symmetric,<br />

cont<strong>in</strong>uous semigroups may be extended to the von Neumann algebra<br />

associated to the GNS representation.<br />

Theorem 2.39. (Weak extension of KMS-symmetric cont<strong>in</strong>uous semigroups)<br />

Let α := {αt : t ∈ R} be a strongly cont<strong>in</strong>uous group of automorphisms of<br />

a C∗-algebra A and ω be a fixed (α, β)-KMS state for some β ∈ R. Let<br />

(πω, Hω,ξω) be the correspond<strong>in</strong>g GNS-representation, �ω the normal extension<br />

of ω to M := πω(A) ′′ and �α := {�αt : t ∈ R} the <strong>in</strong>duced σ-weakly<br />

cont<strong>in</strong>uous group of automorphisms of M.<br />

Suppose that {Φt : t ≥ 0} is a strongly cont<strong>in</strong>uous semigroups on A, (α,β)-<br />

KMS symmetric with respect to ω. Then there exists a unique σ-weakly cont<strong>in</strong>uous<br />

semigroup { � Φt : t ≥ 0} on M determ<strong>in</strong>ed by<br />

�Φt(πω(a)) = πω(Φt(a)) a ∈ A, t≥ 0 . (2.25)<br />

This extension is �ω-KMS symmetric <strong>in</strong> the sense that<br />

�<br />

ˆω �Φt(x)�σ i − 2 (y)<br />

� �<br />

= �ω �σ i + 2 (x) � �<br />

Φt(y)<br />

t ≥ 0 , (2.26)


196 F. Cipriani<br />

for all x, y <strong>in</strong> a σ-weakly dense, �σ-<strong>in</strong>variant ∗ -subalgebra of M. Moreover, if<br />

{Φt : t ≥ 0} is positive, completely positive, Markovian or completely Markovian<br />

then { � Φt : t ≥ 0} shares the same properties.<br />

Proof. Apply<strong>in</strong>g Proposition 2.37 to each <strong>in</strong>dividual map of the family, we<br />

get a semigroup { � Φt : t ≥ 0} of σ-weakly cont<strong>in</strong>uous maps on M. Whatis<br />

left to prove is its σ-weak cont<strong>in</strong>uity as a semigroup. This is <strong>in</strong> fact equivalent<br />

to the weak, and hence to the strong cont<strong>in</strong>uity of the semigroup of<br />

σ(M∗, M)-cont<strong>in</strong>uous maps { � Φt∗ : t ≥ 0} on the predual space M∗, def<strong>in</strong>ed<br />

us<strong>in</strong>g Corollary 2.38 and satisfy<strong>in</strong>g (2.24). Hence, by def<strong>in</strong>ition, we have that<br />

� � Φt∗� = � � Φt�. S<strong>in</strong>ce, moreover, {Φt : t ≥ 0} is a strongly cont<strong>in</strong>uous semigroup,<br />

there exist constants M>0, γ ∈ R such that � � Φt� ≤�Φt� ≤M · e γt<br />

(see [BR1 Proposition 3.1.3]). This gives the bound<br />

� � Φt∗(τ) − τ� ≤�� Φt∗(τ) − � Φt∗(τ ′ )� + � � Φt∗(τ ′ ) − τ ′ � + �τ ′ − τ�<br />

≤�� Φt∗�·�τ − τ ′ � + � � Φt∗(τ ′ ) − τ ′ � + �τ ′ − τ�<br />

≤ (1 + M · e γt ) ·�τ − τ ′ � + � � Φt(τ ′ ) − τ ′ �<br />

for any τ,τ ′ ∈ M∗, by which it is enough to verify the strong cont<strong>in</strong>uity of<br />

{ � Φt∗ : t ≥ 0} on a norm dense subspace of M∗. S<strong>in</strong>ceπω(A) is, by def<strong>in</strong>ition,<br />

σ(M, M∗)-dense <strong>in</strong> M and the symmetric embedd<strong>in</strong>g j�ω : M → M∗ has<br />

norm dense range (Proposition 2.23), it is enough to prove strong cont<strong>in</strong>uity<br />

on j�ω(πω(A)). By (2.24) one has � Φt∗ ◦ j�ω = j�ω ◦ � Φt for all t ≥ 0, so that on<br />

j�ω(πω(A)) the restriction of � Φt∗ acts as<br />

�Φt∗(j�ω(πω(a))) = jω(πω(Φt(a))) .<br />

The strong cont<strong>in</strong>uity of { � Φt∗ : t ≥ 0} on j�ω(πω(A)) then follows from that of<br />

{Φt : t ≥ 0} on A. The proofs of the stated order properties follow easily from<br />

the order properties of the symmetric embedd<strong>in</strong>gs listed <strong>in</strong> Proposition 2.23.<br />

Notice that the strongly cont<strong>in</strong>uous semigroup on the predual space M∗<br />

considered <strong>in</strong> the proof of the previous result, and determ<strong>in</strong>ed by the relation<br />

�Φt∗ ◦ j�ω = j�ω ◦ � Φt<br />

is Markovian with respect to �ω <strong>in</strong> the sense that<br />

or, more explicitly,<br />

t ≥ 0 ,<br />

�Φt∗(�ω) ≤ �ω ∀ t ≥ 0 ,<br />

�ω( � Φt(a)) ≤ �ω(a) ∀ a ∈ M+ , ∀ t ≥ 0 .


Dirichlet Forms on Noncommutative Spaces 197<br />

2.4 Markovian Semigroups on Standard Forms<br />

of von Neumann Algebras<br />

So far we have succeeded <strong>in</strong> extend<strong>in</strong>g maps and semigroups, which are KMS<br />

symmetric with respect to a KMS state ω, fromaC ∗ -algebra to the von Neumann<br />

algebra and to its predual space generated by the GNS representation<br />

(πω , Hω ,ξω). Our next task is to perform a further extension to the Hilbert<br />

space Hω, <strong>in</strong> such a way as to keep, when they hold, positivity and Markovianity.<br />

Notice that, while the presence of the standard cone H + ω makes obvious<br />

the formulation of positivity for maps on Hω, the notion of Markovianity has<br />

to be <strong>in</strong>troduced.<br />

To simplify notations, but also for the sake of generality, we start bywork<strong>in</strong>g <strong>in</strong><br />

the standard form (M , H , H + , J ) of a von Neumann algebra M, onwhicha<br />

fixed normal state ω0 is considered. We shall denote by ξ0 ∈H + the positive,<br />

cyclic and separat<strong>in</strong>g vector represent<strong>in</strong>g it: ω0(·) =(ξ0| ·ξ0). Fundamental<br />

for us will be, <strong>in</strong> particular, the cont<strong>in</strong>uity and the order properties of the<br />

symmetric embedd<strong>in</strong>gs i0, j0 associated to ω0, <strong>in</strong>troduced <strong>in</strong> Section 2.2.<br />

Unless otherwise stated, maps T on H are always assumed to be real, <strong>in</strong> the<br />

sense that T (H J ) ⊆H J .<br />

Def<strong>in</strong>ition 2.40. (Positive and Markovian semigroups on Hilbert<br />

spaces of standard forms) Let(M , H , H + , J ) be a standard form of<br />

the von Neumann algebra M and ξ0 ∈H + the vector represent<strong>in</strong>g a fixed<br />

normal state ω0 on it. A map T : H→His said to be<br />

i) positive if T (H + ) ⊆H + ;<br />

ii) Markovian with respect to ξ0 if it is positive and Tξ0 ≤ ξ0;<br />

iii) completely positive (resp. completely Markovian) if,foreveryn∈N∗ ,<br />

the map<br />

T ⊗ In : H⊗Mn(C) →H⊗Mn(C) (T × In)[ξij] n i,j=1 =[Tξij] n i,j=1<br />

is positive (resp. Markovian) on the Hilbert space H⊗Mn(C) of the standard<br />

form of M ⊗ Mn(C) (<strong>in</strong> the tensor product of Hilbert spaces H⊗Mn(C), the<br />

factor Mn(C) and its scalar product is understood as the Hilbert -Schmidt<br />

standard form of the von Neumann algebra Mn(C); the cyclic vector represent<strong>in</strong>g<br />

the normalized trace be<strong>in</strong>g the suitable multiple of the identity<br />

matrix).<br />

iv) a semigroup {Tt : t ≥ 0} on H is said to be positive (resp. Markovian,<br />

completely positive, completely Markovian) if Tt is positive (resp. Markovian,<br />

completely positive, completely Markovian) for all t ≥ 0.


198 F. Cipriani<br />

It is easy to check that T is Markovian with respect to ξ0 if and only if it<br />

leaves globally <strong>in</strong>variant the order <strong>in</strong>terval [0,ξ0] ⊂H + :<br />

0 ≤ ξ ≤ ξ0 ⇒ 0 ≤ Tξ ≤ ξ0 .<br />

Notice that, while the notion of positive map on H deals with the structure<br />

of ordered Hilbert space only, Markovianity depends upon the choice of a<br />

cyclic vector ξ0 ∈H + .<br />

Given a positive semigroup {Tt : t ≥ 0} on H and a number γ ≥ 0, vectors<br />

η ∈H + such that<br />

e −tγ Ttη ≤ η ∀ t ≥ 0<br />

are called γ-excessive, so that the semigroup is Markovian with respect to ξ0<br />

if and only if this vector is 0-excessive with respect to it.<br />

In the commutative case, where M = L∞ (X, µ), H = L2 (X, µ) andH + =<br />

L 2 +<br />

(X, µ) is the cone of positive, square summable functions with respect to<br />

a positive Radon measure µ on X, positivity has the usual mean<strong>in</strong>g. Then, as<br />

long as the measure is f<strong>in</strong>ite, and one considers as cyclic vector the constant<br />

function ξ0 = 1, both positivity and Markovianity <strong>in</strong> the above sense reduce<br />

to well known classical notions (see [FOT] for example). However, even <strong>in</strong><br />

this commutative sett<strong>in</strong>g, the above def<strong>in</strong>ition makes it possible to consider<br />

as Markovian a wider class of semigroups than those which can be treated<br />

by the classical def<strong>in</strong>ition.<br />

Example 2.41. (Markovianity of Schröd<strong>in</strong>ger semigroups with respect to a<br />

ground-state) Consider, for example, the semigroup e −tH on L 2 (R n ,dx)generated<br />

by a Schröd<strong>in</strong>ger operator H = −div ◦ grad + V associated to a nice<br />

potential function V . By the uniqueness of the Beurl<strong>in</strong>g-Deny-Le Jan decomposition<br />

(see Example 4.19 below), these semigroups are positivity preserv<strong>in</strong>g<br />

and not Markovian <strong>in</strong> the classical sense, unless the potential is positive. However,<br />

if H admits a unique ground-state φ0, i.e. the bottom −γ of its spectrum<br />

is an eigenvalue of multiplicity one, then the correspond<strong>in</strong>g normalized eigenvector<br />

φ0 is strictly positive and γ-excessive. This is the case, for example,<br />

when for some λ ≤ 0 the negative part (V −λ)− of the function V −λ belongs<br />

to the space L n<br />

2 (R n ,dx). In the case when γ ≤ 0, e −tH is then Markovian<br />

with respect to φ0 <strong>in</strong> the above sense. Notice that the associated normal state<br />

ω0 gives the distribution of the position of the particle of the quantum system<br />

described by H <strong>in</strong> its ground-state.<br />

Example 2.42. (Stability of Markovianity under subord<strong>in</strong>ation of semigroups)<br />

Let U : G → B(H) be a strongly cont<strong>in</strong>uous representation of a locally<br />

compact abelian group G leav<strong>in</strong>g the positive cone globally <strong>in</strong>variant:<br />

U(g)H + = H + for all g ∈ G. Suppose that {µt : t ≥ 0} is a symmetric<br />

convolution semigroup of probabilities on G. The subord<strong>in</strong>ate semigroup<br />

�<br />

Ttξ := µt(dg)U(g)ξ ξ ∈H<br />

G


Dirichlet Forms on Noncommutative Spaces 199<br />

is a strongly cont<strong>in</strong>uous, symmetric, contractive and positivity preserv<strong>in</strong>g<br />

semigroup. It is also (completely) Markovian with respect to any cyclic vector<br />

ξ0 ∈H + left <strong>in</strong>variant by the representation. This applies, <strong>in</strong> particular, to<br />

cases <strong>in</strong> which the representation is given by the modular automorphism<br />

group of any faithful, normal state on M. Consider,forexample,thecase<br />

where<br />

G = R µt(dx) :=(4πt) −1/2 e −x2 /4t dx , t > 0 , µ0 := δ0<br />

and U(t) =∆ it<br />

0 where ∆0 is the modular operator correspond<strong>in</strong>g to a cyclic<br />

vector ξ0 ∈H + .Thenwehave<br />

−t| log ∆0|2<br />

Tt = e<br />

t ≥ 0 .<br />

Other example of this construction are discussed <strong>in</strong> [Cip1 Chapter 3].<br />

Notice that, while the def<strong>in</strong>ition of positive map on the Hilbert space<br />

of a standard form only depends upon the order structure of the Hilbert<br />

space, Markovianity is a property deeply connected with the behavior of the<br />

map at the algebra level. The next result will show that Markovian maps<br />

and semigroups on H are exactly those which extend to normal, positive<br />

contractive maps on the von Neumann algebra M and to weakly cont<strong>in</strong>uous,<br />

positive, contractive maps on the predual M∗.<br />

Proposition 2.43. Let Φ : M → M be a (real) l<strong>in</strong>ear map on the von Neumann<br />

algebra M. Then the follow<strong>in</strong>g properties are equivalent:<br />

i) Φ is ω0-KMS symmetric<br />

ω(Φ(x)σ − i<br />

2 (y)) = ω(σ + i<br />

2 (x)Φ(y)) x, y ∈ Mσ ;<br />

ii) the densely def<strong>in</strong>ed l<strong>in</strong>ear map T : D(T ) →Hon the Hilbert space H,<br />

given by<br />

T ◦ i0 = i0 ◦ Φ, D(T ):=i0(Mσ) ,<br />

is symmetric<br />

(Tξ| η) =(ξ| Tη) ξ,η∈ D(T );<br />

iii) the densely def<strong>in</strong>ed l<strong>in</strong>ear map Φ∗ : D(Φ∗) → M∗, on the predual space<br />

M∗, given by<br />

Φ∗ ◦ j0 = j0 ◦ Φ, D(Φ∗) :=j0(Mσ) ,<br />

is <strong>in</strong> duality with Φ:<br />

〈Φ∗(j0(x))|y〉 = 〈j0(x)|Φ(y)〉 x, y ∈ M .<br />

Suppose that the previous symmetry conditions hold true. Then


200 F. Cipriani<br />

iv) Φ,T and Φ∗ are bounded maps and Φ is σ(M, M∗)-cont<strong>in</strong>uous;<br />

v) the Markovianity of Φ, T and Φ∗ are equivalent.<br />

Proof. Notice first that, by Proposition 2.23, D(T )isnormdense<strong>in</strong>H because<br />

Mσ is σ(M , M∗)-dense <strong>in</strong> M and i0 is σ(M , M∗)−σ(H , H) cont<strong>in</strong>uous<br />

with dense range.<br />

Analogously, D(Φ∗)isσ(M∗ , M)dense<strong>in</strong>M∗ because, Mσ is σ(MM∗)-dense<br />

<strong>in</strong> M and j0 is σ(M∗ M) − σ(MM∗) cont<strong>in</strong>uous with dense range. S<strong>in</strong>ce for<br />

all x, y ∈ Mσ we have<br />

ω(Φ(x)σ i − (y)) = (ξ0| Φ(x)∆<br />

2<br />

1<br />

2<br />

0 yξ0) =(i0(Φ(x ∗ ))| i0(y)) = (Ti0(x ∗ )| i0(y))<br />

and<br />

1 + 2<br />

0 Φ(y)ξ0) =(i0(x ∗ )| i0(Φ(y))) = (i0(x ∗ )| Ti0(y)) ,<br />

ω(σ i<br />

2 (x)Φ(y)) = (ξ0| x∆<br />

the ω0-KMS symmetry of Φ and the symmetry of T are equivalent. Analogously,<br />

as for all x, y ∈ Mσ we have<br />

and<br />

ω(Φ(x)σ − i<br />

2 (y)) = (ξ0| Φ(x)∆ 1<br />

2<br />

0 yξ0) =(ξ0| Φ(x)Jy ∗ Jξ0)<br />

=(Jyξ0| Φ(x)ξ0) =〈j0(Φ(x))| y〉<br />

1 + 2<br />

0 Φ(y)ξ0) =(ξ0| xJΦ(y) ∗ Jξ0) =〈j0(x)| Φ(y)〉 ,<br />

ω(σ i<br />

2 (x)Φ(y)) = (ξ0| x∆<br />

the ω0-KMS symmetry of Φ and the duality between the maps Φ and Φ∗ are<br />

also equivalent properties.<br />

Suppose now that one of the above conditions holds true. By the duality<br />

relation with Φ∗, themapΦhas a densely def<strong>in</strong>ed adjo<strong>in</strong>t map, and hence<br />

is σ(M , M∗)-closable. As it is also everywhere def<strong>in</strong>ed, it is σ(M , M∗) and<br />

norm cont<strong>in</strong>uous. Consequently, the map Φ∗ is σ(M∗ , M) and norm cont<strong>in</strong>uous.<br />

The boundedness of T follows from the <strong>in</strong>terpolation bound (2.12) (see<br />

Proposition 2.24).<br />

The last statement concern<strong>in</strong>g the equivalence between Markovianity of Φ, T<br />

and Φ∗ easily follows from the order properties of the symmetric <strong>in</strong>jections<br />

(see Proposition 2.22).<br />

The follow<strong>in</strong>g is the ma<strong>in</strong> result of this section: the class of strongly cont<strong>in</strong>uous,<br />

symmetric semigroups on the Hilbert space H which are Markovian with<br />

respect to ξ0, the class ω0-KMS symmetric, σ(M, M∗)-cont<strong>in</strong>uous Markovian<br />

semigroups on the von Neumann algebra M and the class of ω0-Markovian,<br />

strongly cont<strong>in</strong>uous semigroups on the predual space M∗ are <strong>in</strong> one to one<br />

correspondence.


Dirichlet Forms on Noncommutative Spaces 201<br />

Theorem 2.44. Let (M , H , H + , J ) be a standard form of the von Neumann<br />

algebra M, andξ0∈H + be a cyclic vector represent<strong>in</strong>g a fixed normal state<br />

ω0 on it.<br />

i) Let {Φt : t ≥ 0} be a ω0-KMS symmetric, σ(M, M∗)-cont<strong>in</strong>uous semigroup<br />

on M. Then the relation<br />

Tt ◦ i0 = i0 ◦ Φt<br />

<strong>in</strong>duces a strongly cont<strong>in</strong>uous semigroup on the Hilbert space H; analogously,<br />

the relation<br />

Φ∗t ◦ j0 = j0 ◦ Φt<br />

<strong>in</strong>duces a strongly cont<strong>in</strong>uous semigroup on the predual space M∗.<br />

ii) Suppose that {Tt : t ≥ 0} is a symmetric, strongly cont<strong>in</strong>uous semigroup<br />

on H, Markovian with respect to ξ0. Then the relation<br />

Tt ◦ i0 = i0 ◦ Φt<br />

<strong>in</strong>duces a ω0-KMS symmetric, σ(M, M∗)-cont<strong>in</strong>uous, Markovian semigroup<br />

on M; analogously, the relation<br />

Φ∗t ◦ i0∗ = i0∗ ◦ Tt<br />

<strong>in</strong>duces a strongly cont<strong>in</strong>uous semigroup on the predual space M∗.<br />

Moreover, positivity, complete positivity, Markovianity and complete Markovianity,<br />

are properties shared by all or none of the above semigroups.<br />

Proof. A direct application of Proposition 2.43 verifies both the correct def<strong>in</strong>ition<br />

of semigroups of cont<strong>in</strong>uous maps on the spaces H and M∗, <strong>in</strong>item<br />

i), and on the spaces M and M∗, <strong>in</strong> item ii), and their order properties. The<br />

proof of the cont<strong>in</strong>uity with respect to the parameter t is straightforward and<br />

is left to the reader.<br />

2.5 Ergodic, Positive Semigroups<br />

A classical result, due to G. Frobenius and O. Perron, states that the eigenvalue<br />

�a� of a hermitian irreducible a ∈ Mn(C), with nonnegative entries<br />

is simple and there exists an associated eigenvector with nonnegative coord<strong>in</strong>ates.<br />

This uniqueness result has been extended by L. Gross [G1], <strong>in</strong><br />

an <strong>in</strong>f<strong>in</strong>ite dimensional sett<strong>in</strong>g, to symmetric, positive operators act<strong>in</strong>g on<br />

Hilbert spaces of type L 2 (X, m), where positivity has the usual po<strong>in</strong>twise<br />

mean<strong>in</strong>g. The same author also extended the result <strong>in</strong> the noncommutative<br />

sett<strong>in</strong>g, for positive operators act<strong>in</strong>g on the Segal standard form L 2 (M,τ)<br />

of a f<strong>in</strong>ite von Neumann algebra M (see [Se]). In this framework the author


202 F. Cipriani<br />

discussed existence and uniqueness of ground-states of physical Hamiltonians<br />

for Boson and Fermion systems <strong>in</strong> Quantum Field <strong>Theory</strong> (see also [F]).<br />

In this section we illustrate how to extend the uniqueness result to positive<br />

operators act<strong>in</strong>g on Hilbert spaces of standard forms (M , H , H + , J )ofvon<br />

Neumann algebras M.<br />

Let us recall that a closed sub-cone F ⊆H + is called a face of H + if it is<br />

hereditary, <strong>in</strong> the sense that<br />

ξ ∈F, η ∈H + , η ≤ ξ ⇒ η ∈F.<br />

A face is proper if it is different from the trivial ones: {0} and H + (see [Co1]).<br />

Def<strong>in</strong>ition 2.45. (Ergodic maps and semigroups) A positive map T : H→<br />

H is said to be<br />

i) ergodic if for all ξ,η ∈ H + , ξ,η �= 0, there exists n ∈ N∗ such that<br />

(ξ| T nη) > 0;<br />

ii) <strong>in</strong>decomposable if it leaves <strong>in</strong>variant no proper faces of H + , i.e. for a face<br />

F⊆H + ,ifT (F) ⊆F then necessarily F = {0} or F = H + .<br />

A positive semigroup {Tt : t ≥ 0} on H is said to be<br />

iii) ergodic if for all ξ,η ∈ H + , ξ,η �= 0, there exists t > 0 such that<br />

(ξ| Ttη) > 0;<br />

iv) <strong>in</strong>decomposable if Tt is <strong>in</strong>decomposable for some t>0.<br />

Example 2.46. (Faces of positive cones of commutative von Neumann algebras)<br />

In a cone like L 2 +(X, m), the subcone FB of all positive functions vanish<strong>in</strong>g<br />

on the complement of a Borel set B ⊆ X of positive measure is a face,<br />

and all of them arise <strong>in</strong> this way.<br />

Faces of the self-polar cone H + are <strong>in</strong> one-to-one correspondence with projections<br />

of the von Neumann algebra M (see [Co1]). In fact, sett<strong>in</strong>g Pe := eJeJ<br />

for projections e ∈ Proj(M), one has that<br />

F = Pe(H + )<br />

is a face and all of them can described <strong>in</strong> this way.<br />

Theorem 2.47. (Ergodicity of positive maps and semigroups [Cip3]) Let T :<br />

H→Hbe a symmetric, positive map.<br />

The follow<strong>in</strong>g properties are equivalent:<br />

i) �T � is a simple eigenvalue and there exists a correspond<strong>in</strong>g cyclic eigenvector<br />

<strong>in</strong> H + ;<br />

ii) T is <strong>in</strong>decomposable;


Dirichlet Forms on Noncommutative Spaces 203<br />

iii) T commutes with no proper projection Pe, e ∈ Proj(M);<br />

iv) T is ergodic.<br />

Corollary 2.48. (Ergodicity of positive semigroups [Cip3]) Let {Tt : t ≥<br />

0} be a strongly cont<strong>in</strong>uous, positive, symmetric semigroup on H, andlet<br />

<strong>in</strong>f σ(H) be the <strong>in</strong>fimum of the spectrum of its self-adjo<strong>in</strong>t generator H.<br />

The follow<strong>in</strong>g properties are equivalent:<br />

i) <strong>in</strong>f σ(H) is a simple eigenvalue and there exists a correspond<strong>in</strong>g cyclic<br />

eigenvector <strong>in</strong> H + ;<br />

ii) {Tt : t ≥ 0} is <strong>in</strong>decomposable;<br />

iii) {Tt : t ≥ 0} is ergodic.<br />

We will meet later, <strong>in</strong> Example 2.58 and <strong>in</strong> Corollary 3.4, examples of ergodic<br />

semigroups. Non ergodic Markovian semigroups will appear naturally <strong>in</strong> the<br />

geometric sett<strong>in</strong>g of Chapter 5 (see Remark 5.13).<br />

2.6 Dirichlet Forms and KMS-symmetric, Markovian<br />

Semigroups<br />

A classical result due to A. Beurl<strong>in</strong>g and J. Deny characterizes Markovian<br />

semigroups on Hilbert spaces of square <strong>in</strong>tegrable functions, <strong>in</strong> terms of a<br />

contraction property of the quadratic forms associated to their generators,<br />

with respect to projections onto a suitable convex set.<br />

In this section we will prove an extension of that result <strong>in</strong> the framework of<br />

standard forms of von Neumann algebras. This part is taken from [Cip1]. See<br />

also [GL1,2] for a development of the theory <strong>in</strong> the framework of Haagerup<br />

Lp-spaces. In this section we consider a strongly cont<strong>in</strong>uous, symmetric semigroup<br />

{Tt : t ≥ 0} on the Hilbert space H of a standard form. We will denote<br />

by (H, D(H)) its self-adjo<strong>in</strong>t generator and by (E, F) the associated closed<br />

quadratic form. It will be useful to regard the quadratic form as a lower<br />

semicont<strong>in</strong>uous, convex functional<br />

E : H→(−∞, +∞] so that F := {ξ ∈H: E[ξ] < +∞} .<br />

The quadratic form will be assumed to be J-real <strong>in</strong> the sense that<br />

E[Jξ]=E[ξ] ξ ∈H<br />

and bounded from below <strong>in</strong> the sense that for, some β ∈ R,<br />

E[ξ] ≥ β�ξ� 2<br />

ξ ∈H.


204 F. Cipriani<br />

The doma<strong>in</strong> F is then a J-<strong>in</strong>variant l<strong>in</strong>ear manifold, and it can be shown<br />

that E is determ<strong>in</strong>ed by its restriction to the J-real part F J := F∩H J of its<br />

doma<strong>in</strong>. Notice that the hermitian, sesquil<strong>in</strong>ear form, def<strong>in</strong>ed by polarization,<br />

E : F×F → C E(ξ,η) = 1 � �<br />

E[ξ+η]−E[ξ−η]+iE[ξ+iη]−iE[ξ−iη] ξ,η ∈F,<br />

4<br />

allows us to re-construct the quadratic form as E[ξ] =E(ξ,ξ). As the lower<br />

semicont<strong>in</strong>uity of the quadratic form is equivalent to its closedness, the scalar<br />

products on the doma<strong>in</strong> F<br />

(ξ|η)λ := E(ξ,η)+λ(ξ|η) ξ ∈F,λ>−β.<br />

def<strong>in</strong>e a family of Hilbert structures with equivalent norms given by �ξ� 2 λ =<br />

(ξ|ξ)λ.<br />

To motivate the <strong>in</strong>troduction of the tools needed for the characterizations alluded<br />

to above, let us briefly review the commutative case. There, a Beurl<strong>in</strong>g-<br />

Deny criterion (see [BD2], [Den]) characterizes Markovian semigroups as<br />

those correspond<strong>in</strong>g to quadratic forms which do not <strong>in</strong>crease their values<br />

under the nonl<strong>in</strong>ear map u ↦→ u ∧ 1:<br />

E[u ∧ 1] ≤E[u],<br />

i.e. <strong>in</strong> terms of so called Dirichlet forms.<br />

To extend this criterion to the noncommutative sett<strong>in</strong>g, where, essentially,<br />

the cone L2 +(X, µ) <strong>in</strong>L2 (X, µ) is replaced by the self-polar cone H + <strong>in</strong> H,<br />

one is tempted to replace the constant function 1 by a cyclic vector ξ0 ∈H +<br />

and <strong>in</strong>terprets ξ ∧ ξ0 as <strong>in</strong>f(ξ,ξ0), i.e. <strong>in</strong> terms of a lattice operation. However<br />

(H, H + ) is not a vector lattice apart from the commutative case (see [Io]).<br />

The way to bypass these difficulties can be found by notic<strong>in</strong>g that, <strong>in</strong> the<br />

commutative case, the function u ∧ 1 is the projection of u ∈ L2 R (X, µ) onto<br />

the closed convex set<br />

1 − L 2 R (X, µ) ={v ∈ L2 R (X, µ) :v ≤ 1} .<br />

Def<strong>in</strong>ition 2.49. Let us denote by Proj(ξ,C) the projection of the vector<br />

ξ ∈H J onto the closed convex set C ⊆H J . Then def<strong>in</strong>e, for ξ,η∈H J<br />

ξ ∨ η := Proj(ξ,η + C) and ξ ∧ η := Proj(ξ,η − C) .<br />

We list below, omitt<strong>in</strong>g the proof (see [Cip1], [Io]), the ma<strong>in</strong> properties of<br />

these nonl<strong>in</strong>ear operations.<br />

Lemma 2.50. Let ξ,η∈H J . The follow<strong>in</strong>g properties hold:<br />

i) if sup(ξ,η) (resp. <strong>in</strong>f(ξ,η)) existsthensup(ξ,η) =ξ ∨ η (resp. <strong>in</strong>f(ξ,η) =<br />

ξ ∧ η);


Dirichlet Forms on Noncommutative Spaces 205<br />

ii) ξ ∨ 0=ξ+, ξ ∧ 0=ξ−;<br />

iii) ξ ∨ η = η +(ξ − η)+ and ξ ∧ η = η − (ξ − η)− ;<br />

iv) ξ ∨ η = η ∨ ξ and ξ ∧ η = η ∧ ξ;<br />

v) ξ ∨ η + ξ ∧ η = ξ + η and ξ ∨ η − ξ ∧ η = |ξ − η|;<br />

vi) �ξ ∨ η� 2 + �ξ ∧ η� 2 = �ξ� 2 + �η� 2 .<br />

We may now def<strong>in</strong>e the ma<strong>in</strong> object of a noncommutative potential theory.<br />

Def<strong>in</strong>ition 2.51. (Markovian and Dirichlet forms) Let (M , H , H + ,J)bea<br />

standard form of a von Neumann algebra M and ξ0 ∈ H + separat<strong>in</strong>g vector.<br />

a cyclic and<br />

A J-real quadratic form E : H→(−∞, +∞] issaidtobeMarkovian with<br />

respect to ξ0 if<br />

E[ξ ∧ ξ0] ≤E[ξ] ∀ ξ ∈H J . (2.27)<br />

A closed Markovian form is called a Dirichlet form.<br />

In particular, if E is Markovian and E[ξ] is f<strong>in</strong>ite then E[ξ ∧ ξ0] is f<strong>in</strong>ite too.<br />

In other words, under the Hilbertian projection ξ ↦→ ξ ∧ ξ0, the value of the<br />

quadratic form does not <strong>in</strong>crease. As noticed above, this def<strong>in</strong>ition reduces<br />

to the usual one <strong>in</strong> the commutative sett<strong>in</strong>g. We are go<strong>in</strong>g to see that <strong>in</strong> any<br />

standard form, Dirichlet forms represent an <strong>in</strong>f<strong>in</strong>itesimal characterization of<br />

strongly cont<strong>in</strong>uous, symmetric Markovian semigroups.<br />

Theorem 2.52. (Characterization of Markovian semigroups by Dirichlet<br />

forms) Let (M , H , H + ,J) be a standard form of a von Neumann algebra<br />

M and ξ0 ∈ H + a cyclic vector. Let {Tt : t ≥ 0} be a J-real, symmetric,<br />

contractive, strongly cont<strong>in</strong>uous semigroup on the Hilbert space H and<br />

E : H→[0, +∞] the associated J-real, closed quadratic form. The follow<strong>in</strong>g<br />

properties are equivalent<br />

i) {Tt : t ≥ 0} is Markovian with respect to ξ0;<br />

ii) E is a Dirichlet form with respect to ξ0.<br />

Proof. Let us denote by C the closed convex set ξ0 −H + and by ξC the<br />

projection onto C of a vector ξ ∈H J . Assume property i) and consider ξ ∈<br />

H J .S<strong>in</strong>ceξC ∈ C by def<strong>in</strong>ition and TtξC ∈ C by hypothesis, the properties<br />

of Hilbertian projections then imply that (ξ − ξC|TtξC − ξC) ≤ 0 for all t>0.<br />

In terms of the bounded, positive quadratic forms E t [·] :=t −1 ((I − Tt) ·|·),<br />

this can be written as follows: E t [ξC] ≤E t (ξ,ξC). By the Cauchy-Schwarz<br />

<strong>in</strong>equality we get E t [ξC] ≤E t [ξ] for all t>0 and by the Spectral Theorem,<br />

lett<strong>in</strong>g t → 0, we obta<strong>in</strong> E[ξC] ≤E[ξ].<br />

To prove the converse we start notic<strong>in</strong>g that Markovianity of the semigroup<br />

is equivalent to Markovianity of its resolvent family {Rλ : λ>0} def<strong>in</strong>ed<br />

through Rλ = � ∞<br />

0 e−λt Tt dt for all λ>0: specifically λRλC ⊆ C for all λ>0.<br />

In one direction this follows from the convexity of the set C = ξ −H + and the


206 F. Cipriani<br />

fact that the measures λe−λt dt are probabilities. The opposite � implication<br />

n.<br />

�<br />

n<br />

comes from the well known representation Tt = limn→∞ t R n<br />

t<br />

To show that if ξ ∈ C then ξ := λRλξ ∈ C, consider the functional F : H→<br />

[0, ∞] def<strong>in</strong>ed by<br />

F (η) :=λ −1 E[η]+�η − ξ� 2 .<br />

By the Spectral Theorem one sees<br />

F (η) − F (ξ) =�(λRλ) −1/2 (η − ξ)� 2 ,<br />

so that ξ is the unique m<strong>in</strong>imizer of F . The def<strong>in</strong>ition of ξ C as the nearest<br />

po<strong>in</strong>t to ξ <strong>in</strong> C and the assumption ξ ∈ C imply �ξ C − ξ� ≤�ξ − ξ�. S<strong>in</strong>ce<br />

by assumption E[ξ C] ≤E[ξ] wethenhave<br />

F (ξ C )=λ −1 E[ξ C ]+�ξ C − ξ� 2 ≤ λ −1 E[ξ]+�ξ − ξ� 2 ≤ F (ξ) .<br />

By uniqueness of the m<strong>in</strong>imizer, ξ C = ξ which means ξ ∈ C.<br />

Our second characterization concerns positive semigroups. Apart from its<br />

<strong>in</strong>dependent <strong>in</strong>terest, it will provide an <strong>in</strong>termediate tool for prov<strong>in</strong>g Markovianity.<br />

The result generalizes the so called “second Beurl<strong>in</strong>g-Deny criterion”<br />

(see [BD2], [Den]).<br />

Theorem 2.53. (Quadratic form characterization of positive semigroups)<br />

Let (M, H , H + ,J) be a standard form of the von Neumann algebra M.<br />

Let {Tt : t ≥ 0} be a J-real, symmetric, strongly cont<strong>in</strong>uous, semigroup<br />

on the Hilbert space H and E : H→(−∞, +∞] the associated J-real, closed<br />

quadratic form. The follow<strong>in</strong>g properties are equivalent<br />

i) {Tt : t ≥ 0} is positive;<br />

ii) E[|ξ|] ≤E[ξ] for all ξ ∈H J .Inparticularifξ ∈F J than |ξ| ∈F J ;<br />

iii) ξ ∈F j implies ξ+,ξ− ∈F and E(ξ+,ξ−) ≤ 0.<br />

Proof. Assum<strong>in</strong>g the property i), the follow<strong>in</strong>g identity<br />

E t [ξ+ − ξ−] −E t [ξ+ + ξ−] =4(ξ+|Ttξ−) − 4(ξ+|ξ−) =4(ξ+|Ttξ−)<br />

establishes property ii) for the family of bounded quadratic forms E t depend<strong>in</strong>g<br />

on t>0. Lett<strong>in</strong>g t approach zero we get the property ii) for E. Toprove<br />

the converse we still go via positivity of the resolvent, this property be<strong>in</strong>g<br />

equivalent to positivity of the semigroup by reason<strong>in</strong>g similar to that used <strong>in</strong><br />

the previous theorem and based on convexity.<br />

To prove that for each sufficiently large λ we have Rλ(H + ) ⊆H + ,notice<br />

that the resolvent map Rλ maybeseenastheadjo<strong>in</strong>tI ∗ : H→Fof the<br />

embedd<strong>in</strong>g I : F→Hwhen the doma<strong>in</strong> F is endowed with the graph norm<br />

�·�λ. Letξ be <strong>in</strong> Rλ(H + ). To prove that ξ ∈H + we will show that ξ = |ξ|.<br />

S<strong>in</strong>ce by the Spectral Theorem Rλ(H) ⊆F,wehavethatξ ∈F J . Hence by


Dirichlet Forms on Noncommutative Spaces 207<br />

hypothesis ii) we have |ξ| ∈F J and �|ξ|�λ ≤�ξ�λ. Suppose that η ∈H + is<br />

such that ξ = Rλη = I ∗ η.Then<br />

(|ξ||ξ)λ =(|ξ||I ∗ η)λ =(|ξ||η) ≥|(ξ | η)| = |(ξ | I ∗ η)λ| = �ξ� 2 λ ,<br />

so that by the Cauchy-Schwarz <strong>in</strong>equality we obta<strong>in</strong><br />

�ξ� 2 λ ≤ (|ξ||ξ)λ ≤�|ξ|�λ ·�ξ�λ ≤�ξ� 2 λ<br />

and thus |ξ| = ξ. The equivalence between ii) and iii) easily follows from the<br />

identities 2ξ± = |ξ|±ξ and E[|ξ|]−E[ξ] =E[ξ++ξ−]−E[ξ+−ξ−] =4E(ξ+,ξ−).<br />

The application of the criterion <strong>in</strong> Theorem 2.53 above is frequently simplified<br />

by the follow<strong>in</strong>g situation.<br />

Theorem 2.54. Let E : H→(−∞, +∞] be a non-negative, J-real closed<br />

quadratic form such that<br />

• i) E is non-negative,<br />

• ii) ξ0 ∈F and E(ξ0,ξ) ≥ 0 for all ξ ∈F∩H + ,<br />

• iii) E[|ξ|] ≤E[ξ] ∀ ξ ∈HJ .<br />

Then E is a Dirichlet form with respect to ξ0.<br />

Proof. S<strong>in</strong>ce, by hypothesis i), ξ0 ∈Fand, by Lemma 2.27 iii), ξ ∧ ξ0 =<br />

ξ0 − (ξ − ξ0)−, the Markovianity property E[ξ ∧ ξ0] ≤E[ξ] ofE can re-written<br />

as follows:<br />

E[ξ0 − (ξ − ξ0)−] ≤E[ξ0 +(ξ − ξ0)]<br />

E[ξ0] − 2E(ξ0, (ξ − ξ0)−)+E[(ξ − ξ0)−] ≤E[ξ0] − 2E(ξ0,ξ− ξ0)+E[ξ − ξ0]<br />

−2E(ξ0, (ξ − ξ0)+) ≤E[ξ − ξ0] −E[(ξ − ξ0)+] .<br />

S<strong>in</strong>ce by hypothesis ii) the left hand side of the last <strong>in</strong>equality is negative, to<br />

prove the result it is enough to show that the right hand side is positive, i.e.<br />

E[η+] ≤E[η] for all η ∈H J ,oreven<br />

E[η+] ≤E[η+]+E[η−] − 2E(η+,η−) .<br />

This follows by assumptions i), iii) and the equivalence <strong>in</strong> Theorem 2.30<br />

between statement ii) and iii).<br />

Example 2.55. (Bounded Dirichlet forms) Let (M , H , H + ,J) be a standard<br />

form of the von Neumann algebra M and ξ0 ∈H + a cyclic vector. Consider,<br />

for a set of elements ak ∈ M and positive numbers µk ,νk > 0 , k =1,...,n,<br />

the operators<br />

dk : H→H dk := i(µkak − νkj(a ∗ k)) ,


208 F. Cipriani<br />

the quadratic form<br />

E[ξ] :=<br />

n�<br />

�dkξ� 2<br />

k=1<br />

and its associated symmetric, contractive semigroup {Tt : t ≥ 0}. Thenwe<br />

have<br />

• i) E is J-real if and only if<br />

n�<br />

[µ 2 ka∗k ak − ν 2 kaka ∗ k ] ∈ M ∩ M′ ;<br />

k=1<br />

• ii) if E is J-real then {Tt : t ≥ 0} is positive;<br />

• iii) if E is J-real then E is ξ0-Markovian if and only if<br />

n�<br />

k=1<br />

[µ 2 ka ∗ kak − µkνk(akj(ak)+a ∗ kj(a ∗ k)) + ν 2 kaka ∗ k]ξ0 ∈H + ;<br />

• iv) {Tt : t ≥ 0} is conservative (Ttξ0 = ξ0 for all t ≥ 0) if and only if the<br />

numbers (µk/νk) 2 , k =1,...,n, are eigenvalues of the modular operator<br />

∆ξ0, correspond<strong>in</strong>g the eigenvectors akξ0. In particular, this is the case<br />

when µk = νk for all k =1,...,n and the elements ak belongs to the<br />

centralizer of the state ω0(·) =(ξ0|·ξ0) associated to ξ0,<br />

To verify i) notice that<br />

and<br />

Mω0 := {x ∈ M : ω0(xy) =ω0(yx) ∀ y ∈ M} .<br />

E[ξ] =<br />

E[Jξ]=<br />

n� �<br />

ξ|(µka ∗ k − νkj(ak))(µkak − νkj(a ∗ k ))ξ�<br />

k=1<br />

n� �<br />

ξ|(νkak − µkj(a ∗ k ))(ν∗ kak − µkj(ak))ξ � ,<br />

k=1<br />

so that E is J-real if and only if<br />

n�<br />

k=1<br />

ν 2 k(aka ∗ k + j(a ∗ kak)) =<br />

n�<br />

k=1<br />

µ 2 k(a ∗ kak + j(aka ∗ k)) .<br />

Rearrang<strong>in</strong>g this last equation as �n k=1 [µ2 ka∗k ak−ν 2 kaka∗ k ]=j(� n<br />

k=1 [µ2k a∗k ak−<br />

ν2 kaka∗ k ]) we get the desired condition <strong>in</strong> item i), as a = j(a) fora∈M if and<br />

only if a belongs to the center M ∩ M ′ of M.<br />

Assum<strong>in</strong>g E to be J-real, to prove the positivity property of the semigroup,<br />

we are go<strong>in</strong>g to verify condition iii) <strong>in</strong> Theorem 2.53. Let ξ± ∈H + be two


Dirichlet Forms on Noncommutative Spaces 209<br />

positive orthogonal vectors. Denote by s ′ ± ∈ M ′ the support projection <strong>in</strong><br />

M ′ of the vectors ξ±: these are, by def<strong>in</strong>ition, the projection onto the closure<br />

of the subspaces M ′ ξ± (see [Ara]). S<strong>in</strong>ce s ′ + and s′ − are orthogonal, s′ + s′ − =<br />

s ′ −s ′ + = 0, and they commute with the operators a1 ...,an ∈ M we have<br />

E(ξ+,ξ−) =<br />

n�<br />

k=1<br />

= −<br />

≤ 0<br />

((µkak − νkj(a ∗ k))ξ+|µk(ak − νkj(a ∗ k))ξ−)<br />

n�<br />

k=1<br />

µkνk((akj(ak)+a ∗ k j(a∗ k ))ξ+|ξ−)<br />

by Proposition 2.13 iv).<br />

Assum<strong>in</strong>g E to be J-real, to prove Markovianity of E, we may use Theorem<br />

2.54: s<strong>in</strong>ce the form is nonnegative by construction, and condition iii) of<br />

Theorem 2.54 is satisfied, by the above calculation and by Theorem 2.53,<br />

what is left to be shown is that ξ0 is excessive for E: E(ξ,ξ0) ≥ 0 for all<br />

ξ ∈H + (<strong>in</strong> this case the form doma<strong>in</strong> co<strong>in</strong>cides with the whole of H). A<br />

direct calculation gives<br />

E(ξ,ξ0) =<br />

n�<br />

k=1<br />

� � n� �<br />

= ξ�<br />

(ξ|d ∗ k dkξ0)<br />

k=1<br />

[µ 2 ka ∗ kak − µkνk(akj(ak)+a ∗ kj(a ∗ k)) + ν 2 kj(aka ∗ k)]ξ0<br />

S<strong>in</strong>ce H + is self-dual, the vector ξ0 is excessive for E if and only if the vector<br />

� n<br />

k=1 [µ2 k a∗ k ak − µkνk(akj(ak)+a ∗ k j(a∗ k )) + ν2 k aka ∗ k ]ξ0 is <strong>in</strong> H + .<br />

Conservativity of the semigroup generated by E is clearly equivalent to the<br />

fact that ξ0 belongs to the kernel of the generator and then, by the Spectral<br />

Theorem, to E[ξ0] = 0. This means that dkξ0 =0forallk =1,...,n.Bythe<br />

def<strong>in</strong>ition of the operators dk this happens if and only if<br />

µkakξ0 = νkj(a ∗ k )ξ0<br />

k =1,...,n.<br />

Us<strong>in</strong>g the properties of the modular operators, we may re-write this equation<br />

<strong>in</strong> the form<br />

µkakξ0 = νkj(a ∗ k )ξ0 = νkJa ∗ k ξ0 = νkJJ∆ 1/2<br />

ξ0 akξ0 = νk∆ 1/2<br />

ξ0 akξ0 ,<br />

from which the statement <strong>in</strong> item iv) above follows. When all the coefficients<br />

ak belong to the centralizer of ω0, they strongly commute with the modular<br />

operator ∆ξ0 and the last equation simplifies as follows<br />

µkakξ0 = νk∆ 1/2<br />

ξ0 akξ0 = νkak∆ 1/2<br />

ξ0 ξ0 = νkakξ0 .<br />

�<br />

.


210 F. Cipriani<br />

This is verified if and only if µk = νk for all k =1,...,n.<br />

Notice that, <strong>in</strong> this example, the generator H has the form<br />

H =<br />

n�<br />

k=1<br />

µ 2 ka∗k ak − µkνk(akj(ak)+a ∗ kj(a∗k )) + ν2 kj(aka ∗ k ) ,<br />

similar to that of the L<strong>in</strong>dblad generator of a dynamical, uniformly cont<strong>in</strong>uous<br />

semigroup (see Example 2.26). The similarity is even more evident if, beside<br />

the (say left) action of M on H considered so far,<br />

H×M ∋ (ξ,x) ↦→ ξx ∈H,<br />

we also consider the right one, def<strong>in</strong>ed by<br />

H×M ∋ (ξ,x) ↦→ ξx := j(x ∗ )ξ = Jx ∗ Jξ ∈H.<br />

Us<strong>in</strong>g this action, we have <strong>in</strong> fact that<br />

Hξ =<br />

n�<br />

k=1<br />

µ 2 ka ∗ kakξ − µkνk(akξa ∗ k + a ∗ kξak)+ν 2 kξaka ∗ k .<br />

Notice, moreover, that the maps dk : H → H, by which we constructed<br />

the Dirichlet form E are covariant derivations with respect to the bi-module<br />

action above. In fact, consider<strong>in</strong>g the derivations of M given by<br />

δb : M → M δb(a) :=i[b, a] =i(ba − ab) a, b ∈ M ,<br />

the follow<strong>in</strong>g Leibniz rules hold true:<br />

dk(aξ) =δµiai(a)ξ + adk(ξ) , dk(ξa) =dk(ξ)a + ξδνiai(a) .<br />

Compare this with Chapter 5, equation 5.9. In the tracial case constructions<br />

similar to those above have been considered <strong>in</strong> [GIS] to generate non symmetric<br />

Dirichlet forms and non symmetric Markovian semigroups.<br />

Example 2.56. (An unbounded Dirichlet form) Let (a, D(a)) be a self-adjo<strong>in</strong>t<br />

operator, affiliated to the von Neumann algebra M, and consider the operator<br />

def<strong>in</strong>ed by<br />

da : D(da) →H da := i(a − j(a)) , D(da) :=D(a) ∩ JD(a) ,<br />

together with the positive, J-real quadratic form<br />

E[ξ] :=�daξ� 2 , F = D(da) .<br />

Notice that, s<strong>in</strong>ce a and j(a) commute strongly, the operator da and the<br />

quadratic form E are closable on the doma<strong>in</strong> F = D(da). We are go<strong>in</strong>g


Dirichlet Forms on Noncommutative Spaces 211<br />

to show that if a is affiliated to the centralizer Mω0 then E is Markovian.<br />

This will be done by a monotone approximation procedure. By Haagerup’s<br />

representation lemma (see [H1]), for each fixed vector ξ ∈Hthere exists a<br />

unique, positive, bounded Borel measure ν supported on Sp(a) × Sp(a) such<br />

that �<br />

f(s)g(t) ν(ds, dt) =(ξ| f(a)j(g(a))ξ)<br />

Sp(a)×Sp(a)<br />

for cont<strong>in</strong>uous functions f and g on the spectrum Sp(a) ⊆ R of a. Us<strong>in</strong>gthis<br />

measure we have the follow<strong>in</strong>g representation<br />

�<br />

E[ξ] =<br />

|s − t| 2 ν(ds, dt) .<br />

Sp(a)×Sp(a)<br />

Consider the sequence {fn ∈ Cb(R) :n ∈ N ∗ } of bounded, cont<strong>in</strong>uous functions<br />

def<strong>in</strong>ed by fn(t) =t for t ∈ [−n, +n], while fn(t) =n · sgn(t) otherwise,<br />

and the sequence of bounded quadratic forms:<br />

En[ξ] =�fn(a)ξ − ξfn(a)� 2 .<br />

S<strong>in</strong>ce fn(a) ∈ Mω0, by the previous example, we have that theEn are Dirichlet<br />

forms. Us<strong>in</strong>g the measure ν we may represent these forms as follows:<br />

�<br />

En[ξ] =<br />

|fn(s) − fn(t)| 2 ν(ds, dt) .<br />

Sp(a)×Sp(a)<br />

S<strong>in</strong>ce, as is easy to verify, |fn(s)−fn(t)| ≤|fn+1(s)−fn+1(t)| for all (s, t) ∈ R 2<br />

and n ∈ N ∗ ,wehavethatthequadraticformE is monotone the limit of the<br />

sequence of Dirichlet forms En:<br />

E[ξ] =supEn[ξ]<br />

∀ ξ ∈H.<br />

n<br />

Markovianity of E follows from the Markovianity of the approximat<strong>in</strong>g forms<br />

En. The closure of E is then closed and Markovian, so that it is a Dirichlet<br />

form.<br />

Example 2.57. (Quantum Ornste<strong>in</strong>-Uhlenbeck semigroup, Quantum Brownian<br />

motion, Birth-and-Death process) We describe here the construction of a<br />

Dirichlet form on a type I factor von Neumann algebra B(h) which generates<br />

a Markovian semigroup of those types appear<strong>in</strong>g <strong>in</strong> quantum optics (see [D1]).<br />

Let us consider, on the Hilbert space h := l2 (N), the von Neumann algebra<br />

B(h) and its Hilbert-Schmidt standard form<br />

(B(h) , L 2 (h) , L 2 +(h) ,J)


212 F. Cipriani<br />

discussed <strong>in</strong> Example 2.17. Let {en : n ≥ 0} ⊂h be the canonical Hilbert<br />

basis, and denote by | em〉〈en|, n, m ≥ 0, the partial isometries, hav<strong>in</strong>g Cen<br />

has <strong>in</strong>itial space and Cem as f<strong>in</strong>al one.<br />

For a pair of fixed parameters µ>λ>0, set ν := λ 2 /µ 2 and let<br />

φν(x) :=Tr(ρνx) x ∈B(h) ,<br />

the normal state on B(h) represented by the trace-class operator<br />

ρν := (1 − ν) �<br />

ν n |en〉〈en| .<br />

n≥0<br />

The associated cyclic vector ξν <strong>in</strong> L 2 +(h) isthengivenby<br />

ξν = ρ 1/2<br />

ν<br />

=(1− ν)<br />

1/2 �<br />

n≥0<br />

ν n/2 |en〉〈en| ,<br />

so that φν(x) =〈ξν| xξν〉. Thecreation and annihilation operators a∗ and a<br />

on h are def<strong>in</strong>ed by<br />

a ∗ en := √ � √<br />

nen−1 if n>0;<br />

n +1en+1 , aen :=<br />

0 if n =0.<br />

They are adjo<strong>in</strong>t to one another on their common doma<strong>in</strong><br />

D(a) =D(a ∗ )= � e ∈ h : � √<br />

n|〈e| en〉| 2 < ∞ � ,<br />

n≥0<br />

and satisfy the Canonical Commutation Relation<br />

The quadratic form given by<br />

aa ∗ − a ∗ a = I.<br />

E[ξ] :=�µaξ − λξa ∗ � 2 + �µaξ ∗ − λξ ∗ a ∗ � 2<br />

is well def<strong>in</strong>ed on the dense subspace of L 2 (h) def<strong>in</strong>ed as<br />

D(E) := l<strong>in</strong>ear span{| em〉〈en| ,n,m≥ 0} .<br />

In fact, it is easy to check that for ξ ∈ D(E), the operators aξ, a ∗ ξ, ξa and<br />

ξa ∗ are bounded. It is then possible to prove that the closure of (E ,D(E)) is a<br />

Dirichlet form with respect to ξν, generat<strong>in</strong>g the so called quantum Ornste<strong>in</strong>-<br />

Uhlenbeck Markovian semigroup (see [CFL]).<br />

Moreover, it is easy to check that E vanishes only on the multiples of the vector<br />

ξν, so that, by Corollary 2.48, the associated quantum Ornste<strong>in</strong>-Uhlenbeck<br />

semigroup is ergodic. In [CFL] it was shown also that, the semigroup has the


Dirichlet Forms on Noncommutative Spaces 213<br />

Feller property, <strong>in</strong> the sense that it leaves <strong>in</strong>variant the algebra of compact<br />

operators K(h), and on this C∗-algebra the semigroup is strongly cont<strong>in</strong>uous.<br />

This example exhibits one of the characteristic features of Noncommutative<br />

or Quantum <strong>Probability</strong>, namely, the possibility of accommodat<strong>in</strong>g different<br />

classical processes <strong>in</strong> a s<strong>in</strong>gle quantum process. This po<strong>in</strong>t is illustrated <strong>in</strong><br />

Philippe Biane’s Lectures <strong>in</strong> this volume.<br />

In fact, the quantum Ornste<strong>in</strong>-Uhlenbeck semigroup leaves different maximal<br />

abelian subalgebras of B(h) globally <strong>in</strong>variant ([Ped 2.8]): reduc<strong>in</strong>g it to a<br />

suitable pair of them, one may obta<strong>in</strong>, for example, the semigroups of the<br />

classical Ornste<strong>in</strong>-Uhlenbeck and Birth-and-Death processes (see [CFL]).<br />

When λ = µ, the role of the <strong>in</strong>variant state φν has to be played by the normal,<br />

semif<strong>in</strong>ite trace τ on B(h). However, even <strong>in</strong> this case, us<strong>in</strong>g the theory<br />

developed by [AHK1] (see Chapter 4), it is possible to prove that the closure<br />

of the quadratic form<br />

E[ξ] :=�aξ − ξa ∗ � 2 + �aξ ∗ − ξ ∗ a ∗ � 2<br />

ξ ∈ D(E)<br />

is a Dirichlet form. The associated τ-symmetric Markovian semigroup on<br />

B(h) maybedilated by a Quantum Stochastic processes, known as Quantum<br />

Brownian motion. On a suitable, <strong>in</strong>variant, maximal abelian subalgebra, the<br />

dynamical semigroup reduces to the heat semigroup of a classical Brownian<br />

motion (see [CFL]).<br />

A whole family of Quantum Ornste<strong>in</strong>-Uhlenbeck semigroups, Markovian with<br />

respect to suitable quasi-free states on the CCR algebra, has been <strong>in</strong>troduced<br />

and deeply <strong>in</strong>vestigated <strong>in</strong> [KP].<br />

We would like to illustrate now a construction due to Y.M. Park [P1,3] which,<br />

though similar to that of Example 2.55, is better suited for applications to<br />

Quantum Statistical Mechanics (see Chapter 3).<br />

Def<strong>in</strong>ition 2.58. (Admissible functions) Let us denote by Iλ, thestrip<br />

Iλ := {z ∈ C : |Imz| ≤λ} λ>0 ,<br />

and consider the subspace Mλ of elements x ∈ M for which the map R ∋<br />

t ↦→ σt(x) ∈ M can be extended to an analytic map <strong>in</strong> a doma<strong>in</strong> conta<strong>in</strong><strong>in</strong>g<br />

the strip Iλ.<br />

An analytic function f : D → C, def<strong>in</strong>ed <strong>in</strong> a doma<strong>in</strong> D conta<strong>in</strong><strong>in</strong>g the strip<br />

I1/4, is called admissible if the follow<strong>in</strong>g properties hold true:<br />

• i) f(t) ≥ 0 for all t ∈ R,<br />

• ii) f(t + i/4) + f(t − i/4) ≥ 0, for all t ∈ R<br />

• iii) there exist M>0andp>1, such that the bound<br />

|f(t + is)| ≤M(1 + |t|) −p<br />

holds uniformly for s ∈ [−1/4 , +1/4].


214 F. Cipriani<br />

For example, the function<br />

� �<br />

2<br />

f(t) := (e<br />

π R<br />

k/4t + e −k/4 ) −1 e −k2 /2 −itk<br />

e dk t ∈ R ,<br />

is admissible for all p>1. We will also consider the function f0 : R → R<br />

given by<br />

f0(t) :=<br />

t ∈ R .<br />

1<br />

cosh(2πt) =2(e2πt + e −2πt ) −1<br />

This function, which plays a dist<strong>in</strong>guished role <strong>in</strong> the proof of the Tomita-<br />

Takesaki Theorem (see [Br2], [T1], [V]), is almost admissible <strong>in</strong> the sense that<br />

it has an analytic extension to the <strong>in</strong>terior of I1/4, denoted f0 aga<strong>in</strong>, and, on<br />

the boundary of I1/4, it is a distribution satisfy<strong>in</strong>g<br />

f0(t + i/4) + f0(t − i/4) = δ(t) .<br />

Admissible functions will play the role of coefficients for the class of Dirichlet<br />

forms we are go<strong>in</strong>g to discuss.<br />

Proposition 2.59. ([P1,3]) Let (M , H , H + ,J) be a standard form of the<br />

von Neumann algebra M and ξ0 ∈H + acyclicvector.Considerx∈M1/4, and let f represent an admissible function or equal the function f0. Then the<br />

quadratic form E : H→[0, +∞), def<strong>in</strong>ed by<br />

�<br />

E[ξ] :=<br />

is a Dirichlet form.<br />

�(σt−i/4(x) − j(σt−i/4(x R<br />

∗ )))ξ� 2 f(t) dt ξ ∈H, (2.28)<br />

Proof. In this proof, for the sake of clarity, we will denote the scalar product<br />

<strong>in</strong> H by angle brackets 〈·| ·〉. We will consider the case f = f0 only. The proof<br />

<strong>in</strong> case of admissible functions is similar. Notice that, by the assumption<br />

x ∈ M 1/4, wehave(seeAppendix7.3)<br />

σ t−i/4(x) =σt(σ −i/4(x)) t ∈ R , (2.29)<br />

so that<br />

�σt−i/4(x)� = �σ−i/4(x)� t ∈ R<br />

and E is well def<strong>in</strong>ed on the whole H, asf0 is <strong>in</strong>tegrable. To apply Theorem<br />

2.53, we are go<strong>in</strong>g to verify the property appear<strong>in</strong>g there <strong>in</strong> item iii). Consider<br />

two orthogonal, positive elements ξ± ∈ H + , and decompose E(ξ+| ξ−) as<br />

follows:<br />

E(ξ+| ξ−) =E1(ξ+| ξ−)+E2(ξ+| ξ−)+E ∗ 1 (ξ+| ξ−)+E ∗ 2 (ξ+| ξ−) , (2.30)


Dirichlet Forms on Noncommutative Spaces 215<br />

where<br />

� �<br />

E1(ξ+| ξ−) :=<br />

R<br />

� �<br />

E2(ξ+| ξ−) :=−<br />

R<br />

〈σ t−i/4(x)ξ+| σ t−i/4(x)ξ−〉 +<br />

〈σ t−i/4(x ∗ )ξ−| σ t−i/4(x ∗ )ξ+〉<br />

〈σ t−i/4(x)ξ+| j(σ t−i/4(x ∗ ))ξ−〉 +<br />

〈j(σ t−i/4(x ∗ ))ξ+| σ t−i/4(x)ξ−〉<br />

�<br />

f0(t)dt ,<br />

�<br />

f0(t)dt ,<br />

(2.31)<br />

(2.32)<br />

and E ∗ 1 (ξ+| ξ−), E ∗ 2 (ξ+| ξ−) are obta<strong>in</strong>ed from E1(ξ+| ξ−), E2(ξ+| ξ−), respectively,<br />

replac<strong>in</strong>g x by x ∗ . Reason<strong>in</strong>g as <strong>in</strong> Example 2.55, we get<br />

E1(ξ+| ξ−) =E ∗ 1 (ξ+| ξ−) =0<br />

as ξ+ and ξ− have orthogonal support projection <strong>in</strong> M ′ and σt−i/4(x) ,σt−i/4 (x∗ ) ∈ M. To evaluate the second and fourth part <strong>in</strong> expression 2.30, notice<br />

that<br />

σt−i/4(x) ∗ = σt+i/4(x ∗ ) , j(σt+i/4(x) ∗ )=σt−i/4(j(x)) .<br />

We then have<br />

� �<br />

E2(ξ+| ξ−) =− 〈ξ+| σt+i/4(x R<br />

∗ j(x ∗ ))ξ−〉 +<br />

�<br />

(2.33)<br />

〈ξ+| σt−i/4(xj(x))ξ−〉 f0(t)dt ,<br />

E ∗ � �<br />

2 (ξ+| ξ−) =− 〈ξ+| σt+i/4(xj(x))ξ−〉 +<br />

R<br />

〈ξ+| σt−i/4(x ∗ j(x ∗ �<br />

))ξ−〉 f0(t)dt ,<br />

(2.34)<br />

�<br />

E(ξ+| ξ−) =−<br />

�<br />

〈ξ+| (σi/4 + σ−i/4)(σt(xj(x)+x R<br />

∗ j(x ∗ �<br />

)))ξ−〉 f0(t)dt .<br />

(2.35)<br />

Sett<strong>in</strong>g<br />

�<br />

I0(y) := σt(y)f0(t) dt y ∈ M ,<br />

we may re-write (2.35) as follows:<br />

R<br />

E(ξ+| ξ−) =−〈ξ+| (σ i/4 + σ −i/4)(I0(xj(x)+x ∗ j(x ∗ )))ξ−〉 . (2.36)<br />

Notice that, <strong>in</strong> the formulas above, the modular automorphism group is understood<br />

to act, not only on M, but on the whole algebra B(H) ofbounded<br />

operators on H. It may be shown (see [P1,3], [Br2 Theorem 2.5.14], [T1<br />

Lemma 3.10], [V Lemma 4.1]) that


216 F. Cipriani<br />

�<br />

I0 (σi/4 + σ−i/4)(y) � = y ∀ y ∈ B(H) .<br />

By Proposition 2.13 iv), we then have<br />

E(ξ+| ξ−) =−〈ξ+| (xj(x)+x ∗ j(x ∗ ))ξ−〉 ≤0 . (2.37)<br />

Apply<strong>in</strong>g Theorem 2.53, the semigroup generated by E is positive, and that<br />

the follow<strong>in</strong>g contraction property holds true:<br />

E[|ξ|] ≤E[ξ] ξ ∈H.<br />

As E is, by construction, nonnegative, apply<strong>in</strong>g Theorem 2.54, we then have<br />

that E is a Dirichlet form, provided we verify it vanishes on ξ0. This follows<br />

from the identities<br />

(σ t−i/4(x) − j(σ t−i/4(x ∗ )))ξ0<br />

= ∆ it+1/4 x∆ −it−1/4 ξ0 − J∆ it+1/4 x ∗ ∆ −it−1/4 Jξ0<br />

= i0(σt(x)) − J(i0(σt(x ∗ )))<br />

= i0(σt(x)) − J(i0(σt(x) ∗ ))<br />

= i0(σt(x)) − i0(σt(x))<br />

=0.<br />

Apply<strong>in</strong>g essentially the construction above, we describe next a class of<br />

Dirichlet forms on the CCR and CAR algebras, studied <strong>in</strong> [BKP1,2]. The<br />

correspond<strong>in</strong>g semigroups are �ω-KMS-symmetric with respect to the normal<br />

extensions �ω of the quasi-free KMS-states ω <strong>in</strong>troduced <strong>in</strong> Example 2.29 and<br />

Example 2.30.<br />

Example 2.60. (Dirichlet forms and quasi-free states on CAR algebras)With<strong>in</strong><br />

the framework of Example 2.29, let (πω , Hω ξω) be the GNS-representation<br />

of the CAR algebra U(h0) with respect to the quasi-free state ω, def<strong>in</strong>ed by<br />

the Hamiltonian operator H at the <strong>in</strong>verse temperature β. Strengthen<strong>in</strong>g,<br />

the previous assumptions made on h0, i.e. assum<strong>in</strong>g that<br />

• h0 is a dense subspace of analytic elements for the unitary group R ∋ t ↦→<br />

eitH ,<br />

• e izH ξ ∈ h0 for all ξ ∈ h0 and z ∈ C,<br />

one obta<strong>in</strong>s that, for f ∈ h0, a(f) is entire analytic with respect to the<br />

modular group of ω.<br />

The construction of Proposition 2.59 will be, <strong>in</strong> this particular situation,<br />

modified so ad to take <strong>in</strong>to account the natural Z2-grad<strong>in</strong>g, given by the<br />

<strong>in</strong>volutive, ∗-automorphism γ : U(h0) → U(h0) γ(a(f)) := −a(f) f ∈ h0 .


Dirichlet Forms on Noncommutative Spaces 217<br />

As γ leaves ω <strong>in</strong>variant, by Proposition 2.16, there exists a positive, unitary<br />

operator Uγ : Hω →Hω, such that γ(x) =UγxU −1<br />

γ , x ∈ Mω, Uγξω = ξω and<br />

U 2 γ = I. It is easy to check that Uγxξω = γ(x)ξω for all x ∈ Mω.<br />

Consider now a complete, orthonormal system {gk : k ≥ 1} ⊂h0 and an admissible<br />

function f, normalized so to have <strong>in</strong>tegral one over R. Thequadratic<br />

form E : Hω → [0, +∞]<br />

E[ξ] := �<br />

�<br />

�(σt−i/4(a(gk)) − j(σt−i/4(a k≥1<br />

R<br />

∗ (gk))))Uγξ� 2 f(t) dt ξ ∈Hω ,<br />

(2.38)<br />

is then a Dirichlet form and the associated Markovian semigroup commutes<br />

with the symmetry γ. It can also be shown that the semigroup is <strong>in</strong>dependent<br />

of the choices of the orthonormal basis so that, <strong>in</strong> particular, it commutes<br />

with the modular group. Moreover, the semigroup is ergodic and <strong>in</strong>dependent<br />

of the choice of the normalized admissible function.<br />

This construction applies, for example, to the “Ideal Fermi Gas” where<br />

h := L2 (Rd , dx), h0 is the subspace of functions hav<strong>in</strong>g compactly supported<br />

Fourier transform, H := −H0 − µ · I, H0 denot<strong>in</strong>g the Laplacian operator on<br />

Rd , µ ∈ R and β>0.<br />

Example 2.61. (Dirichlet forms and quasi-free states on CCR algebras) As <strong>in</strong><br />

Example 2.30, let (πω , Hω ξω) be the GNS-representation of the CCR C ∗ -<br />

algebra U(h0) with respect to the quasi-free state ω, def<strong>in</strong>ed by the Hamiltonian<br />

operator H and by the <strong>in</strong>verse temperature β>0. Denote by<br />

Wω(f) :=πω(W (f)) f ∈ h0 ,<br />

the Weyl operators <strong>in</strong> the GNS-representation. As the functions R ∋ t ↦→<br />

ω(W (tf)) are cont<strong>in</strong>uous, for any fixed f ∈ h0, thestateω is regular and the<br />

unitary groups R ∋ t ↦→ Wω(tf) are strongly cont<strong>in</strong>uous (see [BR2 5.2.3]).<br />

Denote by Φω(f) their self-adjo<strong>in</strong>t generators<br />

Wω(tf) =e itΦω(f)<br />

t ∈ R ,<br />

and def<strong>in</strong>e the annihilation and creation operators as follows<br />

aω(f) := Φω(f)+iΦω(if)<br />

√<br />

2<br />

a ∗ ω(f) := Φω(f) − iΦω(if)<br />

√<br />

2<br />

,<br />

.<br />

(2.39)<br />

These are densely def<strong>in</strong>ed, closed operators, adjo<strong>in</strong>t to one another, satisfy<strong>in</strong>g<br />

the Canonical Commutation Relations. Strengthen<strong>in</strong>g aga<strong>in</strong>, the stand<strong>in</strong>g<br />

assumptions made on h0, exactly as we did <strong>in</strong> Example 2.59, one may verify<br />

that


218 F. Cipriani<br />

• the subspace Mσξω is dense <strong>in</strong> the Hilbert space Hω,<br />

• the functions R ∋ t → σt(a(f))ξ and R ∋ t → σt(a∗ (f))ξ admit entire<br />

analytic extensions for all f ∈ h0, ξ ∈ Mσξω<br />

• on vectors <strong>in</strong> Mσξω, these analytic extensions are given by<br />

σz(a(f)) = a(e izH f) σz(a ∗ (f)) = a ∗ (e izH f) f ∈ h0 .<br />

Consider now a complete, orthonormal system {gk : k ≥ 1} ⊂h0 and a<br />

normalized admissible function f. The quadratic form E : Hω → [0, +∞]<br />

E[ξ] := �<br />

�<br />

k≥1<br />

�(σt−i/4(a(gk)) − j(σt−i/4(a R<br />

∗ (gk))))ξ� 2 f(t) dt ξ ∈Hω<br />

(2.40)<br />

is then a Dirichlet form which is <strong>in</strong>dependent of the choice of orthonormal<br />

basis and the normalized, admissible function. Moreover, the correspond<strong>in</strong>g<br />

semigroup is ergodic.<br />

This construction applies, for example, to the “Ideal Bose Gas” where<br />

h := L2 (Rd , dx), h0 is the subspace of functions hav<strong>in</strong>g compactly supported<br />

Fourier transform, H := −H0 + µ · I, H0 denot<strong>in</strong>g the Laplacian operator on<br />

Rd , µ>0andβ>0.<br />

3 Dirichlet Forms <strong>in</strong> Quantum Statistical Mechanics<br />

In this section we apply the theory developed so far to the construction of<br />

Dirichlet forms <strong>in</strong> the framework of Quantum Statistical Mechanics. We illustrate<br />

here a construction due to Y.M. Park [P1], [P2]. The reader may<br />

consult the work of A. Majewski-B. Zegarl<strong>in</strong>ski [MZ1,2] and A. Majewski-R.<br />

Olkiewicz-B. Zegarl<strong>in</strong>ski [MOZ], for alternative approaches based on generalized<br />

conditional expectations.<br />

The problem to face is the dynamical approach to equilibrium. Consideratime<br />

evolution of a quantum system, hence a cont<strong>in</strong>uous group of automorphisms<br />

{αt : t ∈ R} of a C ∗ or von Neumann algebra A. Atanyfixed<strong>in</strong>verse<br />

temperature β, thesetof(α, β)-KMS states, represent<strong>in</strong>g the equilibria of<br />

the dynamical system, are the solution of the KMS equation. For each fixed<br />

(α,β)-KMS state, one would like to construct a dissipative dynamics, hence a<br />

cont<strong>in</strong>uous semigroup {Φt : t ≥ 0} on A, ergodic and and admitt<strong>in</strong>g ω as the<br />

unique <strong>in</strong>variant state. In this section a partial answer to this problem will<br />

be provided, through an L 2 approach based on the construction of a suitable<br />

Dirichlet form. The discussion of the commutative case may be found <strong>in</strong> [Lig]<br />

while constructions of semigroups <strong>in</strong> the framework of Quantum Statistical<br />

Mechanics may found <strong>in</strong> [Mat1], [Mat2], [Mat3].


Dirichlet Forms on Noncommutative Spaces 219<br />

3.1 Quantum Sp<strong>in</strong> Systems<br />

Let us describe now briefly the quantum sp<strong>in</strong> system and their dynamics (see<br />

[BR2 6.2] for a detailed exposition). Let Zd be the d ≥ 1 dimensional lattice<br />

space, whose sites are occupied by sp<strong>in</strong>s 1<br />

2 particles. The observables at site<br />

x ∈ Zd are elements of the algebra<br />

A {x} := M2(C) .<br />

The algebra of observables conf<strong>in</strong>ed <strong>in</strong> the f<strong>in</strong>ite region X ⊂ Zd ,givenby<br />

AX := �<br />

A {x} ,<br />

x∈X<br />

is then the full matrix algebra M 2 |X|(C), where |X| denotes the card<strong>in</strong>ality<br />

of a set X. Let us denote by L the class (ordered by <strong>in</strong>clusion) of all f<strong>in</strong>ite<br />

subsets of Z d .WhenL1 ,L2 ∈Lare disjo<strong>in</strong>t regions, L1 ∩ L2 = ∅, then<br />

AL1∪L2 = AL1 ⊗ AL2 .<br />

On the other hand, if L1 ⊆ L2 and 1L2 denotes the unit of AL2, the algebra<br />

AL1 is identified with the subalgebra AL1 ⊗ 1L2 of AL1∪L2. The normed<br />

algebra of all local observables is given by<br />

A0 := �<br />

X∈L<br />

AX ,<br />

while the quasi-local C∗-algebra A is the norm completion of the local<br />

algebra A0.<br />

The mutual <strong>in</strong>fluence of sp<strong>in</strong>s conf<strong>in</strong>ed to f<strong>in</strong>ite regions is represented by<br />

an <strong>in</strong>teraction, i.e. a family Φ := {ΦX : X ∈L}of self-adjo<strong>in</strong>t elements<br />

ΦX ∈ AX. The bounded derivation<br />

AX ∋ a ↦→ i � ΦY ,a �<br />

a ∈ A,<br />

generates a uniformly cont<strong>in</strong>uous group of automorphisms of A, represent<strong>in</strong>g<br />

the time evolution of the observables subject to the <strong>in</strong>fluence of the particles<br />

conf<strong>in</strong>ed to Y . In order to take <strong>in</strong>to account simultaneously the mutual <strong>in</strong>fluences<br />

among particles <strong>in</strong> different regions, one may hope that a superposition<br />

of these bounded derivations may give rise to a closable derivation (δ, D(δ))<br />

on A, provided the <strong>in</strong>teraction is sufficiently moderate. There are various<br />

conditions on the strength of an <strong>in</strong>teraction provid<strong>in</strong>g derivations generat<strong>in</strong>g<br />

cont<strong>in</strong>uous groups of automorphisms of A. The follow<strong>in</strong>g classical result,<br />

applicable to <strong>in</strong>teractions for which the many body forces are negligible <strong>in</strong><br />

an appropriate sense, is not the most general, as far as the problem of the<br />

existence of the dynamics is concerned. It implies, however, the existence of


220 F. Cipriani<br />

a f<strong>in</strong>ite group velocity for the dynamics of the systems, which will be used <strong>in</strong><br />

the sequel to construct the Dirichlet form generat<strong>in</strong>g the Markovian approach<br />

to the equilibrium. We will denote by D(X) the diameter of a set X.<br />

Theorem 3.1. ([BR2 Theorem 6.2.4, 6.2.11], [LR]) Consider, on the quasilocal<br />

C∗-algebra A, an <strong>in</strong>teractions satisfy<strong>in</strong>g, for a fixed λ>0,<br />

�<br />

�Φ�λ := sup |X|4 |X| e λD(X) �ΦX� < +∞ . (3.1)<br />

Then, sett<strong>in</strong>g<br />

x∈Zd x∈X∈L<br />

D(δ) :=A0 δ(a) := �<br />

X∩Y �=∅<br />

i � ΦY ,a �<br />

a ∈ AX , X ∈L, (3.2)<br />

a closable derivation is correctly def<strong>in</strong>ed. Its closure generates a strongly cont<strong>in</strong>uous<br />

group α Φ := {α Φ t : t ∈ R} of automorphisms of A, satisfy<strong>in</strong>g the<br />

follow<strong>in</strong>g properties<br />

i) analyticity: the time evolutions R ∋ t → α Φ t (a) of local observables a ∈ A0,<br />

admit analytic extensions to doma<strong>in</strong>s conta<strong>in</strong><strong>in</strong>g the strip Iβ0, whereβ0 :=<br />

λ<br />

2�Φ�λ ;<br />

ii) f<strong>in</strong>ite group velocity: denot<strong>in</strong>g by d(x, X) distance of the site x ∈ Z d from<br />

the region X ∈L, we have for all a ∈ A {x} ,b∈ AX ,t∈ R<br />

�[α Φ t<br />

(a),b]� ≤2�a�·�b�·|X|·e−<br />

� �<br />

λd(x,X))−2|t|�Φ�λ<br />

. (3.3)<br />

Example 3.2. (Is<strong>in</strong>g and Heisenberg models) Suppose that, <strong>in</strong> addition to<br />

an external potential, represented by a one-body <strong>in</strong>teraction, particles may<br />

<strong>in</strong>fluence each other only through a two-body <strong>in</strong>teraction. This means that<br />

Φ(X) = 0 whenever |X| ≥3, so that the norms �Φ�λ <strong>in</strong> (3.1) are f<strong>in</strong>ite if and<br />

only if<br />

sup<br />

x∈Zd �Φ({x})� < +∞ , sup<br />

x∈Zd �<br />

e λ|x−y| �Φ({x, y})� < +∞ .<br />

y∈Z d<br />

These are just uniform bounds on the energy of the particles occupy<strong>in</strong>g the<br />

sites of the lattice. In case Φ is translationally <strong>in</strong>variant, this reduces to<br />

�<br />

e λ|y| �Φ({0,y})� < +∞ .<br />

y∈Z d<br />

Us<strong>in</strong>g the Pauli’s matrices σ x j ∈ A {x}, j =0, 1, 2, 3, at a fixed site x ∈ Z d :<br />

σ x 0 =<br />

� �<br />

10<br />

01<br />

σ x 1 =<br />

� �<br />

01<br />

10<br />

σ x 2 =<br />

� �<br />

0 −i<br />

i 0<br />

σ x 3 =<br />

� �<br />

1 0<br />

,<br />

0 −1


Dirichlet Forms on Noncommutative Spaces 221<br />

one may choose, for example,<br />

with h ∈ R and<br />

Φ({0}) :=hσ 0 3 , Φ({0,x}) :=<br />

3�<br />

i=1<br />

3� �<br />

e λ|x| |Ji(x)| < +∞ .<br />

i=1 x∈Zd Ji(x)σ 0 i σ x i ,<br />

These choices are referred to as the anisotropic Heisenberg model when Ji �= 0<br />

and Ji �= Jj for some pair i, j; theisotropic Heisenberg model when J1 = J2 =<br />

J3 �= 0;theX − Y model when Ji �= 0,i =1, 2 but J3 =0;theIs<strong>in</strong>g model<br />

when J1 = J2 =0.<br />

3.2 Markovian Approach to Equilibrium<br />

Let ω be a (αΦ ,β)-KMS-state at <strong>in</strong>verse temperature β>0, correspond<strong>in</strong>g to<br />

: t ∈ R}, hence to the <strong>in</strong>teraction Φ satisfy<strong>in</strong>g the assump-<br />

the evolution {αΦ t<br />

tion of the previous theorem. Let (πω, Hω,ξω) be the GNS representations of<br />

the state ω, M the von Neumann algebra πω(A) ′′ and (M, Hω , H + ω ,Jω) the<br />

correspond<strong>in</strong>g standard form.<br />

The normal extension of ω to M is a (�α Φ ,β)-KMS-state w.r.t. the normal<br />

extension {�α Φ t : t ∈ R} of the time evolution on A. Hence the modular group<br />

{σt : t ∈ R} of the cyclic vector ξω co<strong>in</strong>cides with {�α Φ −βt : t ∈ R}. This<br />

implies, <strong>in</strong> particular, that the time evolutions of the local observables <strong>in</strong> M<br />

admit analytic extensions to doma<strong>in</strong>s conta<strong>in</strong><strong>in</strong>g the strip Iβ0/β: πω(A0) ⊂ M β0/β .<br />

To construct a Dirichlet form on the GNS Hilbert space Hω, we first apply<br />

Proposition 2.59 with f a fixed admissible function, or with f = f0.<br />

Denote by Ex,j the quadratic forms associated to the self-adjo<strong>in</strong>t elements<br />

a x j := πω(σ x j ):<br />

�<br />

Ex,j[ξ] =<br />

�(σt−i/4(a R<br />

x j ) − j(σt−i/4(a x j )))ξ�2 f(t)dt . (3.4)<br />

Notice that, by Proposition 2.59, these forms are correctly def<strong>in</strong>ed, and <strong>in</strong><br />

fact completely Dirichlet forms, provided the ratio β0/β is greater than 1/2.<br />

Theorem 3.3. Let Φ be an <strong>in</strong>teraction on the quasi-local algebra A satisfy<strong>in</strong>g<br />

the bound (3.1), for a fixed λ>0. Suppose that ω is a (α Φ ,β)-KMS-state


222 F. Cipriani<br />

correspond<strong>in</strong>g to the evolution {αΦ t : t ∈ R} at an <strong>in</strong>verse temperature β>0<br />

satisfy<strong>in</strong>g<br />

β< λ<br />

. (3.5)<br />

�Φ�λ<br />

Then, for any p-admissible function f with p>d+1 and for f = f0, the<br />

quadratic functional<br />

E : Hω → [0, +∞] E[ξ] := �<br />

x∈Zd j=0<br />

3�<br />

Ex,j[ξ] (3.6)<br />

is a completely Dirichlet form on Hω, with respect to the cyclic vector ξω<br />

represent<strong>in</strong>g ω.<br />

Proof. The condition β< λ<br />

�Φ�λ guarantees that β0/β > 1<br />

2 , so that the selfadjo<strong>in</strong>t,<br />

local elements {ax j : x ∈ Zd ,j = 0, 1, 2, 3} belong to M1/2. By<br />

Proposition 2.59, the quadratic forms Ex,j are completely Dirichlet forms. As<br />

the functional E is def<strong>in</strong>ed as the supremum of bounded, completely Dirichlet<br />

forms, it shares with them the completely Markovian property. What we have<br />

to prove is that E is densely def<strong>in</strong>ed. We will show that E is f<strong>in</strong>ite on the<br />

symmetric embedd<strong>in</strong>g <strong>in</strong> Hω of the local algebra πω(A0).<br />

By the boundedness of the symmetric embedd<strong>in</strong>gs, see Proposition 2.22, one<br />

has �iω(b)� ≤�b�, for all b ∈ M. We have for all b ∈ M ,a= a∗ ∈ M1/2 and then<br />

�(σ t−i/4(a) − j(σ t−i/4(a)))iω(b)� = �iω([σt(a),b])� ≤�[σt(a),b]� ,<br />

E[iω(b)] ≤<br />

3� �<br />

�<br />

j=0 x∈Zd R<br />

�[σt(a x j ),b]� 2 f(t) dt b ∈ M . (3.7)<br />

Assume now that b ∈ πω(A0), or, more precisely, that b ∈ πω(AX) forsome<br />

fixed X ∈L. For each fixed <strong>in</strong>dex j =0, 1, 2, 3, let us split the sum over<br />

x ∈ Zd , <strong>in</strong> the right hand side of (3.7) as follows:<br />

�<br />

�<br />

(3.8)<br />

where<br />

x∈Z d<br />

I 1 j := �<br />

I 2 j := �<br />

�[σt(a<br />

R<br />

x j ),b]�2f(t) dt = I 1 j + I2 j<br />

�<br />

N∈N d(x,X)=N<br />

�<br />

N∈N d(x,X)=N<br />

�<br />

�<br />

|t|≤ Nβ 0<br />

2<br />

|t|> Nβ 0<br />

2<br />

�[σt(a x j ),b]� 2 f(t) dt ,<br />

�[σt(a x j ),b]� 2 f(t) dt .


Dirichlet Forms on Noncommutative Spaces 223<br />

By Theorem 3.1 ii), and <strong>in</strong> particular the bound (3.3), if b = πω(c) forsome<br />

c ∈ AX, as Pauli matrices have unit norm, we have<br />

�[σt(a x j ),b]� = �[σt(πω(σ x j )),πω(c)]� = �[πω(α Φ −βt (σx j )),πω(c)]�<br />

≤�[α Φ −βt (σx j ),c]�<br />

� �<br />

λd(x,X))−2|t|�Φ�λ<br />

≤ 2 ·�c�·|X|·e −<br />

t ∈ R ,<br />

(3.9)<br />

S<strong>in</strong>ce on the d-dimensional square lattice Z d the follow<strong>in</strong>g bound holds true<br />

us<strong>in</strong>g (3.9) we have that<br />

I 1 j ≤ 4 ·�c� 2 �<br />

2<br />

·|X|<br />

≤ 4 ·�c� 2 ·|X| 2 ·<br />

♯{x ∈ Z d : d(x, X) =N} �N d−1<br />

�<br />

N∈N d(x,X)=N<br />

� �<br />

R<br />

�<br />

� �<br />

f(t)dt<br />

|t|≤ Nβ 0<br />

2<br />

�<br />

N∈N d(x,X)=N<br />

N →∞,<br />

e −2<br />

� �<br />

λN−2|t|�Φ�λ f(t)dt<br />

e −Nλ < +∞ .<br />

(3.10)<br />

S<strong>in</strong>ce �[σt(ax j ),b]� ≤2�b� for all t ∈ R and j =0, 1, 2, 3, <strong>in</strong> case f = f0, the<br />

decay properties of f0 imply that I2 j is f<strong>in</strong>ite. In case of a general p-admissible<br />

function<br />

f(t) ≤ M · (1 + |t|) −p<br />

t ∈ R<br />

for some M>0andp>d+ 1, so that we still have<br />

I 2 j ≤ 4 ·�b�2 · M · � �<br />

�<br />

(1 + |t|) −p dt<br />

≤ 4 ·�b� 2 · M · �<br />

N∈N d(x,X)=N<br />

�<br />

N∈N d(x,X)=N<br />

|t|> Nβ 0<br />

2<br />

� 2<br />

p − 1<br />

��<br />

1+ Nβ0<br />

2<br />

� 1−p<br />

< +∞ .<br />

(3.11)<br />

Corollary 3.4. ([P2 Theorem 2.1]) Under the assumptions of Theorem 3.3<br />

the follow<strong>in</strong>g properties are equivalent:<br />

i) ω is an extremal (αΦ ,β)-KMS-state;<br />

ii) ω is a factor state;<br />

iii) the Markovian semigroup {Tt : t ≥ 0} is ergodic.<br />

Proof. The equivalence between properties i) and ii) is a well known property<br />

of KMS-states and may be found, for example, <strong>in</strong> [BR2 Theorem 5.3.30].<br />

Before establish<strong>in</strong>g the equivalence between properties ii) and iii), notice first<br />

that, the subspace Kω ⊆Hω, where the Dirichlet form (3.6) is vanish<strong>in</strong>g,<br />

co<strong>in</strong>cides both with the subspace where the associated semigroup acts as the<br />

identity<br />

{ξ ∈Hω : Ttξ = ξ,t≥ 0} ,


224 F. Cipriani<br />

as well as with the eigenspace of the generator H correspond<strong>in</strong>g to the eigenvalue<br />

zero<br />

{ξ ∈Hω : Hξ =0} .<br />

Notice also that, s<strong>in</strong>ce the <strong>in</strong>volutive algebra generated by the Pauli matrices<br />

{σx j : x ∈ Zd ,j =0, 1, 2, 3}, is norm dense <strong>in</strong> the quasi-local C∗-algebra A,<br />

the <strong>in</strong>volutive algebra generated by {πω(σx j ):x ∈ Zd ,j =0, 1, 2, 3} is σweakly<br />

dense <strong>in</strong> the von Neumann algebra M. By a more general result, due<br />

Y.M. Park [P2 Theorem 2.1], the structure of the Dirichlet form (3.6) is such<br />

that its vanish<strong>in</strong>g subspace is precisely<br />

Kω = Z(M)ξω ,<br />

i.e. the closure of the symmetric image <strong>in</strong> Hω of the center Z(M) :=M∩M ′ of<br />

the algebra M. Suppose now that ω is a factor state, so that Z(M) =C · 1M.<br />

Then the eigenvector ξω is simple and, by Corollary 2.48, the semigroup is<br />

ergodic. Conversely, if the semigroup is ergodic, Corollary 2.48 implies that<br />

ξω is simple, so that the center Z(M) is trivial and ω is a factor state.<br />

4 Dirichlet Forms and Differential Calculus<br />

on C ∗ -algebras<br />

In this chapter we show that, on algebras endowed with semi-f<strong>in</strong>ite traces,<br />

Dirichlet forms are <strong>in</strong> one-to-one correspondence with differential calculi. On<br />

one hand this correspondence provides a tool to construct Markovian semigroups<br />

symmetric with respect to traces. On the other hand this allows to<br />

analyze Dirichlet spaces from a geometric po<strong>in</strong>t of view. The mean<strong>in</strong>g of this<br />

correspondence will be clarified through several examples <strong>in</strong> this chapter. The<br />

correspondence will be fundamental to prove the geometric applications of<br />

the last two chapters of these notes.<br />

4.1 Dirichlet Forms and Markovian Semigroups<br />

with Respect to Traces<br />

Our aim, <strong>in</strong> this section, is to show that any regular Dirichlet form on a<br />

C ∗ -algebra A endowed with a reasonable trace gives rise to a first order differential<br />

calculus on A. The matter of this section is taken from [CS1]. Under<br />

the stronger hypothesis that the doma<strong>in</strong> of the self-adjo<strong>in</strong>t generator associated<br />

to the Dirichlet form conta<strong>in</strong>s a dense sub-algebra, it was previously<br />

proved <strong>in</strong> [S2,3]. The use of derivation <strong>in</strong> the construction of unbounded<br />

Dirichlet forms was also considered <strong>in</strong> [DL2] and [GIS].<br />

Let A be a C ∗ -algebra and τ a densely def<strong>in</strong>ed, faithful, semi-f<strong>in</strong>ite, lower<br />

semi-cont<strong>in</strong>uous trace on it. When τ is f<strong>in</strong>ite, the theory developed <strong>in</strong> the


Dirichlet Forms on Noncommutative Spaces 225<br />

previous section applies and we have at our disposal good notions of positivity<br />

and Markovianity, for quadratic forms and semigroups, as well as a one to one<br />

correspondence between them. In the f<strong>in</strong>ite trace sett<strong>in</strong>g, the theory simplifies<br />

due to the fact that the modular operator reduces to the identity so that,<br />

essentially, positivity has the same mean<strong>in</strong>g <strong>in</strong> the algebra as <strong>in</strong> the standard<br />

Hilbert space.<br />

We now briefly <strong>in</strong>dicate the changes needed to cover the case of C∗-algebras with a densely def<strong>in</strong>ed, faithful, semi-f<strong>in</strong>ite, lower semi-cont<strong>in</strong>uous trace τ :<br />

A+ → [0, +∞].<br />

In the C ∗ -algebra A def<strong>in</strong>e the subset L := {a ∈ A : τ(a ∗ a) < ∞}. It<br />

is easy to see that L is a left ideal <strong>in</strong> A, N := L ∗ L is a subalgebra of<br />

A and τ extends to a l<strong>in</strong>ear functional on N .LetL 2 (A, τ) be the Hilbert<br />

space def<strong>in</strong>ed complet<strong>in</strong>g L with respect to the pre-Hilbert space structure<br />

given by the sesquil<strong>in</strong>ear form associat<strong>in</strong>g τ(a ∗ b)toa, b ∈L.Denot<strong>in</strong>gby<br />

ητ : L → L 2 (A, τ) the natural <strong>in</strong>jection with dense range, the Gelfand-<br />

Naimark-Segal (GNS) representation is def<strong>in</strong>ed by<br />

πτ : A →B(L 2 (A, τ)) (ητ (b)|πτ (a)ητ (c)) := τ(b ∗ ac)<br />

for a ∈ A and b, c ∈L.WedenotebyM or L∞ (A, τ) the von Neumann<br />

algebra πτ (A) ′′ generated by A <strong>in</strong> this representation and by the same symbol<br />

τ the normal extension to M of the trace on A. Moreover, we shall denote by<br />

the same symbols L, N we used before for the trace τ on A, the corrispond<strong>in</strong>g<br />

spaces associated to its normal extension on M.<br />

From now on we shall not dist<strong>in</strong>guish between a f<strong>in</strong>ite trace element of M<br />

and its image <strong>in</strong> the L2 (A, τ).<br />

Sometimes, we will denote by L∞ + (A, τ) the positive cone M+ .<br />

In the trace case the <strong>in</strong>volution a ↦→ a∗ of M extends to an anti-unitary map<br />

on the GNS space L2 (A, τ). A self-dual closed, convex cone <strong>in</strong> the Hilbert<br />

space L2 (A, τ) is then realized as the closure of the subset<br />

L 2 + (A, τ) :=L+<br />

and (M,L2 (A, τ)),L2 + (A, τ),J) is a standard form of M.<br />

In particular, the positive part (resp. the modulus) of a self-adjo<strong>in</strong>t element<br />

a ∈ M ∩ L2 (A, τ), w.r.t. the cone L2 +(A, τ), co<strong>in</strong>cides with the positive part<br />

a+ (resp. the modulus |a| = a+ + a−) ofa<strong>in</strong> the algebra M.<br />

In the framework of this standard form we can apply directly the part of<br />

the theory developed <strong>in</strong> section 2.6 concern<strong>in</strong>g positive semigroups and their<br />

characterization <strong>in</strong> terms of quadratic forms. In particular Theorem 2.53 rema<strong>in</strong>s<br />

valid <strong>in</strong> this context.<br />

To extend the notion of Markovianity for semigroup and forms to the present<br />

semi-f<strong>in</strong>ite sett<strong>in</strong>g, we have only to face the fact that sets like<br />

{ξ ∈ L 2 (A, τ) :ξ ≤ ξτ } , {ξ ∈ L 2 +(A, τ) :ξ ≤ ξτ }


226 F. Cipriani<br />

(on which we based the def<strong>in</strong>itions of Markovian forms and Markovian semigroups<br />

w.r.t. states) have to be modified due to the possible lack of a cyclic<br />

vector ξτ ∈ L 2 (A, τ) represent<strong>in</strong>g the semi-f<strong>in</strong>ite trace τ. Def<strong>in</strong><strong>in</strong>g the closed<br />

convex set C ⊂ L 2 (A, τ)<br />

C := {a ∈L: a = a ∗ ≤ 1M} , (4.1)<br />

we shall denote by a∧1 the Hilbertian projection of a onto the closed, convex<br />

set C. Incasea = a ∗ ∈ A ∩ L 2 (A, τ) the projection a ∧ 1 is an element of the<br />

algebra A and co<strong>in</strong>cides with the one obta<strong>in</strong>ed, by the cont<strong>in</strong>uous functional<br />

calculus of A, compos<strong>in</strong>g the cont<strong>in</strong>uous function R ∋ t ↦→ f(t) =t∧1 witha.<br />

Def<strong>in</strong>ition 4.1. (Dirichlet form with respect to a faithful, semif<strong>in</strong>ite, lowersemicont<strong>in</strong>uous<br />

trace)<br />

A quadratic functional E : L2 (A, τ) → (−∞, +∞] issaidtobe<br />

• i) J-real if E[Jξ]=E[ξ] for all ξ ∈ L2 (A, τ);<br />

• ii) Markovian if<br />

E[ξ ∧ 1] ≤E[ξ] ξ = Jξ ∈ L 2 (A, τ) ,<br />

where ξ ∧ 1 denotes the projection of ξ onto the closed, convex set C;<br />

• iii) Dirichlet form if it is a lower semi-cont<strong>in</strong>uous, Markovian functional;<br />

• iv) completely Dirichlet form if all of its canonical extensions En to<br />

L 2 (Mn(A),τn)<br />

En[[aij] n i,j=1] :=<br />

n�<br />

i,j=1<br />

E[aij] [aij] n i,j=1 ∈ L 2 (Mn(A),τn)<br />

are Dirichlet form for all n ≥ 1; here τn denotes the trace τn := τ ⊗ trn on<br />

the C ∗ -algebra Mn(A) :=A ⊗ Mn(C);<br />

• v) regular if the subspace B := A ∩F is norm dense <strong>in</strong> the C ∗ -algebra A<br />

and a form core for (E, F); here the subspace<br />

F := {ξ ∈L 2 (A, τ) :E[ξ] < +∞}<br />

denotes the doma<strong>in</strong> of the quadratic form or functional;<br />

• vi) a C ∗ -Dirichlet form if it is a regular, completely Dirichlet form;<br />

• vii) a J-real, symmetric, strongly cont<strong>in</strong>uous, semigroup {Tt : t ≥ 0} on<br />

L 2 (A, τ) will be called Markovian if it leaves globally <strong>in</strong>variant the subset<br />

C ∩ L 2 + (A, τ) ={a ∈L:0≤ a ≤ 1M};<br />

• viii) the semigroup is said to be completely positive (resp. Markovian)<br />

if all of its canonical extensions to L 2 (Mn(A),τn) are positive (resp.<br />

Markovian).


Dirichlet Forms on Noncommutative Spaces 227<br />

As announced before, the one to one correspondence between completely<br />

Dirichlet forms, completely Markovian semigroups on L 2 (A, τ) and completely<br />

Markovian semigroups on the von Neumann algebra M still holds true<br />

start<strong>in</strong>g from a densely def<strong>in</strong>ed, faithful, semi-f<strong>in</strong>ite, lower semi-cont<strong>in</strong>uous<br />

trace τ on a C ∗ -algebra A or at least from a densely def<strong>in</strong>ed, faithful, semif<strong>in</strong>ite,<br />

normal trace τ on a von Neumann algebra M.<br />

4.2 Modules and Derivations on C ∗ -algebras<br />

and Associated Dirichlet Forms<br />

In this section we show how the construction of the model Dirichlet form, i.e.<br />

the Dirichlet <strong>in</strong>tegral<br />

�<br />

E[u] = |∇u(x)| 2 m(dx) ,<br />

can be generalized <strong>in</strong> terms of derivations on C ∗ -algebras.<br />

R d<br />

Def<strong>in</strong>ition 4.2. (Modules and Derivations on C ∗ -algebras) An A–bimodule<br />

over the C ∗ -algebra A is a Hilbert space H with a left representation and a<br />

right representation of A, which commute. Equivalently, it can be thought as<br />

arepresentationoftheC ∗ –algebra A ⊗max A ◦ , maximal tensor product of A<br />

and its opposite C ∗ -algebra A ◦ .<br />

The left action of a ∈ A on ξ ∈Hwill be denoted by aξ ∈H, while the right<br />

action by ξa ∈H. S<strong>in</strong>ce the actions commute, we may denote (aξ)b = a(ξb)<br />

by a unique symbol aξb for a, b ∈ A and ξ ∈H.<br />

The bimodule H is said to be symmetric if there exists an isometric, antil<strong>in</strong>ear<br />

<strong>in</strong>volution J : H→Hexchang<strong>in</strong>g the right and left actions of A<br />

J (aξb) =b ∗ J (ξ)a ∗<br />

∀ a, b ∈ A, ξ ∈H.<br />

A derivation ∂ : D(∂) → H is def<strong>in</strong>ed to be a l<strong>in</strong>ear map def<strong>in</strong>ed on a<br />

subalgebra D(∂) ofA, satisfy<strong>in</strong>g the follow<strong>in</strong>g Leibniz rule:<br />

∂(ab) =∂(a)b + a∂(b) a, b ∈ D(∂).<br />

The derivation is called symmetric if its doma<strong>in</strong> D(∂) is a self-adjo<strong>in</strong>t subalgebra,<br />

the A–bimodule (H , J ) is symmetric and<br />

∂(a ∗ )=J (∂a) a ∈ D(∂) .<br />

Example 4.3. (Gradient and square <strong>in</strong>tegrable vector fields) The standard<br />

example of a symmetric derivation is the gradient operator<br />

∇ : C ∞ c (M) → L2 (TM)


228 F. Cipriani<br />

of a Riemannian manifold M, def<strong>in</strong>ed on the subalgebra of smooth functions<br />

and tak<strong>in</strong>g values <strong>in</strong> the Hilbert space L 2 (TM) of square <strong>in</strong>tegrable vector<br />

fields. Bounded cont<strong>in</strong>uous functions on M, act<strong>in</strong>g by po<strong>in</strong>twise multiplication<br />

on vector fields, give rise to a module with respect to which the Leibniz<br />

rule holds true.<br />

Example 4.4. (Derivations associated to mart<strong>in</strong>gales and stochastic calculus)<br />

An important example of derivation appears naturally <strong>in</strong> the theory of symmetric<br />

Hunt processes {Xt : t ≥ 0} on locally compact spaces X (i.e. the<br />

probabilistic counterpart of the theory of classical Dirichlet forms) and <strong>in</strong><br />

particular, <strong>in</strong> the decomposition theory of additive functionals (see [FOT]).<br />

There, the algebra is the space C0(X) of cont<strong>in</strong>uous functions f on X, the<br />

bimodule H is the space M of square <strong>in</strong>tegrable mart<strong>in</strong>gales M of f<strong>in</strong>ite energy,<br />

and the left and right actions are given by a suitable def<strong>in</strong>ed stochastic<br />

<strong>in</strong>tegral f • M. The derivation ∂f is given by the f<strong>in</strong>ite energy mart<strong>in</strong>gale<br />

part of M [f] of the cadlag additive functional<br />

A [f]<br />

t = f(Xt) − f(X0) t ≥ 0 .<br />

We are go<strong>in</strong>g to see that derivations naturally gives rise to Dirichlet forms.<br />

Theorem 4.5. (Construction of Dirichlet forms by derivations) Let (∂,D(∂))<br />

be a densely def<strong>in</strong>ed, symmetric derivation on the C ∗ -algebra A, tak<strong>in</strong>g values<br />

<strong>in</strong> the symmetric Hilbert A–bimodule (H, J ). Suppose that D(∂) is conta<strong>in</strong>ed<br />

and dense <strong>in</strong> L 2 (A, τ) and that (∂,D(∂)) is a closable operator from L 2 (A, τ)<br />

to H.<br />

Then the closure of the quadratic form (E, F) given by<br />

E : L 2 (A, τ) → [0, ∞) E[a] =�∂a� 2 H , a ∈F:= D(∂) (4.2)<br />

is a C ∗ -Dirichlet form. The generator ∆ of the completely Markovian semigroup<br />

Tt = e −t∆ associated to E, can be represented as the composition of the<br />

“divergence” ∂ ∗ of the “gradient” ∂:<br />

∆ = ∂ ∗ ◦ ∂.<br />

In other words, the derivation ∂ appears as a differential square root of ∆.<br />

To establish the Markovianity of E we are go<strong>in</strong>g to prove a cha<strong>in</strong> rule formula<br />

for derivations which generalizes the familiar one we have seen <strong>in</strong> Example 4.3.<br />

For any self-adjo<strong>in</strong>t element a of A, we shall denote by La (resp. Ra)<br />

the unique representation of C(sp(a)), the C∗-algebra of cont<strong>in</strong>uous, complex<br />

valued functions on sp(a) (the spectrum of a), such that<br />

La(f)ξ =<br />

�<br />

f(a)ξ if f(0) = 0<br />

ξ if f ≡ 1<br />

f ∈ C(sp(a)) ξ ∈H


Dirichlet Forms on Noncommutative Spaces 229<br />

and<br />

�<br />

Ra(f)ξ =<br />

ξf(a) iff(0) = 0<br />

ξ if f ≡ 1<br />

f ∈ C(sp(a)) ξ ∈H.<br />

La ⊗ Ra will be the tensor product representation of C(sp(a)) ⊗ C(sp(a)) =<br />

C(sp(a) × sp(a)). When I ⊆ R is a closed <strong>in</strong>terval and f ∈ C1 (I), we will<br />

denote by ˜ f ∈ C(I × I) thedifference quotient on I × I, sometimes called the<br />

quantum derivative of f, def<strong>in</strong>ed by<br />

˜f(s, t) =<br />

� f(s)−f(t)<br />

s−t<br />

if s �= t<br />

f ′ (s) if s = t.<br />

(4.3)<br />

Next lemma shows that the doma<strong>in</strong> of a closed derivation with values <strong>in</strong><br />

a Hilbertian bimodule is closed under C 1 -functional calculus. The case of<br />

the standard bimodule L 2 (A, τ) was considered <strong>in</strong> [GIS], while the case of a<br />

C ∗ -bimodule was analyzed <strong>in</strong> [S3].<br />

Lemma 4.6. (Functional calculus <strong>in</strong> the doma<strong>in</strong> of a derivation) Let (H , J )<br />

be a symmetric Hilbert A–bimodule, (∂,D(∂)) be a symmetric derivation def<strong>in</strong>ed<br />

on an <strong>in</strong>volutive subalgebra of A <strong>in</strong>to H, andleta = a∗ ∈ A. Then<br />

i) for any polynomial f on R, f(a) belongs to D(∂), the follow<strong>in</strong>g cha<strong>in</strong>-rule<br />

holds true<br />

∂(f(a)) = (La ⊗ Ra)( ˜ f) ∂(a) , (4.4)<br />

and we have the bound<br />

�∂(f(a))� ≤�f ′ � C(sp(a))�∂(a)� a ∈ D(∂); (4.5)<br />

ii) if (∂,D(∂)) is closable as operator from A to H, then its closure is a<br />

derivation;<br />

iii) if (∂,D(∂)) is a closed derivation from A to H, the above cha<strong>in</strong>-rule holds<br />

true for all f ∈ C 1 (sp(a)) such that f(0) = 0. In particular the doma<strong>in</strong> of a<br />

closed derivation is closed under C 1 -functional calculus.<br />

Proof. i) When f is a polynomial it is easy to verify (4.4) by direct <strong>in</strong>spection<br />

and then derive (4.5). Statement ii) follows directly from the Leibniz rule and<br />

the cont<strong>in</strong>uity of the left and right actions. To prove the statement <strong>in</strong> iii),<br />

we may approximate the derivative f ′ of a function f ∈ C1 (sp(a)) vanish<strong>in</strong>g<br />

at the orig<strong>in</strong>, by polynomials f ′ n , uniformly on sp(a), and then approximate<br />

uniformly f by the primitive polynomials fn vanish<strong>in</strong>g at the orig<strong>in</strong>. Apply<strong>in</strong>g<br />

(4.5) we have that ∂fn(a) is a Cauchy-sequence <strong>in</strong> H so that, s<strong>in</strong>ce the<br />

derivation has been assumed to be closed, ∂f(a) belongs to that doma<strong>in</strong> and<br />

(4.4) holds true for all f <strong>in</strong> C 1 (sp(a)).<br />

Proof. of Theorem 4.5. Let us denote by (E, F) the closure <strong>in</strong> L 2 (A, τ) ofthe<br />

form (E, F) def<strong>in</strong>ed <strong>in</strong> the statement of the theorem. First notice that the


230 F. Cipriani<br />

form is J-real due to the fact the derivation is symmetric. Follow<strong>in</strong>g the same<br />

reason<strong>in</strong>g as the one used <strong>in</strong> the proof of Lemma 4.6 iii), we start by prov<strong>in</strong>g<br />

the bound<br />

E[f(a)] ≤�f ′ � 2 C(sp(a)) E[a] a ∈F (4.6)<br />

for all f ∈ C1 (sp(a)) such that f(0) = 0. Let us fix an element a ∈F and<br />

denote I the spectrum sp(a). Consider the same polynomial approximation<br />

fn (vanish<strong>in</strong>g at the orig<strong>in</strong>) of the function f. S<strong>in</strong>cefnand f ′ n are uniformly<br />

convergent to f and f ′ , respectively, they are uniformly bounded sequence <strong>in</strong><br />

C(I). An application of the bound (4.5) and the estimate<br />

�fn(a)� 2 2 = τ(|fn(a)| 2 ) ≤�fn� 2 C(I) τ(|a|2 )=�fn� 2 C(I)) �a�2 2 ,<br />

tell us that (fn(a),∂fn(a)) is bounded sequence with respect to the graph<br />

norm of the operator (∂,D(∂)). Hence, possibly consider<strong>in</strong>g a suitable subsequence,<br />

it is weakly convergent to some (b, ξ) ∈ F×H.S<strong>in</strong>cefn(a) isalso<br />

converg<strong>in</strong>g <strong>in</strong> the norm of A to f(a), we have the identification b = f(a) and<br />

E[f(a)] = lim<br />

n→∞ E[fn(a)] ≤ lim<br />

n→∞ �f ′ n�2C(I) E[a] =�f ′ � 2 C(I)) E[a] .<br />

In particular f(a) ∈ F for all a ∈F= D(∂) andf ∈ C 1 (sp(a)) such that<br />

f(0) = 0. We will denote Lip 0(R) the algebra of all real Lipschitz functions<br />

f on the real l<strong>in</strong>e, vanish<strong>in</strong>g at the orig<strong>in</strong>, normed by the <strong>in</strong>fimum �f� Lip0 (R)<br />

of all numbers λ such that |f(s) − f(t)| ≤λ|s − t|, s, t ∈ R. Approximat<strong>in</strong>g<br />

locally uniformly the elements of Lip 0(R) by functions <strong>in</strong> C(R) vanish<strong>in</strong>g at<br />

the orig<strong>in</strong>, we obta<strong>in</strong> from (4.6) the estimate<br />

E[f(a)] ≤�f� 2 Lip0 (sp(a)) E[a] a ∈F f ∈ Lip0 (sp(a)) . (4.7)<br />

S<strong>in</strong>ce F is a form core for (E, F), the estimate (3.7) holds true for all a ∈ F.<br />

We get the Markovianity of (E, F) just choos<strong>in</strong>g as Lipschitz function f(t) =<br />

t∧1. Complete Markovianity follows just apply<strong>in</strong>g the result to the extension<br />

of the derivation to matrix algebras over A.<br />

4.3 Modules and Derivations Associated to Regular<br />

Dirichlet Forms<br />

In this section we are go<strong>in</strong>g to show how Dirichlet spaces, commutative or<br />

not, give rise to a differentiable structure on a C ∗ -algebra. More precisely,<br />

we construct an essentially unique, closable derivation on a C ∗ -algebra, represent<strong>in</strong>g<br />

the Dirichlet form as <strong>in</strong> Theorem 4.5.<br />

Keep<strong>in</strong>g <strong>in</strong> m<strong>in</strong>d the model case of the Dirichlet <strong>in</strong>tegral on Euclidean spaces<br />

or Riemannian manifolds, what we are go<strong>in</strong>g to do is to reconstruct the<br />

gradient operator and the Hilbert space of square summable vector fields<br />

from the energy functional.


Dirichlet Forms on Noncommutative Spaces 231<br />

The fundamental h<strong>in</strong>t for the search of a differential calculus associated<br />

to a Dirichlet form is the follow<strong>in</strong>g observation made <strong>in</strong> [DL1]. The orig<strong>in</strong>al<br />

proof, based on the Kadison-Schwartz <strong>in</strong>equality, works for quadratic forms<br />

associated to 1 1<br />

2-Markovian semigroups. It is more general than the one we offer<br />

here, which is directly based on the fundamental properties of the Dirichlet<br />

form, but requires 2-Markovianity.<br />

Proposition 4.7. Let (E,D(E)) be a completely Dirichlet form on L 2 (A, τ).<br />

Then B := A ∩ D(E) is an <strong>in</strong>volutive subalgebra of M, called the Dirichlet<br />

algebra of (E,D(E)). Incase(E,D(E)) is a C ∗ –Dirichlet form, B is a dense,<br />

<strong>in</strong>volutive sub–algebra of A and a form core for (E ,D(E)).<br />

Proof. Let a = a ∗ ∈Bbe such that �a� =1sothat<br />

a2 = a −<br />

2 0<br />

dt a ∧ t.<br />

By convexity, lower semicont<strong>in</strong>uity and Markovianity of E, wehave<br />

�<br />

E<br />

� 1 � � 1<br />

dt a ∧ t ≤ dt E[a ∧ t] ≤E[a]<br />

� 1<br />

0<br />

0<br />

so that a2 ∈B. By scal<strong>in</strong>g the conclusion rema<strong>in</strong>s true for all self-adjo<strong>in</strong>t<br />

a ∈B. This implies that, if b = b∗ and c = c∗ are self-adjo<strong>in</strong>t elements <strong>in</strong> B,<br />

bc + cb =(b + c) 2 − b 2 − c 2 ∈B.<br />

In turn, s<strong>in</strong>ce b + c and b − c are self-adjo<strong>in</strong>t, we have<br />

so that<br />

b 2 − c 2 =<br />

(b + c)(b − c)+(b − c)(b + c)<br />

2<br />

(b + ic) 2 =(b 2 − c 2 )+i(bc + cb) ∈B.<br />

Decompos<strong>in</strong>g a generic element a ∈Bas<br />

a =<br />

a + a∗<br />

2<br />

+ ia − a∗<br />

2i<br />

,<br />

∈B,<br />

we conclude that a2 ∈B.Ifa, b are generic elements <strong>in</strong> B, consider<strong>in</strong>g the<br />

matrix � �<br />

0 a<br />

∈ M2(B) ,<br />

b 0<br />

and apply<strong>in</strong>g the above result to the extension E2 of E on M2(A), we obta<strong>in</strong><br />

� � � �2 ab 0 0 a<br />

= ∈ M2(B)<br />

0 ba b 0<br />

so that ab ∈B.


232 F. Cipriani<br />

Next step will be the construction of a Hilbert space on which the algebra<br />

A will act, naturally associated to the Dirichlet form.<br />

Proposition 4.8. The sesquil<strong>in</strong>ear form on the algebraic tensor product B⊗<br />

B, l<strong>in</strong>ear <strong>in</strong> the right-hand side and conjugate l<strong>in</strong>ear <strong>in</strong> the left-hand one,<br />

which, to c ⊗ d and a ⊗ b, associates<br />

1�<br />

� ∗ ∗ ∗ ∗<br />

E(c, abd )+E(cdb ,a) −E(db ,c a) , (4.8)<br />

2<br />

is positive def<strong>in</strong>ite and def<strong>in</strong>es a pre-Hilbert structure on B⊗B.<br />

Proof. By the complete Markovianity of the semigroup Tt = e−t∆ associated<br />

to E the resolvent maps<br />

(I + ε∆) −1 =<br />

I<br />

I + ε∆ =<br />

� ∞<br />

0<br />

e −t Tεt dt ε > 0,<br />

are bounded, completely positive, self-adjo<strong>in</strong>t contractions on L 2 (A, τ) and<br />

bounded, completely positive, normal contractions on M; by the St<strong>in</strong>espr<strong>in</strong>g<br />

construction ([Sti]), there exists, for a fixed ε>0, a Hilbert space Kε, a<br />

normal representation πε of M <strong>in</strong>to B(Kε) and a l<strong>in</strong>ear operator Wε from<br />

L 2 (A, τ) <strong>in</strong>toKε, with norm less than one, such that<br />

I<br />

I + ε∆ (a) =W ∗ ε πε(a)Wε<br />

for any a <strong>in</strong> M. It is then easy to check, for a, b, c, d ∈ M, theidentity<br />

d ∗ ∆<br />

I + ε∆ (c)∗ab+d ∗ c ∗ ∆<br />

∆<br />

(a)b − d∗<br />

I + ε∆ I + ε∆ (c∗a)b =<br />

1<br />

ε d∗� �∗� � 1<br />

Wεc − πε(c)Wε Wεa − πε(a)Wε b +<br />

ε d∗c ∗ (I − W ∗ ε Wε)ab .<br />

This implies the positivity of the matrix<br />

�<br />

∆<br />

I + ε∆ (aj) ∗ ai + a ∗ ∆<br />

j<br />

I + ε∆ (ai) − ∆<br />

I + ε∆ (a∗j ai)<br />

�<br />

i,j=1...n<br />

(4.9)<br />

(4.10)<br />

<strong>in</strong> Mn(A), for any n <strong>in</strong> N ∗ , a1,...,an <strong>in</strong> B and, s<strong>in</strong>ce the expression <strong>in</strong> (4.8)<br />

is the limit of the one <strong>in</strong> (4.9), the positivity of the sesquil<strong>in</strong>ear form.<br />

Def<strong>in</strong>ition 4.9. (Hilbert space structure) The symbol H0 will denote the<br />

Hilbert space obta<strong>in</strong>ed from B⊗B after separation and completion with<br />

respect to the sem<strong>in</strong>orm provided by the positive sesquil<strong>in</strong>ear form of<br />

Proposition 4.9. The scalar product <strong>in</strong> H0 will be denoted by (·|·)H0 (or (·|·)<br />

if no confusion can arise). We will denote by a ⊗E b the canonical image <strong>in</strong><br />

H0 of the elementary tensor a ⊗ b of B⊗B. The Hilbert space structure is<br />

then uniquely identified by the follow<strong>in</strong>g expression


Dirichlet Forms on Noncommutative Spaces 233<br />

�<br />

c ⊗E d | a ⊗E b �<br />

=<br />

H0<br />

1 � �<br />

∗ ∗ ∗ ∗<br />

E(c, abd )+E(cdb ,a) −E(db ,c a) (4.11)<br />

2<br />

on any pair of elementary vectors a ⊗E b, c ⊗E d ∈H0. Inparticular,wehave<br />

||a ⊗E b|| 2 1�<br />

� ∗ ∗ ∗ ∗<br />

H0 = E(a, abb )+E(abb ,a) −E(bb ,a a) a, b ∈B. (4.12)<br />

2<br />

The follow<strong>in</strong>g result guarantees that there exists on H0, a natural structure<br />

of right module over A.<br />

Proposition 4.10. (Right A-module structure) There exists a structure of<br />

right A-module on H0, i.e. a representation of the opposite C ∗ –algebra of A<br />

<strong>in</strong>to B(H0), characterized by<br />

� n �<br />

ai ⊗E bi<br />

i=1<br />

�<br />

b =<br />

n�<br />

i=1<br />

ai ⊗E bib , ∀n ∈ N ∗ ,a1,...,an,b1,...,bn ∈B,b∈B.<br />

(4.13)<br />

Proof. The positivity of the matrix <strong>in</strong> (4.10) provides the positivity of the<br />

operator<br />

n�<br />

b<br />

i,j=1<br />

∗ j<br />

� ∆<br />

I + ε∆ (aj) ∗ ai + a ∗ j<br />

from which we get the <strong>in</strong>equality<br />

n�<br />

�<br />

ai ⊗E bib�<br />

i=1<br />

2 H0<br />

= 1<br />

2 lim<br />

ε↓0 τ<br />

�<br />

b ∗� �n<br />

≤ 1<br />

2 �b�2 � �n<br />

lim τ<br />

ε↓0<br />

= �b� 2 �<br />

n�<br />

i=1<br />

b<br />

i,j=1<br />

∗ j ( ∆<br />

I + ε∆ (aj) ∗ ai + a ∗ j<br />

i,j=1<br />

b ∗ j<br />

ai ⊗E bi� 2 H0 .<br />

� ∆<br />

I + ε∆ (aj) ∗ ai + a ∗ j<br />

∆<br />

I + ε∆ (ai) − ∆<br />

I + ε∆ (a∗j ai)<br />

�<br />

∆<br />

I + ε∆ (ai) − ∆<br />

I + ε∆ (a∗ � �<br />

j ai))bi b<br />

∆<br />

I + ε∆ (ai) − ∆<br />

I + ε∆ (a∗ � �<br />

j ai) bi<br />

By cont<strong>in</strong>uity we can then extend the right multiplication by any b ∈ A to<br />

the whole H0. One easily checks, on the l<strong>in</strong>ear span of the elementary vectors,<br />

that this operation <strong>in</strong>tertw<strong>in</strong>es the algebraic structures of A and B(H0) and<br />

their <strong>in</strong>volutions:<br />

(c ⊗E d| a ⊗E be)H0 =(c ⊗E de ∗ | a ⊗E b)H0 a, b, c, d, e ∈B.<br />

A natural left A–module structure of H0, i.e. a representation of A <strong>in</strong>to<br />

B(H0), is constructed with the help of the follow<strong>in</strong>g lemma.<br />

bi


234 F. Cipriani<br />

Lemma 4.11. For any n <strong>in</strong> N∗ ,anya, a1,...,an,b1,...,bn <strong>in</strong> B, one has<br />

n�<br />

n�<br />

n�<br />

� aai ⊗E bi − a ⊗E aibi�H0 ≤�a�·�<br />

i=1<br />

i=1<br />

i=1<br />

ai ⊗E bi�H0 . (4.14)<br />

Proof. For a fixed ε>0let(πε,Kε) andWεbe the representation of M and<br />

the operator from L2 (A, τ) toKε, respectively, considered <strong>in</strong> the proof of<br />

I<br />

Proposition 4.8. Recall <strong>in</strong> particular that I+ε∆ (a) =W ∗ ε πε(a)Wε for any a<br />

<strong>in</strong> M. Us<strong>in</strong>g these operators one easily checks that<br />

n�<br />

i,j=1<br />

b ∗ j<br />

−<br />

� ∆<br />

I + ε∆ (aaj) ∗ aai +(aaj) ∗<br />

n�<br />

b<br />

i,j=1<br />

∗ j a ∗ j<br />

=<br />

n�<br />

∆<br />

I + ε∆ (aai) − ∆<br />

I + ε∆ (a∗j a ∗ �<br />

aai)<br />

�<br />

∆<br />

I + ε∆ (a)∗a + a ∗ ∆ ∆<br />

(a) −<br />

I + ε∆ I + ε∆ (a∗ �<br />

a) aibi<br />

b<br />

i,j=1<br />

∗ j (Wεaj − πε(aj)Wε) ∗ a ∗ a(Wεai − πε(ai)Wε)bi<br />

≤�a� 2<br />

n�<br />

i,j=1<br />

which implies<br />

n�<br />

�<br />

� aai ⊗E bi − a ⊗E aibi� 2<br />

i=1<br />

1<br />

= lim<br />

ε↓0 2 τ<br />

� n �<br />

i,j=1<br />

−<br />

b ∗ ∆<br />

j (<br />

n�<br />

i,j=1<br />

≤�a� 2 · lim <strong>in</strong>f<br />

ε↓0<br />

≤�a� 2 ·�<br />

n�<br />

i=1<br />

b ∗ j (Wεaj − πε(aj)Wε) ∗ (Wεai − πε(ai)Wε)bi ,<br />

i<br />

I + ε∆ (aaj) ∗ ∗ ∆<br />

aai +(aaj)<br />

I +ε∆ (aai) − ∆<br />

I + ε∆ (a∗j a∗aai))bi I + ε∆ (a)∗a + a ∗ ∆ ∆<br />

(a) −<br />

I + ε∆ I + ε∆ (a∗ �<br />

a))aibi<br />

b ∗ j a ∗ j ( ∆<br />

1<br />

2ε τ<br />

� n �<br />

i,j=1<br />

ai ⊗E bi� 2<br />

b ∗ j (Wεaj − πε(aj)Wε) ∗ (Wεai − πε(ai)Wε)bi<br />

(the last <strong>in</strong>equality follow<strong>in</strong>g from identity (4.9)).<br />

Proposition 4.12. (Left A-module structure) There exists a structure of left<br />

A-module on H0, i.e. a representation of the C ∗ –algebra A <strong>in</strong>to B(H0), characterized<br />

by<br />

a(b ⊗E c) =ab ⊗E c − a ⊗E bc, a,b,c∈B. (4.15)<br />

�<br />

bi


Dirichlet Forms on Noncommutative Spaces 235<br />

Proof. By the previous lemma, any a ∈Bdef<strong>in</strong>es a bounded operator <strong>in</strong><br />

B(H0) whose action a(b ⊗E c) onb ⊗E c is given by ab ⊗E c − a ⊗E bc, sothat<br />

(4.15) holds true. Extend<strong>in</strong>g, by cont<strong>in</strong>uity, these maps to the whole A, itis<br />

easy to check that this yields a morphism of algebras from A <strong>in</strong>to B(H0). To<br />

show that it respects <strong>in</strong>volutions, one has to check that<br />

for all a, b, c, d, e <strong>in</strong> B.<br />

(e(a ⊗E b) |c ⊗E d)H0 =(a ⊗E b |e ∗ (c ⊗E d))H0 (4.16)<br />

We are now <strong>in</strong> the position to state the ma<strong>in</strong> result of this section. Recall<br />

that an A–bimodule is a Hilbert space H0 with a left representation and a<br />

right representation of A which commute (or, equivalently, a representation<br />

of the C ∗ –algebra A ⊗max A ◦ ).<br />

Theorem 4.13. (A-bimodule structure) Let (E,D(E)) be a C ∗ –Dirichlet form<br />

on L 2 (A, τ). There exists a canonical A–bimodule structure on the Hilbert<br />

space H0, characterized by<br />

a(b ⊗E c) =ab ⊗E c − a ⊗E bc (4.17)<br />

(b ⊗E c)a = b ⊗E ca (4.18)<br />

for all a, b and c belong<strong>in</strong>g to the Dirichlet algebra B.<br />

Proof. We have only to prove that the right and left actions constructed<br />

before commute. By cont<strong>in</strong>uity, it is enough to show that they commute on<br />

the elementary tensors. In fact we have:<br />

for all a1,a2,b,c∈B.<br />

(a1(b ⊗E c))a2 =(a1b ⊗E c − a1 ⊗E bc)a2<br />

= a1b ⊗E ca2 − a1 ⊗E bca2<br />

= a1(b ⊗E ca2)<br />

= a1((b ⊗E c)a2)<br />

Once the canonical A–bimodule H0 associated to the C ∗ –Dirichlet form<br />

(E,D(E)) is available, we lead our efforts to the construction of the derivation<br />

on the Dirichlet algebra B.<br />

Lemma 4.14. For a ∈B,themapla, def<strong>in</strong>ed on the l<strong>in</strong>ear span of the set<br />

of elementary tensor products of elements of B by<br />

la(b ⊗E c) := 1<br />

�<br />

E(a, bc)+E(b<br />

2<br />

∗ ,ca ∗ ) −E(b ∗ �<br />

a, c) b, c ∈B, (4.19)<br />

extends to a cont<strong>in</strong>uous l<strong>in</strong>ear form on H0, with norm not greater than E[a] 1<br />

2 .


236 F. Cipriani<br />

Proof. Fix n <strong>in</strong> N ∗ , a, b1,...,bn,c1,...,cn <strong>in</strong> B and consider the represen-<br />

tation I<br />

I+ε∆ (a) =W ∗ ε πε(a)Wε used <strong>in</strong> the proof of Proposition 4.8. We then<br />

have<br />

la<br />

� n �<br />

bi ⊗E ci<br />

i=1<br />

�<br />

1<br />

= lim<br />

ε↓0 2ε τ<br />

�<br />

(Wεa − πe(a)Wε) ∗� �n<br />

+ a ∗ (I − W ∗ ε Wε)<br />

Let us apply the Cauchy-Schwarz <strong>in</strong>equality<br />

i=1<br />

� n �<br />

i=1<br />

(Wεbi − πε(bi)Wε)ci<br />

bici<br />

��<br />

.<br />

|τ(α ∗ εβε + γ ∗ ε δε)| ≤τ(α ∗ εαε + γ ∗ ε γε) 1/2 τ(β ∗ ε βε + δ ∗ ε δε) 1/2<br />

to αε = Wεa−πε(a)Wε , βε = �<br />

and δε =(I − W ∗ ε<br />

1<br />

�<br />

+<br />

i (Wεbi −πε(bi)Wε)ci , γε =(I −W ∗ ε Wε) 1/2 a<br />

Wε) 1/2 �<br />

i bici. We have, as <strong>in</strong> the proof of Proposition 4.8<br />

limε↓0<br />

2ε τ(α∗ε αε + γ ∗ ε γε) =<br />

limε↓0<br />

1<br />

2 τ<br />

�<br />

∆<br />

I + ε∆ (a∗ )a + a ∗ ∆ ∆<br />

(a) −<br />

I + ε∆ I + ε∆ (a∗ �<br />

a) ≤E(a, a)<br />

and<br />

1<br />

lim<br />

ε↓0 2ε τ(β∗ ε βε + δ ∗ εδε) n�<br />

=|| bi ⊗E ci||H0 .<br />

i=1<br />

� �<br />

� �n<br />

Hence we proved that �la i=1 bi<br />

��<br />

�� � �n ⊗E ci ≤ E[a] ·�<br />

which, by cont<strong>in</strong>uity, the statement follows.<br />

i=1 bi ⊗E ci�H0, from<br />

Notice that, s<strong>in</strong>ce B is a form core, la can be uniquely def<strong>in</strong>ed for any<br />

a ∈ D(E). The above result allows us to consider a map, which will be subsequently<br />

verified to be a derivation, provid<strong>in</strong>g a first (possibly left-degenerate)<br />

representation of Dirichlet forms.<br />

Def<strong>in</strong>ition 4.15. We will denote by ∂0 : B→H0 the follow<strong>in</strong>g l<strong>in</strong>ear map:<br />

for any a ∈Blet ∂0(a) be the vector <strong>in</strong> H0 associated, by the Riesz representation<br />

theorem, to the cont<strong>in</strong>uous l<strong>in</strong>ear functional la provided by the<br />

previous lemma. More explicitly, this map is characterized by<br />

(∂0(a)|b ⊗E c)H0 = la(b ⊗E c)<br />

= 1<br />

�<br />

E(a, bc)+E(b<br />

2<br />

∗ ,ca ∗ ) −E(b ∗ �<br />

a, c)<br />

b, c ∈B.<br />

(4.20)<br />

Notice that ||∂0(a)|| 2 ≤E[a] for all a <strong>in</strong> B and then, s<strong>in</strong>ce B is a form core<br />

H0<br />

for E, ∂0 can be extended to the whole form doma<strong>in</strong> D(E).<br />

Proposition 4.16. (Possibly left-degenerate representation of C ∗ –Dirichlet<br />

forms) Let (E,D(E)) be a C ∗ –Dirichlet form. Then we have:


Dirichlet Forms on Noncommutative Spaces 237<br />

i) ∂0 : B→H0 is a A–bimodule derivation;<br />

ii) for any a <strong>in</strong> B and b <strong>in</strong> A one has<br />

∂0(a)b = a ⊗E b ; (4.21)<br />

iii) the right representation of A is non degenerate: H0A is dense <strong>in</strong> H0;<br />

iv) for any a <strong>in</strong> B, one has<br />

E(a, a) −�∂0(a)� 2 H0<br />

1<br />

= lim<br />

ε↓0 2 τ<br />

�<br />

∆<br />

I + ε∆ (a∗ �<br />

a) . (4.22)<br />

By the def<strong>in</strong>ition of ∂0 and Theorem 4.13, it is easy to check the identities<br />

and<br />

(∂0(a1a2) − a1∂0(a2)+∂0(a1)a2|c ⊗E d)H0 =0<br />

(∂0(a)|c ⊗ db ∗ )H0 =(a ⊗E b|c ⊗E d)H0 ,<br />

for all a1,a2,a,b,c,d <strong>in</strong> B. From these identities the statements i) and ii)<br />

follow by cont<strong>in</strong>uity. The ma<strong>in</strong> po<strong>in</strong>t <strong>in</strong> statement iii) is to prove the existence<br />

of the limit. We omit the long proof and refer to [CS1 Proposition 4.4 iii)].<br />

Corollary 4.17. (Representation of Dirichlet forms generat<strong>in</strong>g conservative<br />

semigroups) Let (E,D(E)) be a C∗ –Dirichlet form correspond<strong>in</strong>g to a conservative<br />

Markovian semigroup:<br />

Tt(1M) =1M<br />

t ≥ 0 .<br />

Then we have<br />

E[a] =�∂0(a)� 2 a ∈ D(E) .<br />

H0<br />

To remedy the possible degeneracy of the left A-module structure of H0, and<br />

the asymmetric aspect of the right hand side of (4.22), we have to analyze<br />

further the structure of the A-bimodule H0. The result is that a suitable<br />

weight on A will take care of the possible degeneracy of the left action.<br />

Theorem 4.18. (Non degenerate representation of C ∗ –Dirichlet forms) Let<br />

(E,D(E)) be a C ∗ –Dirichlet form. Then there exists a symmetric derivation<br />

(∂1 , H1) <strong>in</strong> a nondegenerate A-bimodule and a weight K on A such that the<br />

follow<strong>in</strong>g representation holds true<br />

E(ξ,ξ) =�∂1(ξ)� 2 1<br />

H1 +<br />

2 K(ξ∗ξ + ξξ ∗ ) ξ ∈ D(E) . (4.23)<br />

In the conservative case K =0and (∂0 , H0) =(∂1 , H1)<br />

We omit the long proof of the result and refer the <strong>in</strong>terested reader to [CS1<br />

Theorem 8.1]. It is also possible to show [CS1 Theorem 8.2], that even the<br />

quadratic form on the right hand side of (4.23), depend<strong>in</strong>g on the weight K,


238 F. Cipriani<br />

may be described <strong>in</strong> terms of a derivation. The A-bimodule correspond<strong>in</strong>g to<br />

K is completely degenerate.<br />

Under the stronger hypothesis that the doma<strong>in</strong> of the generator ∆ conta<strong>in</strong>s a<br />

subalgebra dense <strong>in</strong> A, the last result has been obta<strong>in</strong>ed <strong>in</strong> [S2,3]. In that case<br />

the tangent bimodule supports a richer structure of a C∗-bimodule (see [Co2]<br />

for the def<strong>in</strong>ition). The simplest example of such a situation is the one of the<br />

Dirichlet <strong>in</strong>tegral of a Riemannian manifold M. There one has not only the<br />

Hilbert bimodule structure of square <strong>in</strong>tegrable vector fields L2 (TM)butalso<br />

the C∗-bimodule C0(TM) of cont<strong>in</strong>uous vector fields vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity.<br />

The fact that Dirichlet forms may be represented <strong>in</strong> terms of modules over<br />

C∗-algebras, allows to study them along the various decomposition theories<br />

of representations of C∗-algebras (see [Dix1]). To illustrate this po<strong>in</strong>t of view,<br />

we re-consider a famous formula due to A. Beurl<strong>in</strong>g-J. Deny [BD2] and Y. Le<br />

Jan [LJ] concern<strong>in</strong>g the decomposition of Dirichlet forms on locally compact<br />

spaces.<br />

Example 4.19. Dirichlet forms on commutative C ∗ –algebras: Beurl<strong>in</strong>g–<br />

Deny–Le Jan decomposition revisited I. Let X be a locally compact, metric<br />

space and m a positive Radon measure on X with full topological support.<br />

The Beurl<strong>in</strong>g–Deny–Le Jan decomposition (see also [FOT]) of a regular<br />

Dirichlet form (E,D(E)) on L 2 (X, m), represents E as a sum<br />

E = E (c) + E (j) + E (k)<br />

of Markovian forms, on the Dirichlet algebra B, called the diffusive, jump<strong>in</strong>g<br />

and kill<strong>in</strong>g parts, respectively.<br />

This decomposition may be obta<strong>in</strong>ed <strong>in</strong> a algebraic way from the representation<br />

of a Dirichlet form <strong>in</strong> terms of bimodules valued derivation, notic<strong>in</strong>g that<br />

a bimodule H on C0(X) is just a representation of the algebra C0(X × X).<br />

• The kill<strong>in</strong>g part has the form<br />

E (k) �<br />

(a, a) = |a| 2 dk<br />

for a positive Radon measure k on X. It corresponds to the degenerate<br />

part of the C0(X)-bimodule associated with E by our construction.<br />

• The jump<strong>in</strong>g part has the form<br />

E (j) �<br />

(a, a) = |a(x) − a(y)| 2 J(dx, dy)<br />

X×X−d<br />

for some symmetric, positive Radon measure J supported off the diagonal<br />

dX of X × X. It obviously appears as the form a ↦→ ||∂j(a)|| 2 ,where∂j is<br />

the universal derivation<br />

X<br />

∂j(a) =a ⊗ 1 − 1 ⊗ a


Dirichlet Forms on Noncommutative Spaces 239<br />

with values <strong>in</strong> the C0(X)-bimodule L 2 (X × X \ dX,J). Here left action of<br />

a ∈ C0(X) onf ∈ L 2 (X × X \ dX,J)isgivenby<br />

(a · f)(x, y) =a(x)f(x, y) J−a.e. (x, y) ∈ X × X \ dX ,<br />

while the right action is given by<br />

(f · a)(x, y) =f(x, y)a(y) J−a.e. (x, y) ∈ X × X \ dX .<br />

• The diffusive part is characterized by its strong local property:<br />

E (c) [a + b] =E (c) [a]+E (c) [b]<br />

whenever a, b ∈Band a is constant <strong>in</strong> a neighborhood of the support of<br />

b. By our analysis this part has the form a ↦→ ||∂c(a)|| 2 for a derivation<br />

Hc<br />

∂c with values <strong>in</strong> a C0(X)–monomodule Hc (where left and right actions<br />

co<strong>in</strong>cide). This is precisely the part of the C0(X)-bimodule associated with<br />

E by our construction, supported on the diagonal dX of X × X. Asthe<br />

representation theory of C0(X) is noth<strong>in</strong>g but measure theory on X, this<br />

means that the derivation ∂c may be decomposed accord<strong>in</strong>gly. The Hilbert<br />

module Hc may be represented as a Hilbert <strong>in</strong>tegral Hc = � ⊕<br />

X Hxµ(dx) of<br />

Hilbert modules {Hx : x ∈ X}, µ be<strong>in</strong>g a suitable Borel measure on X,<br />

where the C0(X)-module structure of Hx, for each fixed x ∈ X, isgivenby<br />

(a · ξ) =a(x)ξ a ∈ C0(X) , ξ ∈Hx .<br />

The derivation ∂c then decomposes as a direct <strong>in</strong>tegral ∂c = � ⊕<br />

X ∂x c µ(dx)<br />

of derivations {∂ x c : x ∈ X} and the correspond<strong>in</strong>g Leibniz rule appears as<br />

follows<br />

∂ x c (ab)(x) =a(x)∂ x c (b)(x)+b(x)∂ x c (a)(x) µ−a.s. x ∈ X.<br />

It may be proved that the class of the C0(X)-module Hc corresponds to<br />

the class of the measure µ and both are identified by the E-polar subsets<br />

of X.<br />

This illustrates how <strong>in</strong> the commutative situation the analysis we developed<br />

so far, not only gives back the classical Beurl<strong>in</strong>g–Deny–Le Jan decomposition,<br />

but also provides an algebraic <strong>in</strong>terpretation of each of the three parts plus<br />

a description of the diffusion piece <strong>in</strong> terms of a local derivation.<br />

Example 4.20. Dirichlet forms on group C ∗ –algebras associated to<br />

functions of negative type. Let Γ be a countable discrete group, with<br />

unit e ∈ Γ , and denote its elements by s,t,.... On the Hilbert space l 2 (Γ ),<br />

endowed with its natural basis {ɛt : t ∈ Γ }, consider the left and right regular<br />

representations def<strong>in</strong>ed as follows:


240 F. Cipriani<br />

λ : Γ → B(l 2 (Γ )) λ(s)ɛt := ɛst s, t ∈ Γ<br />

ρ : Γ → B(l 2 (Γ )) ρ(s)ɛt := ɛts s, t ∈ Γ.<br />

The same symbols, λ and ρ, will denote also the associated representations<br />

of the convolution algebra l 1 (Γ ). The reduced C ∗ -algebra of Γ is def<strong>in</strong>ed as<br />

the norm closure,<br />

C ∗ red (Γ )=λ(l1 (Γ )) ⊆ B(l 2 (Γ )) ,<br />

of the family of convolution operators associated to λ. The unique trace-state<br />

is def<strong>in</strong>ed by<br />

τ : C ∗ red(Γ ) → C τ(λ(s)) := δs,e s ∈ Γ.<br />

S<strong>in</strong>ce τ(λ(s) ∗ λ(t)) = δs,t, the von Neumann algebra associated to the GNS<br />

representation of τ is isomorphic to the von Neumann algebra generated by<br />

the left regular representation λ, i.e. the weak closure λ(l 1 (Γ )) ′′ of λ(l 1 (Γ ))<br />

<strong>in</strong> B(l 2 (Γ )). The correspond<strong>in</strong>g standard form, described <strong>in</strong> Example 2.20<br />

lies <strong>in</strong> the Hilbert space l 2 (Γ ). The positive cone co<strong>in</strong>cides with the subset<br />

of positive def<strong>in</strong>ite sequences, while the symmetric embedd<strong>in</strong>g of λ(l 1 (Γ )) ′′<br />

<strong>in</strong>to l 2 (Γ ), extends the map λ(a) ↦→ a from λ(l 1 (Γ )) to l 2 (Γ ).<br />

Consider now a positive, cont<strong>in</strong>uous function d : G → [0, +∞], of negative<br />

type, <strong>in</strong> the sense that<br />

n�<br />

i,j=1<br />

cicjd(s −1<br />

i sj) ≤ 0 s1 ,... ,sn ∈ Γ<br />

whenever c1 ,... ,cn ∈ C are such that � n<br />

i=1 ci = 0. Examples, <strong>in</strong>clud<strong>in</strong>g<br />

word-length functions of f<strong>in</strong>itely generated groups, and <strong>in</strong> particular on free<br />

and Coxeter groups, may be found <strong>in</strong> [Boz], where it is also shown that the<br />

class of these groups is stable under free products. The case of the wordlength<br />

functions on the free groups is explicitly illustrated <strong>in</strong> Section 12.2 of<br />

the Philippe Biane’s Lectures <strong>in</strong> this volume.<br />

A positive, cont<strong>in</strong>uous function of negative type def<strong>in</strong>es a semigroup {T d t :<br />

t ≥ 0} of completely positive contractions of C∗ red (Γ ), characterized by<br />

T d t (λ(f)) = λ(e−td f) f ∈ l 1 (Γ ) , t ≥ 0 .<br />

The dilation of this completely Markovian semigroup corresponds to a quantum<br />

stochastic process on C∗ red (Γ ) which is studied <strong>in</strong> Section 12 of the<br />

Phillippe Biane’s Lectures <strong>in</strong> this volume. The semigroup, which can be seen<br />

to be of strong Feller type [S8], is easily seen to be symmetric with respect<br />

to the trace-state, and the associated completely Dirichlet form is given by<br />

Ed : l 2 (Γ ) → [0, +∞) Ed[a] = �<br />

s∈Γ<br />

d(s)|a(s)| 2 .


Dirichlet Forms on Noncommutative Spaces 241<br />

The form is regular because, the group algebra CΓ ⊆ l 1 (Γ ) ∩ l 2 (Γ ) of f<strong>in</strong>ite<br />

support functions, on which E is clearly f<strong>in</strong>ite, is both a form core and a norm<br />

dense subspace of C∗ red (Γ ).<br />

To describe the derivation associated to the Dirichlet form Ed, we recall that<br />

positive, cont<strong>in</strong>uous, functions of negative type d on Γ , may be represented<br />

<strong>in</strong> terms of 1–cocycles of unitary representations of Γ . There exists <strong>in</strong> fact<br />

a unitary representation π : Γ → B(H) and a 1–cocycle c : Γ → H, by<br />

def<strong>in</strong>ition a map satisfy<strong>in</strong>g<br />

such that<br />

c(st) =c(s)+π(s)c(t) s, t ∈ Γ,<br />

�c(s)� 2 H = d(s)<br />

(c(s) | c(t))H = 1<br />

2 (d(t)+d(s) − d(s−1 t)) .<br />

As customary, the same symbol π will denote both the unitary representation<br />

of Γ and the <strong>in</strong>duced representation of C∗ red (Γ ). Def<strong>in</strong><strong>in</strong>g a map ∂ : D(∂) →<br />

H ⊗ l2 (Γ )by<br />

D(∂) :=λ(CΓ ) , ∂(λ(f)) := c ⊗ f, f∈ CΓ ,<br />

it is easy to check that the follow<strong>in</strong>g representation holds true<br />

E[a] =�∂(a)� 2 H⊗l 2 (Γ )<br />

a ∈ D(∂) .<br />

A C ∗ red (Γ )–bimodule structure on the Hilbert space H ⊗ l2 (Γ ) is def<strong>in</strong>ed<br />

by consider<strong>in</strong>g the left and right actions given by the representations πl :=<br />

π ⊗ λ and πr := id ⊗ ρ, respectively (the symbol id denotes the identity<br />

representation of Γ on H). This bimodule structure turns out to be symmetric<br />

with respect to the anti-l<strong>in</strong>ear <strong>in</strong>volution given by<br />

J (ξ ⊗ a) :=ξ ⊗ J(a) ξ ⊗ a ∈ H ⊗ l 2 (Γ ) ,<br />

where J(a)(s) = a(s −1 ), s ∈ Γ , is just the <strong>in</strong>volution associated to the<br />

standard cone. To verify the Leibniz rule it is enough to prove that for all<br />

s, t ∈ Γ<br />

∂(λ(ɛs) · λ(ɛt)) = πl(λ(ɛs))(∂(λ(ɛt))) + πr(λ(ɛt))(∂(λ(ɛs))) . (4.24)<br />

In fact we have<br />

πl(λ(ɛs))(∂(λ(ɛt))) + πr(λ(ɛt))(∂(λ(ɛs))) =<br />

πl(λ(ɛs))(c(t) ⊗ ɛt)) + πr(λ(ɛt))(c(s) ⊗ ɛs)) =<br />

π(λ(ɛs))c(t) ⊗ λ(ɛs)ɛt + c(s) ⊗ ρ(ɛt)ɛs =<br />

π(s)c(t) ⊗ ɛst + c(s) ⊗ ɛst =<br />

(c(s)+π(s)c(t)) ⊗ ɛst =


242 F. Cipriani<br />

c(st) ⊗ ɛst =<br />

∂(λ(ɛst)) =<br />

∂(λ(ɛs) · λ(ɛt)) .<br />

S<strong>in</strong>ce d vanishes on e ∈ Γ , and the identity of C ∗ red (Γ )isλ(δe), it is easy<br />

to see that the semigroup is conservative, so that the tangent bimodule H0<br />

associated with the Dirichlet form as <strong>in</strong> Corollary 4.17 is a sub–C∗ red (Γ )–<br />

bimodule of H ⊗ l2 (Γ ) and the kill<strong>in</strong>g weight vanishes identically.<br />

Example 4.21. Clifford Dirichlet form of free fermion systems. The<br />

follow<strong>in</strong>g one is historically the first example of noncommutative Dirichlet<br />

form. It was studied by L. Gross from the po<strong>in</strong>t of view of application to<br />

Quantum Field <strong>Theory</strong> [G1,2]. Let h be an <strong>in</strong>f<strong>in</strong>ite dimensional, separable,<br />

real Hilbert space and Cl(h) the complexification of the Clifford algebra over<br />

h. See Section 5.2 below for a detailed construction of the Clifford algebra,<br />

when h is f<strong>in</strong>ite dimensional. It is well known that Cl(h) isasimpleC∗ –<br />

algebra with a unique trace state τ. The associated von Neumann algebra<br />

be<strong>in</strong>g hyperf<strong>in</strong>ite type II1 factor.<br />

By the Chevalley–Segal isomorphism, here denoted by D, L2 (A, τ) canbe<br />

canonically identified with (the complexification of) the antisymmetric Fock<br />

space Γ (h) overh. L. Gross [G1,2] showed that the Second Quantization<br />

Γ (Ih) of the identity operator Ih on h is isomorphic to the Number operator<br />

N = D−1Γ (Ih)D, which is the generator of a conservative C∗ –semigroup over<br />

Cl(h). To describe the structure of N, let{ei : i ∈ N} be an orthonormal base<br />

of h and let {Ai : i ∈ N} be the correspond<strong>in</strong>g set of annihilation operators<br />

on Γ (h). For each i ∈ N the operator Di := D−1AiD, def<strong>in</strong>ed on the doma<strong>in</strong><br />

D( √ N), is a densely def<strong>in</strong>ed, closed operator with values <strong>in</strong> L2 (A, τ) and<br />

N = �<br />

D ∗ i Di = D −1 ( �<br />

A ∗ i Ai)D.<br />

i∈N<br />

Moreover, D( √ N) is a sub–algebra of Cl(h) and on it the follow<strong>in</strong>g graded<br />

Leibniz rules hold true:<br />

i∈N<br />

Di(ab) =Di(a))b + γ(a)(Di(b)) .<br />

Here γ is the extension to L 2 (A, τ) of the canonical <strong>in</strong>volution of Cl(h) which<br />

is the unique extension of the map v ↦→ −v on h. This shows that, by consider<strong>in</strong>g<br />

on L 2 (A, τ) the GNS right action of Cl(h) and the new left action<br />

given by γ, we obta<strong>in</strong> a sequence of closed derivations on Cl(h) withvalues<br />

<strong>in</strong> L 2 (A, τ). The C ∗ –Dirichlet form then is given by the formula<br />

E[a] =||N 1/2 (a)|| 2 H<br />

�<br />

= ||Di(a)|| 2 L2 (A,τ) ,<br />

where the tangent bimodule is now a sub-bimodule of �<br />

i∈N L2 (A, τ) andthe<br />

derivation is �<br />

i∈N Di.<br />

i∈N


Dirichlet Forms on Noncommutative Spaces 243<br />

Example 4.22. Heat semigroup on the noncommutative torus. This is<br />

a classical example com<strong>in</strong>g from Noncommutative Geometry [Co2], Let Aθ<br />

be the non commutative 2-torus, i.e. the universal C ∗ -algebra generated by<br />

two unitaries U and V , satisfy<strong>in</strong>g the relation<br />

VU = e 2iπθ UV .<br />

Let τ : Aθ → C be the tracial state given by<br />

τ(U n V m )=δn,0δm,0<br />

n, m ∈ Z .<br />

The heat semigroup {Tt : t ≥ 0} on Aθ is characterized by<br />

Tt(U n V m )=e −t(n2 +m 2 ) U n V m<br />

n, m ∈ Z .<br />

It is τ-symmetric, and the associated Dirichlet form is given by<br />

� �<br />

E αn,mU n V m�<br />

= �<br />

n,m∈Z<br />

n,m∈Z<br />

The derivation associated to E is the direct sum<br />

∂(a) =∂1(a) ⊕ ∂2(a)<br />

of the follow<strong>in</strong>g derivations ∂1 and ∂2 def<strong>in</strong>ed by<br />

(n 2 + m 2 )|αn,m| 2 .<br />

∂1(U n V m )=<strong>in</strong>U n V m , ∂2(U n V m )=imU n V m<br />

n, m ∈ Z .<br />

The heat semigroup is clearly conservative, and the Aθ–bimodule H0 associated<br />

with E, as <strong>in</strong> Corollary 4.17, is a sub–bimodule of the direct sum of two<br />

standard bimodules L 2 (A, τ) ⊕ L 2 (A, τ).<br />

4.4 Derivations and Dirichlet Forms<br />

<strong>in</strong> Free <strong>Probability</strong><br />

Free <strong>Probability</strong> is a noncommutative probabilistic theory, developed by<br />

D. Voiculescu, based on a suitable choice of the notion of <strong>in</strong>dependence of<br />

random variables, known as free <strong>in</strong>dependence. The orig<strong>in</strong> of this type of <strong>in</strong>dependence<br />

may be found <strong>in</strong> the asymptotic behavior of empirical spectral<br />

distributions of large random matrices and <strong>in</strong> properties shared by certa<strong>in</strong><br />

subgroups of free groups. Among the various fields of applications, a number<br />

of already avaible results suggest that the theory is a useful tool <strong>in</strong> approach<strong>in</strong>g<br />

long stand<strong>in</strong>g problems <strong>in</strong> von Neumann algebra theory. An <strong>in</strong>troduction<br />

to the matter may be found <strong>in</strong> [Voi1].


244 F. Cipriani<br />

In the framework of Free <strong>Probability</strong> an important role is played by the relativeFreeEntropyχ<br />

∗ which is a free analogue of the Shannon’s entropy. A<br />

possible approach to the construction of the Free Entropy is based the free<br />

analogue of the Fisher’s <strong>in</strong>formation Φ ∗ which <strong>in</strong> turn depends on the noncommutative<br />

generalization of the Hilbert transform. In this section we show<br />

how the noncommutative Hilbert transform is related to a suitable derivation<br />

and a correspond<strong>in</strong>g Dirichlet form (see [Voi1], [Voi2], [Bi]).<br />

Let (M,τ) be a noncommutative probability space, i.e. a von Neumann algebra<br />

M with a faithful and normal trace τ on it. Let 1 ∈ B ⊆ M be a<br />

∗ -subalgebra and let X = X ∗ ∈ M be a noncommutative random variable.<br />

Denote by B[X] ⊆ M the ∗ -subalgebra generated by B and X and consider<br />

on B[X] ⊗alg B[X] theB[X]-bimodule structure given by<br />

c · (a ⊗ b) :=ca ⊗ b<br />

(a ⊗ b) · c := a ⊗ cb .<br />

If X and B are algebraically free <strong>in</strong> the sense that no algebraic relation exist<br />

between them, there exists a unique derivation ∂X : B[X] → B[X] ⊗alg B[X]<br />

such that<br />

∂XX =1⊗ 1<br />

∂Xb =0 b ∈ B.<br />

In other words, if B[X] is regarded as the algebra of noncommutative polynomials<br />

<strong>in</strong> the variable X with coefficients belong<strong>in</strong>g to the algebra B, the<br />

operator ∂X acts as the partial derivation with respect to the noncommutative<br />

variable X and the elements of B play the role of constants. More<br />

explicitly one has<br />

∂X(b0Xb1X...Xbn) =<br />

=<br />

n�<br />

b0Xb1 ...bk−1(1 ⊗ 1)bkX...Xbn<br />

k=1<br />

n�<br />

b0Xb1 ...bk−1 ⊗ bkX...Xbn<br />

k=1<br />

b0 ,...bn ∈ B.<br />

Denot<strong>in</strong>g by W ∗ (B[X]) the von Neumann subalgebra of M generated by<br />

B[X] we may consider on it the restriction of the trace τ and the standard<br />

representation <strong>in</strong> L 2 (W ∗ (B[X]),τ). The derivation may be considered as a<br />

densely def<strong>in</strong>ed map from the Hilbert space L 2 (W ∗ (B[X]),τ) to the Hilbert<br />

bimodule L 2 (W ∗ (B[X]) ⊗ W ∗ (B[X]),τ ⊗ τ).<br />

Under the assumption that 1 ⊗ 1 ∈ D(∂ ∗ X )itmaybeprovedthat<br />

• the derivation is closable<br />

• B[X] ⊂ D(∂ ∗ X ∂X) and <strong>in</strong> particular that X ∈ D(∂ ∗ X ∂X).<br />

The element<br />

J (X : B) :=∂ ∗ X(1 ⊗ 1) ∈ L 2 (W ∗ (B[X]),τ)


Dirichlet Forms on Noncommutative Spaces 245<br />

is then called the noncommutative Hilbert transform of X with respect to B<br />

and the square of its norm<br />

Φ ∗ (X : B) :=�J (X : B)� 2 2 = �∂∗ X 1 ⊗ 1�2 2 = �∂∗ X ∂X(X)� 2 2<br />

is by def<strong>in</strong>ition the relative free <strong>in</strong>formation of X with respect to B. Incase<br />

B = C and µX is the distribution of X given by<br />

�<br />

f(t)µX(dt) :=τ(f(X)) f ∈ C0(R) ,<br />

R<br />

the algebra W ∗ (C[X]) is identified with L∞ (R ,µX) andtheHilbertspace<br />

L2 (W ∗ (C[X]) ,τ) becomes L2 (R ,µX). The elements f ∈ C[X] arejustpolynomials<br />

on R and ∂Xf co<strong>in</strong>cides with the difference quotient (4.3)<br />

∂Xf(s, t) =<br />

� f(s)−f(t)<br />

s−t<br />

if s �= t<br />

f ′ (s) if s = t.<br />

In case the Radon-Nikodym derivative p := dµX<br />

dλ with respect to the Lebesgue<br />

measure λ exists <strong>in</strong> L3 (R ,λ), it is possible to prove that J (X : C) is,upto<br />

afactor2π, the Hilbert transform Hp of p:<br />

Hp(t) :=p.v. 1<br />

�<br />

p(s)<br />

ds .<br />

π t − s<br />

Apply<strong>in</strong>g Theorem 4.5 one has that the closure of the quadratic form<br />

EX[ξ] :=�∂Xξ� 2<br />

R<br />

F := B[X]<br />

def<strong>in</strong>ed on L2 (W ∗ (B[X]),τ) is a completely Dirichlet form generat<strong>in</strong>g a completely<br />

Markovian, conservative, tracially symmetric C∗ 0 -semigroup on the<br />

von Neumann algebra W ∗ (B[X]). The Dirichlet form EX is deeply connected<br />

with the relative free Fisher <strong>in</strong>formation and with the free entropy. In fact<br />

it has been shown by Ph. Biane [Bi] that the Hessian of the free entropy<br />

co<strong>in</strong>cides with EX on the doma<strong>in</strong> where the relative free Fisher <strong>in</strong>formation<br />

is f<strong>in</strong>ite. Moreover, the Dirichlet form EX satisfies the Po<strong>in</strong>caré <strong>in</strong>equality if<br />

and only if the random variable X is centered, has unital covariance and a<br />

semi-circular distribution ([Bi]).<br />

5 Noncommutative <strong>Potential</strong> <strong>Theory</strong> and Riemannian<br />

Geometry<br />

In this chapter we will explore certa<strong>in</strong> relationships between noncommutative<br />

potential theory and classical geometry (see [CS2], [DR1,2]). More precisely<br />

we will show the strong <strong>in</strong>terplay among noncommutative Dirichlet


246 F. Cipriani<br />

<strong>in</strong>tegrals and heat semigroups on one hand, and the sign of the curvature of<br />

a Riemannian manifold on the other hand.<br />

In any complete Riemannian manifold M without boundary, whose Ricci<br />

curvature is bounded from below, the heat semigroup e −t∆ , generated by<br />

the Laplace–Beltrami operator ∆, is Markovian, i.e. it is a strongly cont<strong>in</strong>uous,<br />

positivity preserv<strong>in</strong>g, contraction semigroup on the C ∗ -algebra C0(M)<br />

of complex cont<strong>in</strong>uous functions vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity.<br />

From an <strong>in</strong>f<strong>in</strong>itesimal po<strong>in</strong>t of view, one has, correspond<strong>in</strong>gly, that the<br />

Dirichlet <strong>in</strong>tegral E[f] := �<br />

M |∇f|2 is a Dirichlet form, i.e. a quadratic,<br />

lower semicont<strong>in</strong>uous functional on C0(M) satisfy<strong>in</strong>g the contraction property<br />

E[f ∧ 1] ≤E[f].<br />

These fundamental properties of classical potential theory are <strong>in</strong>dependent<br />

of the curvature of the space, even though the Remannian metric is, together<br />

with the concept of differential, the ma<strong>in</strong> <strong>in</strong>gredient for the def<strong>in</strong>ition of the<br />

gradient operator and the energy functional.<br />

<strong>Classical</strong> potential theory of Riemannian manifolds, suggests however that, if<br />

a relationship among curvature and some k<strong>in</strong>d of positivity exists, it should<br />

emerge consider<strong>in</strong>g energy functionals, heat equations and their solutions, on<br />

vector bundles over the manifold.<br />

The first candidates are obviously the exterior bundle Λ∗M, the Hodge-de<br />

Rham Laplacian operator ∆HdR on it and the heat semigroup e−t∆HdR it<br />

generates. However, while on scalar functions we have a natural notion of<br />

positivity, on exterior forms this is not obvious.<br />

The way to tie together the useful tools of exterior differential calculus with a<br />

good notion of positivity is to look at exterior forms, i.e. cont<strong>in</strong>uous sections<br />

C0(Λ ∗ M) of the exterior bundle Λ ∗ M, from the po<strong>in</strong>t of view of its isomorphic<br />

Clifford bundle Cl(M). Precisely, we will work with exterior forms as elements<br />

of the Clifford algebra C ∗ 0(M) , i.e. cont<strong>in</strong>uous sections of the Clifford bundle,<br />

vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity:<br />

C ∗ 0 (Λ∗M) ∼ = C ∗ 0 (M) .<br />

This is the most natural <strong>in</strong>volutive algebra, <strong>in</strong>deed a C∗-algebra, extend<strong>in</strong>g<br />

the algebra C0(M) of cont<strong>in</strong>uous functions, vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity on M.<br />

In this framework E.B. Davies and O. Rothaus proved <strong>in</strong> [DR1] that on<br />

any complete Riemannian manifold, the quadratic form associated to the<br />

covariant derivation ∇ on C ∗ 0 (M)<br />

�<br />

EB[σ] :=<br />

|∇σ| 2 dm<br />

M<br />

is a noncommutative Dirichlet form on C∗ 0 (M), with respect to a natural<br />

trace, so that the heat semigroup e−t∆B generated by the Bochner-Laplacian<br />

∆B := ∇ ∗ ◦∇


Dirichlet Forms on Noncommutative Spaces 247<br />

is a Markovian semigroup. In this chapter we provide a proof of the above<br />

result verify<strong>in</strong>g that ∇ is a closed derivation <strong>in</strong> the sense of Chapter 3.<br />

LaterweconsidertheDiracoperatorDon C∗ 0 (M). It was <strong>in</strong>troduced <strong>in</strong> the<br />

Riemannian sett<strong>in</strong>g by M. F. Atiyah and G. B. S<strong>in</strong>ger [AtS] (see also [LM]<br />

and reference there<strong>in</strong>), <strong>in</strong> their work on the celebrated Index Theorem (see<br />

[LM]). In the semi-Riemannian sett<strong>in</strong>g of relativistic quantum theory it was<br />

previously considered by P.A.M. Dirac [Dir].<br />

The Dirac operator is the fundamental elliptic operator of a Riemannian<br />

manifold; from the po<strong>in</strong>t of view of the exterior algebra it is just the sum of<br />

the exterior differential and its adjo<strong>in</strong>t<br />

D ∼ = d + d ∗ .<br />

Its square D 2 , the so called Dirac-Laplacian, is just a version, on the Clifford<br />

algebra, of the Hodge-de Rham-Laplacian on exterior forms<br />

D 2 ∼ = (d + d ∗ ) 2 = d ∗ d + dd ∗ =: ∆HdR .<br />

A first difference between these two Laplacians is that while ∆B reveals the<br />

geometric aspects of M, the operator D 2 is connected with the topological<br />

ones. For example, <strong>in</strong> the kernel of ∆B one f<strong>in</strong>ds the parallel forms, while the<br />

harmonic forms are conta<strong>in</strong>ed <strong>in</strong> the kernel of D 2 .<br />

The second ma<strong>in</strong> difference is that, while, by the Davies-Rothaus theorem,<br />

the heat semigroup e−t∆ is always Markovian, the semigroup e−tD2 need<br />

not be.<br />

We will show that the heat semigroup e−tD2, generated by the Dirac-<br />

Laplacian D2 , is Markovian if and only if the curvature operator of M is<br />

nonnegative. Equivalently, the quadratic form of the operator D2 ED : L 2 (C ∗ 0 (M) ,τ) −→ [0, +∞] ED[σ] =<br />

�<br />

M<br />

|Dσ| 2 ,<br />

is a noncommutative Dirichlet form exactly when the curvature operator � R<br />

of M is nonnegative.<br />

In this way, a geometric property, like the positive sign of the curvature operator,<br />

is shown to be equivalent to a potential theoretic property of the Dirac<br />

energy functional ED.<br />

Theorem 5.1. Let (M,g) be a complete Riemannian manifold without<br />

boundary. The follow<strong>in</strong>g properties are equivalent:<br />

i) {e−tD2 : t ≥ 0} is a completely Markovian semigroup, i.e. a strongly<br />

cont<strong>in</strong>uous, semigroup of completely positive contractions on C∗ 0(M);


248 F. Cipriani<br />

ii) {e−tD2 : t ≥ 0} is a completely Markovian semigroup on the Hilbert space<br />

L2 (Cl(M), g) of square <strong>in</strong>tegrable sections of the Clifford bundle;<br />

iii) the Dirac form ED is a C∗ –Dirichlet form on L2 (C∗ 0 (M), τ);<br />

iv) the curvature operator is nonnegative: � R ≥ 0.<br />

The methods we use are based on the Bochner Identity, on the correspondence<br />

between noncommutative Dirichlet forms and Markovian semigroups, and<br />

on the characterization of Dirichlet forms <strong>in</strong> terms of derivations which we<br />

presented <strong>in</strong> Chapter 4.<br />

Specializ<strong>in</strong>g the above general sett<strong>in</strong>g, let us review some applications.<br />

Example 5.2. (Dirichlet forms and topology of Riemann surfaces) Let Σ be<br />

a compact, connected, orientable surface. Then there exists a metric g on Σ<br />

such that the Dirac form ED is a Dirichlet form if and only if Σ is either<br />

homeomorphic to the sphere S2 or to the torus T 2 .<br />

In fact <strong>in</strong> dimension two the curvature operator is just the multiplication<br />

operator by the Gauss curvature function kg; if our surface is homeomorphic<br />

to a sphere or a torus, it then carries a metric of nonnegative curvature and<br />

the correspond<strong>in</strong>g Dirac form ED is a C∗-Dirichlet form. On the other hand<br />

if g is a metric for which ED is a C∗-Dirichlet form, then its Gauss curvature<br />

is nonnegative so that, by the Gauss–Bonnet formula χ(M) = 1<br />

�<br />

2π M kg, its<br />

Euler characteristic is nonnegative.<br />

Example 5.3. (Dirichlet forms on hypersurfaces) Another example <strong>in</strong> which<br />

the Dirac form ED is Markovian is the one of a codimension one convex,<br />

hypersurface M ⊂ R n+1 .<br />

Example 5.4. (Dirichlet forms and topology of Cartan’s symmetric spaces)<br />

Manifolds which have a nonnegative curvature operator have been <strong>in</strong>tensively<br />

studied (and, at the end classified) s<strong>in</strong>ce the works of E. Cartan on symmetric<br />

spaces. Comb<strong>in</strong><strong>in</strong>g these result with the above equivalences we obta<strong>in</strong> that<br />

if M is a compact simply connected Riemannian manifold whose Dirac form<br />

ED is a complete Dirichlet form then M is homeomorphic to the product of<br />

spaces of the follow<strong>in</strong>g type:<br />

i) Euclidean spaces,<br />

ii) Spheres,<br />

iii) Projective spaces,<br />

iv) Symmetric spaces of compact type.<br />

In the follow<strong>in</strong>g sections we will expla<strong>in</strong> carefully the <strong>in</strong>gredients, of geometric<br />

and analytic nature needed for the proof of Theorem 5.1. Very good references<br />

for all the background we are go<strong>in</strong>g to summarize are [LM], [Pet].


Dirichlet Forms on Noncommutative Spaces 249<br />

5.1 Clifford Algebras<br />

Let (E,g) be a real, Euclidean vector space, with f<strong>in</strong>ite dimension dim E. The<br />

Clifford algebra Cl(E,g) associated to (E,g) can be def<strong>in</strong>ed as the unital,<br />

associative, complex algebra generated by a unit 1 and the elements of E<br />

subject to the relations : e1·e2+e2·e1+2g(e1,e2)·1 = 0 for all e1,e2 <strong>in</strong> E. The<br />

product of two elements σ1,σ2 ∈ Cl(E,g) will be denoted by σ1 · σ2 ,andits<br />

unit by 1. There is a natural, vector space isomorphism φ : Λ ∗ (E) → Cl(E,g)<br />

between the complexified exterior algebra Λ ∗ (E) :=Λ ∗ R (E) ⊗R C and the<br />

Clifford algebra, characterized by<br />

φ(1Λ∗ (E)) =1,<br />

(5.1)<br />

φ(v1 ∧ v2 ∧···∧vr) =v1 · v2 ···vr<br />

for an orthogonal family v1,v2,...,vr ∈ E and r = 1,...,dim E. Consequently<br />

dimC Cl(E,g)=2dim E and if {ei : i =1,...,dim E} is an orthonormal<br />

basis for E, then the unit 1 and the set {ei1 · ei2 ·····eir :1≤ i1 <<br />

i2 < ···


250 F. Cipriani<br />

bundle whose fibers Clx(M) atpo<strong>in</strong>tsx ∈ M are just the Clifford algebras<br />

of the Euclidean spaces (TxM,gx).<br />

The above algebraic isomorphism between the Clifford algebra and the exterior<br />

algebra can be raised to a natural isomorphism of Euclidean bundles<br />

Cl(M) � Λ ∗ (M) between the Clifford bundle and the complexified exterior<br />

bundle Λ ∗ (M). The great advantage of this is that the exterior algebra of<br />

cont<strong>in</strong>uous differential forms, vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity on M, when considered as<br />

the space of sections of the Clifford bundle vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity acquires a<br />

natural structure of C ∗ -algebra.<br />

Def<strong>in</strong>ition 5.5. (Clifford algebra of a Riemannian manifold.) The space<br />

C∗ 0 (M) of sections of the Clifford bundle Cl(M) is called the Clifford algebra<br />

of the Riemannian manifold (M,g). The multiplication of two sections<br />

σ1 ,σ2, or sp<strong>in</strong>ors as they are also called, is def<strong>in</strong>ed po<strong>in</strong>twise (σ1 · σ2)(m) =<br />

(σ1)(m) · (σ2)(m) by the Clifford product <strong>in</strong> each Clifford algebra Clm(M).<br />

Endowed with the po<strong>in</strong>twise multiplication and <strong>in</strong>volution <strong>in</strong> the fibers, it<br />

is a C∗ –algebra with respect to the norm which is the supremum of operator<br />

norms on the fibers:<br />

�σ� := sup �σ(m)� . (5.3)<br />

m∈M<br />

Its center conta<strong>in</strong>s the algebra C0(M) of cont<strong>in</strong>uous complex functions vanish<strong>in</strong>g<br />

at <strong>in</strong>f<strong>in</strong>ity on M.<br />

The Clifford algebra carries a natural faithful, semif<strong>in</strong>ite l.s.c. trace τ :<br />

C ∗ 0 (M)+ → [0, +∞] obta<strong>in</strong>ed by glu<strong>in</strong>g together the traces {τm : m ∈ M} on<br />

the fibers by the Riemannian volume measure dm on M:<br />

�<br />

τ(σ) :=<br />

τm(σ(m)) dm σ ∈ C<br />

M<br />

∗ 0 (M)+ . (5.4)<br />

The GNS Hilbert space L2 (C∗ 0(M),τ) associated to τ co<strong>in</strong>cides with the<br />

Hilbert space L2 (Cl(M)), with scalar product<br />

(σ2|σ1) :=τ(σ ∗ 2 · σ1)<br />

�<br />

= τm(σ2(m)<br />

M<br />

∗ · σ1(m))dm<br />

�<br />

(5.5)<br />

= 〈σ2(m)|σ1(m)〉 Cl(TmM,gm) · dm.<br />

M<br />

The left and right regular representations of the Clifford algebra onto itself<br />

naturally extend, by cont<strong>in</strong>uity, to commut<strong>in</strong>g representations <strong>in</strong> the Hilbert<br />

space L2 (Cl(M)) = L2 (C∗ 0 (M),τ). These actions, denoted by (a, ξ, b) → aξb,<br />

satisfy<br />

�aξb�2 ≤�a�·�ξ�2 ·�b� a, b ∈ C ∗ 0 (M) ξ ∈ L2 (Cl(M)) (5.6)<br />

<strong>in</strong> such a way that L 2 (Cl(M)) is naturally endowed with a structure of a<br />

C ∗ 0(M)–bimodule.


Dirichlet Forms on Noncommutative Spaces 251<br />

Extend<strong>in</strong>g, aga<strong>in</strong> by cont<strong>in</strong>uity, the <strong>in</strong>volution of C ∗ 0(M) toL 2 (C ∗ 0(M),τ),<br />

we get an antil<strong>in</strong>ear, isometric <strong>in</strong>volution J : L 2 (C ∗ 0 (M),τ) → L2 (C ∗ 0 (M),τ)<br />

which exchanges the left and right actions of the Clifford algebra,<br />

J(aξb) =a ∗ J(ξ)b ∗<br />

a, b ∈ C ∗ 0(M) ξ ∈ L 2 (C ∗ 0(M),τ) , (5.7)<br />

so that L 2 (C ∗ 0 (M),τ)is<strong>in</strong>factasymmetric C∗ 0 (M)–bimodule.<br />

5.3 Covariant Derivative, Bochner<br />

and Dirac-Laplacians<br />

We now describe the basic elliptic operators whose potential theoretic properties<br />

we want to analyze.<br />

The Levi-Civita connection on the tangent bundle TM <strong>in</strong>duces a hermitian<br />

connection <strong>in</strong> the associated Clifford bundles. We shall denote the associated<br />

covariant derivative, act<strong>in</strong>g on the spaces of smooth sections of the bundle,<br />

by the symbol<br />

∇ : C ∞ c (Cl(M)) → C∞ c (Cl(M) ⊗ T ∗ M) . (5.8)<br />

By def<strong>in</strong>ition the covariant derivative satisfies the follow<strong>in</strong>g rules<br />

∇(fσ)=σ ⊗ df + f∇σ f ∈ C ∞ c (M) ,σ ∈ C ∞ c (Cl(M))<br />

X(σ1|σ2) =(∇Xσ1|σ2)+(σ1|∇X σ2) σ1 ,σ2 ∈ C ∞ c (Cl(M)) ,X ∈ C ∞ c (TM) ,<br />

(5.9)<br />

where ∇X denotes the composition of ∇ with the contraction associated to<br />

the vector field X.<br />

At the Hilbert space level the covariant derivative ∇ can be understood<br />

as a closed operator from L2 (C∗ 0(M),τ)=L2 (Cl(M)) to L2 (Cl(M) ⊗ T ∗M), def<strong>in</strong>ed on the Sobolev space H1 (M) ofelementsofL2 (Cl(M)) whose distributional<br />

covariant derivative belongs to L2 (Cl(M) ⊗ T ∗M). The Bochner-Lapalcian is then the nonnegative self-adjo<strong>in</strong>t operator on<br />

L2 (Cl(M)) def<strong>in</strong>ed as<br />

△B := ∇ ∗ ∇ . (5.10)<br />

Its associated closed quadratic form co<strong>in</strong>cides with<br />

EB := H 1 (M) → [0, ∞)<br />

EB[σ] :=�∇σ� 2 L2 (Cl(M)⊗T ∗M) σ ∈ H 1 (M) .<br />

(5.11)<br />

Under the natural isomorphism C ∗ 0 (M) � Λ∗ (M), the above def<strong>in</strong>ed operators<br />

∇, △B can be naturally identified with the more usual covariant<br />

derivative and Bochner-Laplacian act<strong>in</strong>g on the exterior bundle.


252 F. Cipriani<br />

The Dirac operator D : C ∞ (C ∗ 0(M)) → C ∞ (C ∗ 0(M)) is def<strong>in</strong>ed as<br />

Dσ :=<br />

n�<br />

ei ·∇eiσ (5.12)<br />

i=1<br />

at any po<strong>in</strong>t m ∈ M, where{ei : i =1,...,n} denotes any orthonormal base<br />

of TmM and “ · ” denotes the product <strong>in</strong> the Clifford algebra Cl(TmM,gm). At<br />

the Hilbert space level D is a self-adjo<strong>in</strong>t operator on L 2 (Cl(M)) with doma<strong>in</strong><br />

H 1 (M). The operator D 2 on L 2 (Cl(M)) is called the Dirac-Laplacian. Itis<br />

a nonnegative self-adjo<strong>in</strong>t operator whose closed quadratic form is given by<br />

ED : H 1 (M) → [0, ∞)<br />

ED[σ] :=�Dσ� 2 L 2 (Cl(M))<br />

σ ∈ H 1 (Cl(M)) .<br />

(5.13)<br />

Under the natural isomorphism Cl(M) � Λ ∗ (M), the Dirac operator D and<br />

the Dirac-Laplacian D 2 correspond to the operator d+d ∗ and to the Hodge-de<br />

Rham-Laplacian ∆HdR := dd ∗ + d ∗ d, respectively.<br />

5.4 Leibniz Rule for the Covariant Derivative<br />

Here we show that the Bochner-Laplacian generates a completely Markovian<br />

semigroup on the Clifford algebra of a Riemannian manifold [DR1 Theorem<br />

13]. The proof is different from the orig<strong>in</strong>al one and it relies on the Leibniz<br />

property of the covariant derivative.<br />

On the Hilbert space L2 (Cl(M)⊗T ∗M) there are cont<strong>in</strong>uous and commut<strong>in</strong>g<br />

left and right actions of the Clifford algebra C∗ 0 (M) which are characterized<br />

by the rule<br />

ξ · (ζ ⊗ ω) · η =(ξ · ζ · η) ⊗ ω, (5.14)<br />

for all ξ,ζ,η ∈ C ∗ 0 (M) and all sections ω of the cotangent bundle T ∗ M.<br />

Moreover, s<strong>in</strong>ce T ∗ M is the complexification of a real vector bundle, there is<br />

on L 2 (T ∗ M) a canonical conjugation whose action we denote by ω → ω. The<br />

map<br />

J : L 2 (Cl(M) ⊗ T ∗ M) → L 2 (Cl(M) ⊗ T ∗ M)<br />

J (ζ ⊗ ω) :=ζ ∗ ⊗ ω<br />

(5.15)<br />

is then easily seen to be a natural, antil<strong>in</strong>ear, isometric <strong>in</strong>volution, which<br />

<strong>in</strong>tertw<strong>in</strong>es the left and right actions of C ∗ 0 (M) onL2 (Cl(M) ⊗ T ∗ M):<br />

J (ξ · ζ ⊗ ω · η) :=η ∗ ·J(ζ ⊗ ω) · ξ ∗ . (5.16)<br />

Summariz<strong>in</strong>g, we endowed the Hilbert space L 2 (Cl(M)⊗T ∗ M) with a canonical<br />

structure of symmetric C ∗ 0 (M)–bimodule.


Dirichlet Forms on Noncommutative Spaces 253<br />

Theorem 5.6. Let (M,g) be a Riemmanian manifold without boundary and<br />

C ∗ 0 (M) its associated Clifford C∗ -algebra. Then the covariant derivative def<strong>in</strong>ed<br />

on the Dirichlet algebra B(M) :=C ∗ 0 (M) ∩ H1 (M)<br />

∇ : B(M) → L 2 (Cl(M) ⊗ T ∗ M) (5.17)<br />

is a symmetric derivation <strong>in</strong> the sense that B(M) is a ∗ –algebra, the Leibniz<br />

rule<br />

∇(σ1 · σ2) =(∇σ1) · σ2 + σ1 · (∇σ1) σ1 ,σ2 ∈B(M) (5.18)<br />

holds true and that ∇ <strong>in</strong>tertw<strong>in</strong>es the <strong>in</strong>volutions J and J<br />

∇(Jσ)=∇(σ ∗ )=J (∇σ) σ ∈B(M) . (5.19)<br />

As a consequence, the Bochner-Laplacian ∆B generates a completely Markovian<br />

semigroup, both on the Hilbert space L 2 (Cl(M) of square <strong>in</strong>tegrable sections<br />

of the Clifford bundle and on the Clifford C ∗ -algebra C ∗ 0 (M).<br />

Proof. The symmetry property (5.19) follows directly from the def<strong>in</strong>ition of<br />

J <strong>in</strong> (5.15). The Leibniz property of ∇ is just a consequence of the metric<br />

property (5.9) of the Levi-Civita connection. In fact, for any vector field X,<br />

from (5.9) and the properties of the Clifford product we have<br />

X(σ|σ) =(∇Xσ|σ)+(σ|∇Xσ)<br />

∇X(σ · σ) =(∇Xσ) · σ + σ · (∇Xσ) σ ∈B(M) .<br />

S<strong>in</strong>ce the contraction iX, with respect to a vector field X, commutes with<br />

the left and right actions of the Clifford algebra, and ∇X = iX ◦∇,wehave<br />

iX(∇(σ · σ)) = iX((∇σ) · σ + σ · (∇σ)) σ ∈B(M) .<br />

As this is true for any vector field, we have<br />

∇(σ · σ) =(∇σ) · σ + σ · (∇σ) σ ∈B(M) ,<br />

from which (5.18) follows by polarization.<br />

5.5 Curvature Operator and Bochner Identity<br />

A large number of results <strong>in</strong> topology and geometry on Riemannian manifolds,<br />

follows from the so called Bochner method, based on the splitt<strong>in</strong>g of the<br />

Dirac Laplacian as the sum of the Bochner-Laplacian and a piece depend<strong>in</strong>g<br />

the curvature of the manifold. This second part is described <strong>in</strong> terms of<br />

(a bundle of) f<strong>in</strong>ite dimensional self-adjo<strong>in</strong>t operators called the curvature<br />

operator.


254 F. Cipriani<br />

The curvature tensor R associated to the Levi–Civita connection of (M,g)<br />

can be def<strong>in</strong>ed, as a smooth section of ⊗ 4 T ∗ M,atthepo<strong>in</strong>tm ∈ M by the<br />

formula<br />

Rm(v1,v2,v3,v4) =(Rm(v1,v2)v3|v4)TmM v1,v2,v3,v4 ∈ TmM, (5.20)<br />

where, for fixed vector fields v1, v2 , Rm(v1,v2) is the endomorphism of TmM<br />

given by<br />

Rm(v1,v2)v = −(∇v1∇v2v −∇v2∇v1v −∇ [v1,v2]v)(m) v ∈ C ∞ c (TM)<br />

(5.21)<br />

which depends only on values of v1, v2 and v at po<strong>in</strong>t m.<br />

The symmetries of the curvature tensor (not <strong>in</strong>clud<strong>in</strong>g the Bianchi’s identities)<br />

allow to <strong>in</strong>terpret Rm as a self-adjo<strong>in</strong>t operator � Rm on the Hilbert space<br />

Λ2 mM accord<strong>in</strong>g to the formula:<br />

( � Rm(v1 ∧ v2)| v3 ∧ v4) Λ 2 m M := Rm(v1,v2,v3,v4) v1,v2,v3,v4 ∈ TmM .<br />

(5.22)<br />

Def<strong>in</strong>ition 5.7. (Curvature operator of a Riemannian manifold) The field<br />

�R := { � Rm : m ∈ M} is called the curvature operator of the Riemannian<br />

manifold (M,g). It is said to be nonnegative if, at each po<strong>in</strong>t m of M, � Rm<br />

is nonnegative operator <strong>in</strong> the spectral sense, i.e. if all of its eigenvalues are<br />

nonnegative. This property will be denoted by � R ≥ 0.<br />

The curvature operator is related to the sectional curvature K : Gr(2,M) →<br />

R of M, def<strong>in</strong>ed on the Grassmannian Gr(2,M) of 2–planes of M by the<br />

formula<br />

Km(πm) =( � Rm(v1 ∧ v2)|v1 ∧ v2) Λ 2 m M πm ∈ Grm(2,M) (5.23)<br />

for any orthonormal base {v1,v2} of the 2–plane πm at m. The curvature tensor<br />

R, the curvature operator � R and the sectional curvature K are equivalent<br />

algebraic objects <strong>in</strong> the sense that each of them determ<strong>in</strong>es algebraically the<br />

others. Attention should be paid to the fact that, while � R ≥ 0 implies K ≥ 0,<br />

the opposite is not always true. The reason for that is that at some po<strong>in</strong>t m,<br />

�Rm may not admit a base of decomposable eigenvectors. Conditions for this<br />

are well known and some of them will be used later <strong>in</strong> the work.<br />

The proof of Theorem 5.1 is based on a repeated use of the Bochner Identity<br />

(see [LM Theorem 8.2])<br />

D 2 = ∆B + ΘR<br />

(5.24)<br />

by which the Dirac-Laplacian D 2 decomposes as a sum of the Bochner-<br />

Laplacian ∆B and a symmetric operator ΘR depend<strong>in</strong>g only on the curvature<br />

operator � R. In terms of quadratic forms on L 2 (Cl(M)), the decomposition<br />

reads as follows (see [LM Theorem 8.6])


Dirichlet Forms on Noncommutative Spaces 255<br />

ED = EB + QR<br />

(5.25)<br />

where QR is the quadratic form QR[σ] :=(σ|ΘR(σ)) L2 (Cl(M)) of the operator<br />

ΘR. The quadratic form QR can be explicitly represented as a superposition<br />

�<br />

QR[σ] = QR(m)[σm] · dm σ ∈ L 2 (Cl(M)) (5.26)<br />

M<br />

of quadratic forms on the Hilbert spaces Cl(TmM,gm) givenby<br />

QR(m)[σ] = 1<br />

4<br />

n(n−1)/2 �<br />

α,β=1<br />

� �Rm(ξα) | ξβ<br />

�<br />

Λ 2 m M · � [ξα,σ] | [ξβ,σ] �<br />

Cl(TmM,gm)<br />

(5.27)<br />

where {ξα : α =1,...,n(n − 1)/2} is any orthonormal base of Λ2 mM.Ifwe choose a base {ηα : α =1,...,n(n−1)/2} of orthonormal eigenvectors of � Rm<br />

correspond<strong>in</strong>g to the eigenvalues {µα : α =1,...,n(n − 1)/2}, then (2.38)<br />

becomes<br />

QR(m)[σ] = 1<br />

n(n−1)/2 �<br />

µα� [ηα,σ] �<br />

4<br />

2 2 . (5.28)<br />

α=1<br />

5.6 Completely Unbounded and Approximately<br />

Bounded Parts of a Dirichlet Form<br />

Here we describe a general decomposition of derivations and Dirichlet forms,<br />

which will be useful for the proof of the ma<strong>in</strong> result of this chapter.<br />

Def<strong>in</strong>ition 5.8. Let ∂ : D(∂) →Hbe a derivation, densely def<strong>in</strong>ed on a<br />

subalgebra of a C∗-algebra A (see Def<strong>in</strong>ition 4.2) and denote by LA−A(H),<br />

the von Neumann algebra of all bounded operators on H, commut<strong>in</strong>g both<br />

with the left and right actions of A.<br />

• An element B ∈LA−A(H) issaidtobe∂-bounded if the map<br />

B ◦ ∂ : D(∂) →H<br />

extends to a bounded map from A to H.<br />

• A projection p <strong>in</strong> LA−A(H) issaidtobeapproximately ∂-bounded if it is<br />

the <strong>in</strong>creas<strong>in</strong>g limit of a net of ∂-bounded projections.<br />

• A projection p <strong>in</strong> LA−A(H) issaidtobecompletely ∂-unbounded if 0 is<br />

the only ∂-bounded projection smaller than p.<br />

• The derivation (∂,D(∂) , H) issaidtobebounded (resp. approximately<br />

bounded, completely unbounded) if the identity operator 1H is bounded<br />

(resp. approximately bounded, completely unbounded).


256 F. Cipriani<br />

Example 5.9. (Structure of bounded completely Dirichlet forms on amenable<br />

C ∗ -algebras) By def<strong>in</strong>ition (see [Co2]), amenable C ∗ algebras are those on<br />

which any bounded derivation ∂ : A → X, tak<strong>in</strong>g values <strong>in</strong> a dual Banach<br />

bimodule X (i.e. a bimodule X such that, as a Banach space it is the dual<br />

X =(X∗) ∗ of some other X∗), are necessarily <strong>in</strong>ner: thereexistsξ ∈ X such<br />

that<br />

∂(a) =aξ − ξa a ∈ A.<br />

Recall that by results due to A. Connes [Co3] and U. Haagerup [H3], A is<br />

amenable if and only if is nuclear. S<strong>in</strong>ce the derivations represent<strong>in</strong>g C ∗ -<br />

Dirichlet forms take values <strong>in</strong> Hilbert bimodule (see Chapter 4), we conclude<br />

that if A is amenable, τ : A → C is a f<strong>in</strong>ite trace on A, andE : L 2 (A,τ) →<br />

[0, +∞) is a bounded C ∗ -Dirichlet form, then it can be represented as follows<br />

for some ξ ∈H.<br />

E[a] =�aξ − ξa� 2 H<br />

a ∈ A<br />

Proposition 5.10. (Decomposition of derivations and Dirichlet forms [CS2<br />

Lemma 4.3]) Let ∂ : D(∂) →Hbe a derivation, densely def<strong>in</strong>ed on a subalgebra<br />

of a C ∗ -algebra A. Then there exists <strong>in</strong> LA−A(H), the greatest approximately<br />

∂-bounded projection Pab. Every ∂-bounded operator B satisfies<br />

B ◦ Pab = B.<br />

As a consequence, the derivation (∂,D(∂) , H) splits, canonically, as direct<br />

sum<br />

∂ = ∂cu ⊕ ∂ab<br />

of a smallest completely unbounded derivation<br />

∂cu : D(∂) →Hcu ∂cu := (1H − Pcu) ◦ ∂, Hcu := (1H − Pab)H<br />

and the greatest approximately bounded derivation<br />

∂ab : D(∂) →Hab ∂ab := Pab ◦ ∂, Hab := PabH .<br />

Apply<strong>in</strong>g the decomposition above to the derivation associated to a Dirichlet<br />

form E, one gets a decomposition of E as a sum of two, not necessarily closable,<br />

Markovian forms<br />

E = Ecu + Eab<br />

called the completely unbounded part and the approximately bounded one,<br />

respectively. To clarify the mean<strong>in</strong>g of the above splitt<strong>in</strong>g, let us apply it <strong>in</strong><br />

the commutative sett<strong>in</strong>g.<br />

Example 5.11. (Decomposition of Dirichlet forms on commutative C ∗ -algebras:<br />

the Beurl<strong>in</strong>g-Deny-Le Jan decomposition revisited II). By the classical


Dirichlet Forms on Noncommutative Spaces 257<br />

Beurl<strong>in</strong>g-Deny-Le Jan decomposition BDLJ (see [FOT]), a regular Dirichlet<br />

form (E,D(E)) on L 2 (X, m), splits as a sum<br />

E = E (c) + E (j) + E (k)<br />

of three Markovian forms, called the diffusive, jump<strong>in</strong>g and kill<strong>in</strong>g parts,<br />

respectively. In Example 4.19, we this decomposition was obta<strong>in</strong>ed comb<strong>in</strong><strong>in</strong>g,<br />

the representation of Dirichlet forms <strong>in</strong> terms of derivations, with the basic<br />

results of representation theory of commutative C ∗ -algebras. Compar<strong>in</strong>g the<br />

BDLJ decomposition with the one of Proposition 5.9, it is not difficult to<br />

have the follow<strong>in</strong>g identifications<br />

Ecu = E (c) , Eab = E (j) + E k .<br />

This result may suggests to <strong>in</strong>terpret the completely unbounded part of<br />

Dirichlet form E on a C ∗ -algebra A, as the strongly local part of E.<br />

5.7 Positive Curvature Operator and Markovianity<br />

of the Dirac-Dirichlet Form<br />

In this section we collect the ma<strong>in</strong> po<strong>in</strong>ts <strong>in</strong> the proof of Theorem 5.1. The<br />

equivalence among items ii) and iii) is just the general correspondence between<br />

noncommutative Dirichlet form and tracially symmetric Markovian<br />

semigroups. Property ii) follows from i) by the results of Section 2.2, 2.3,<br />

specialized and simplified to the tracial sett<strong>in</strong>g. The reverse implication is a<br />

matter of elliptic regularity, i.e. Sobolev estimates, and it is a consequence of<br />

the regularization properties of the heat semigroups of Feller type.<br />

To show that iii) is a consequence of iv), let us consider the Bochner Identity<br />

<strong>in</strong> the form (5.25). As the curvature operator is assumed to be nonnegative,<br />

by (5.27), quadratic from QR is cont<strong>in</strong>uous superposition of Dirichlet forms<br />

associated to bounded derivations, i.e. the commutators appear<strong>in</strong>g <strong>in</strong> (5.28).<br />

Hence, the quadratic form QR itself is a Dirichlet form. S<strong>in</strong>ce, by Theorem<br />

5.6, the quadratic form of the Bochner-Laplacian EB is a Dirichlet form too,<br />

so is their sum ED.<br />

The proof of the fact that property iv) follows from property iii) is based,<br />

essentially, on the decomposition of Dirichlet forms <strong>in</strong>troduced <strong>in</strong> 5.6.<br />

Suppose that ED is a Dirichlet form. Write the curvature operator as the<br />

difference of its positive and negative parts: � R = � R+ − � R−. Correspond<strong>in</strong>gly<br />

we have the decomposition Q � R = Q � R+ − Q � R− ,wheretheformsQ � R± are<br />

def<strong>in</strong>ed analogously to Q � R .<br />

Consider now the follow<strong>in</strong>g decomposition of the Dirac form:<br />

ED = E+ −E− , (5.29)


258 F. Cipriani<br />

where<br />

E+ :=EB + Q � R+<br />

E− :=Q � R− .<br />

(5.30)<br />

Obviously we then have that E+ = ED + E− appears as a sum of Dirichlet<br />

forms, and we may apply the analysis developed <strong>in</strong> Section 5.6. Clearly<br />

On the other hand<br />

By identification we have<br />

(E+)cu =(EB + QR+ � )cu = EB ,<br />

(E+)ab =(EB + QR+ � )ab = QR+ � .<br />

(ED + E−)cu =(ED + Q � R− )cu =(ED)cu ,<br />

(ED + E−)ab =(ED + Q � R− )ab =(ED)ab + Q � R−<br />

Q � R+ =(E+)ab =(ED + Q � R− )ab =(ED)j + Q � R−<br />

(5.31)<br />

(5.32)<br />

(5.33)<br />

so that the quadratic form associated to the curvature operator can be identified<br />

with the jump<strong>in</strong>g part (i.e. approximately bounded part) of the Dirac<br />

form:<br />

Q � R = Q � R+ − Q � R− =(ED)ab . (5.34)<br />

This proves, <strong>in</strong> particular, that Q � R is a C ∗ –Dirichlet form. By a purely algebraic<br />

result, concern<strong>in</strong>g specific properties of Clifford algebras over f<strong>in</strong>ite<br />

dimensional Euclidean spaces, one obta<strong>in</strong>s that the coefficients µα <strong>in</strong> (5.28)<br />

have to be nonnegative, so that the curvature operator has to be nonnegative<br />

too.<br />

We conclude this section with an application of the above ma<strong>in</strong> result to<br />

the topology of Riemannian spaces with nonnegative curvature operator.<br />

Proposition 5.12. Let M be a compact Riemannian manifold with nonnegative<br />

curvature operator � R ≥ 0. Then the space<br />

kerD<br />

of harmonic sp<strong>in</strong>ors is a f<strong>in</strong>ite dimensional C ∗ -algebra hence a f<strong>in</strong>ite sum of<br />

full matrix algebras<br />

kerD = Mn1(C) ⊕ Mn2(C) ⊕···⊕Mnk (C) ,<br />

where the <strong>in</strong>teger k ∈ N ∗ is the number of maximal ideals of kerD. Moreover,<br />

the evaluation maps<br />

ex :kerD → Cl(TxM) ex(σ) =σ(x) x ∈ M


Dirichlet Forms on Noncommutative Spaces 259<br />

are <strong>in</strong>jective morphisms of C ∗ -algebras. In particular the sum of the Betti<br />

numbers bi(M) of M is a sum of squares<br />

dimM �<br />

i=1<br />

bi(M) =<br />

and if M is irreducible and dimM is even this number co<strong>in</strong>cides with Euler’s<br />

characteristic χ(M) of M.<br />

Proof. By elliptic regularity, kerD is a subspace of the Clifford algebra (<strong>in</strong><br />

fact its elements are smooth sections of the Clifford bundle). Comb<strong>in</strong><strong>in</strong>g<br />

Theorem 5.1 and Theorem 4.18 with the observation that the kernel of a<br />

derivation on a C ∗ -algebra is a C ∗ -subalgebra, we get that kerD is a f<strong>in</strong>ite<br />

dimensional C ∗ -subalgebra of C ∗ 0(M). The structure theorem of f<strong>in</strong>ite dimensional<br />

C ∗ -algebras (see [T2]) allows to conclude that kerD is f<strong>in</strong>ite sum of<br />

full matrix algebras. Moreover, s<strong>in</strong>ce the curvature operator is nonnegative,<br />

by the Bochner’s identity (5.25), harmonic sp<strong>in</strong>ors are parallel (vanish<strong>in</strong>g covariant<br />

derivative). Hence for any fixed po<strong>in</strong>t x ∈ M, the evaluation map<br />

kerD ∋ σ ↦→ σ(x) ∈ Cl(TxM) is an <strong>in</strong>jective morphism of C ∗ -algebras.<br />

The statement concern<strong>in</strong>g the sum of the Betti numbers follows observ<strong>in</strong>g<br />

that kerD is isomorphic to the space of harmonic forms of M and<strong>in</strong>turn,<br />

by the de Rham’s Theorem (see [LM] for example), to the cohomology of M.<br />

The last statement follows because on irreducible manifolds of even dimension<br />

hav<strong>in</strong>g nonnegative curvature operator, the odd Betti numbers vanish (see<br />

[Pet page 241]).<br />

Remark 5.13. If a Riemannian manifold whose curvature operator is nonnegative,<br />

admits a non constant harmonic form, then the heat semigroup e −tD2<br />

generated by the Dirac Laplacian is nonergodic.<br />

6 Dirichlet Forms and Noncommutative Geometry<br />

In the previous chapter we observed that on Riemannian manifolds there<br />

exist noncommutative Dirichlet forms and associated Markovian semigroups<br />

reflect<strong>in</strong>g geometrical or topological properties of the space.<br />

In that sett<strong>in</strong>g, the powerful tools of differential and <strong>in</strong>tegral calculus, as well<br />

as tensor algebra, allow to construct topological <strong>in</strong>variants, as for example<br />

the de Rham’s currents describ<strong>in</strong>g the de Rham homology of X <strong>in</strong> terms of<br />

curvatures. Recall, as a prototype, the case of a Riemann surfaces Σ, whose<br />

Euler characteristic may be computed as the <strong>in</strong>tegral of the Gauss curvature<br />

of a Riemannian metric on Σ (see Example 5.2).<br />

In this chapter we collect some result to corroborate the idea that Dirichlet<br />

spaces, commutative or not, exibits a geometry. The idea is to exploit the<br />

k�<br />

l=1<br />

n 2 l


260 F. Cipriani<br />

first order differential calculus, constructed <strong>in</strong> Chapter 4, to develop tools<br />

of differential topology and cyclic cohomology. We limit the presentation to<br />

commutative C ∗ -algebras.<br />

6.1 Dirichlet Spaces as Banach Algebras<br />

In this section (X, m) will denote a metrizable, compact, Hausdorff space,<br />

endowed with a positive Radon measure of full topological support. We will<br />

consider a regular Dirichlet form (E, F) onL2 (X, m) and denote by B the<br />

associated Dirichlet algebra C(X) ∩F (see Proposition 4.7). The first h<strong>in</strong>t<br />

that a Dirichlet structure on a topological space X may carriy topological<br />

<strong>in</strong>formation comes from the follow<strong>in</strong>g result (see [Cip4]).<br />

Proposition 6.1. (The Dirichlet algebra is semisimple) The Dirichlet algebra<br />

B := C(X) ∩F,normedby<br />

�a� := �a�∞ + � E[a] a ∈B,<br />

is an <strong>in</strong>volutive, semisimple Banach algebra. It is unital if and only if E[1] = 0<br />

(i.e. the Dirichlet space is conservative). As a consequence, the Dirichlet algebra<br />

B has a unique Banach algebra topology: any norm, mak<strong>in</strong>g B a Banach<br />

algebra, is equivalent to the above one.<br />

Proof. By represent<strong>in</strong>g E by a derivation, from the Leibniz rule one obta<strong>in</strong>s<br />

for all a, b ∈B:<br />

� �<br />

E[ab] ≤ E[a] ·�b�∞ + �a�∞ · � E[b] a, b ∈B (6.1)<br />

so that<br />

�ab� = �ab�∞ + � E[ab] ≤�a�∞�b�∞ + � E[a] ·�b�∞ + �a�∞ · � E[b]<br />

≤ (|a�∞ + � E[a])(|b�∞ + � E[b])<br />

= �a��b� .<br />

Clearly �1� =1ifandonlyifE[1]=0sothatwegetthatB is an <strong>in</strong>volutive<br />

normed algebra which is complete s<strong>in</strong>ce the Dirichlet form is closed. Denot<strong>in</strong>g<br />

by ρ(a) = limn→∞ �a n � 1/n the spectral radius of an element a of B, weget<br />

immediately that �a�∞ ≤ ρ(a) s<strong>in</strong>ce�a�∞ ≤�a� . To prove the converse<br />

notice that from (6.1) one has � E[a n ] ≤ n�a� n−1<br />

∞ · � E[a] for all n ≥ 1. Then<br />

we have<br />

which implies<br />

�a n �∞ = �a n �∞ + � E[a n ] ≤�a� n ∞ + n�a� n−1<br />

∞ · � E[a]<br />

ρ(a) ≤ lim<br />

n→∞ (�a�n∞ + n�a�n−1 ∞ · � E[a]) 1/n = �a�∞ .


Dirichlet Forms on Noncommutative Spaces 261<br />

Hence the spectral radius ρ(a) of an element of the Dirichlet algebra B co<strong>in</strong>cides<br />

with its uniform norm �a�∞, which is the spectral radius of the semisimple<br />

algebra C(X). This proves that B is semisimple. The uniqueness of the<br />

norm topology of semisimple Banach algebras is a known result (see [Ric]).<br />

The above result still holds true for locally compact spaces X. To carry over<br />

the proof one needs to consider the so called the extended Dirichlet space of<br />

E (see [FOT], [Cip4]).<br />

A first consequence of the uniqueness above is the fact that strongly local<br />

Dirichlet forms are determ<strong>in</strong>ed by their Dirichlet algebras. Theresultmay<br />

be considered as a generalization, to Dirichlet spaces, of the fact that, on<br />

a differerentiable, paracompact manifold, Riemannan metrics giv<strong>in</strong>g rise to<br />

Dirichlet <strong>in</strong>tegrals hav<strong>in</strong>g the same doma<strong>in</strong> (the first order Sobolev space)<br />

are quasi-conformally equivalent.<br />

Theorem 6.2. (Quasi-conformal equivalence of Dirichlet forms hav<strong>in</strong>g the<br />

same doma<strong>in</strong>) Let (E1, F1) and (E2, F2) be two strongly local, regular Dirichlet<br />

spaces on L 2 (X, m) hav<strong>in</strong>g common Dirichlet algebra B = C(X) ∩F1 =<br />

C(X) ∩F2.<br />

Then F1 = F2 and the Dirichlet forms are quasi-equivalent <strong>in</strong> the sense that<br />

there exists k>0 such that<br />

1<br />

k E1[a] ≤E2[a] ≤ k E1[a] a ∈F1 = F2 . (6.2)<br />

Proof. By Corollary 6.2 we have, for some k>0<br />

1<br />

k (�a�∞ + � E1[a]) ≤�a�∞ + E2[a] ≤ k (�a�∞ + � E1[a]) a ∈F1 = F2 .<br />

(6.3)<br />

Represent<strong>in</strong>g the forms by some derivations ∂1 and ∂2 and apply<strong>in</strong>g the<br />

cha<strong>in</strong> rule, one obta<strong>in</strong>s the identity<br />

Ei[φλ(a)] + Ei[ψλ(a)] = Ei[a] a ∈B<br />

where φλ(t) :=λ−1 (1 − s<strong>in</strong>(λt)), ψλ(t) :=λ−1 (1 − cos(λt)) for all t ∈ R and<br />

λ �= 0. S<strong>in</strong>ce limλ→0 �φλ(a)�∞ = limλ→0 �ψλ(a)�∞ = 0 we observe that (6.2)<br />

follows from (6.3)<br />

The second consequence we draw from the norm uniqueness result concerns<br />

the differential topology of a Dirichlet space. The fact that the spectral<br />

radius <strong>in</strong> the Dirichlet algebra co<strong>in</strong>cides with the uniform norm has another<br />

important consequence: an element a ∈Bis <strong>in</strong>vertible <strong>in</strong> B if and only if it is<br />

<strong>in</strong>vertible <strong>in</strong> C(X). This implies a generalization to regular Dirichlet spaces<br />

of a well known result of H. Whitney <strong>in</strong> differential topology accord<strong>in</strong>g to<br />

which f<strong>in</strong>ite-dimensional, locally trivial, vector bundles over a smooth manifold<br />

admit a smooth structure. The <strong>in</strong>terested reader may f<strong>in</strong>d the proof <strong>in</strong><br />

[Cip4].


262 F. Cipriani<br />

Theorem 6.3. (K-<strong>Theory</strong> and vector bundles over regular Dirichlet spaces)<br />

Let (E, F) be a regular Dirichlet form on L2 (X, m). Then we have<br />

i) the K-theory K∗(B) of the Dirichlet algebra B co<strong>in</strong>cides with the Ktheory<br />

K∗(C(X)) of the algebra C(X) and <strong>in</strong> turn with the topological<br />

K-theory K∗ (X) of the topological space X;<br />

ii) every f<strong>in</strong>ite-dimensional locally trivial vector bundle E over X admits a<br />

Dirichlet structure, namely, a compatible vector bundle atlas whose transition<br />

matrices have entries <strong>in</strong> the Dirichlet algebra B.<br />

6.2 Fredholm Module of a Dirichlet Space<br />

In this section we show how to use the differential calculus associated to<br />

Dirichlet form, to construct Fredholm modules, <strong>in</strong> the sense of Atiyah [At],<br />

on Dirichlet spaces. Notice that Fredholm modules and their summability<br />

properties lie at the core of Noncommutative Geometry of A. Connes [Co2].<br />

The construction of Fredholm modules (F, H) on compact topological spaces<br />

X is a generalization of the theory of elliptic differential operators on compact<br />

manifolds [At]. In its (odd) form, one requires that the elements f of the<br />

algebra of cont<strong>in</strong>uous functions C(X) are represented as bounded operators<br />

π(f) on a Hilbert space H on which there is a dist<strong>in</strong>guished self-adjo<strong>in</strong>t<br />

operator F of square one (F 2 = 1), the symmetry, such that the commutators<br />

[F, π(f)] are compact operators.<br />

A nice example of an (even) Fredholm module on an even dimensional<br />

oriented conformalmanifoldV hasbeenconstructedbyA.Connes-D.Sullivan- N.<br />

Teleman <strong>in</strong> [CST] (see also [Co2 Chapter IV, 4-α Theorem 2]). The Fredholm<br />

module is uniquely determ<strong>in</strong>ed by the oriented conformal structure, which<br />

can be characterized for example by the Ln-norm on 1-forms on V ,where<br />

n := dimV . In the particular case where V is an oriented Riemann surface,<br />

the conformal structure is uniquely determ<strong>in</strong>ed by the Dirichlet <strong>in</strong>tegral<br />

�<br />

E[a] = |∇a| 2 dm<br />

V<br />

on the algebra of cont<strong>in</strong>uous functions over V . We are go<strong>in</strong>g to see that the<br />

above construction can be carried out on general Dirichlet spaces.<br />

Def<strong>in</strong>ition 6.4. (Phase operator of a Dirichlet form) Let us consider a regular<br />

Dirichlet form (E, F) onL 2 (X, m). Let ∂ : B→Hbe the associated<br />

derivation, def<strong>in</strong>ed on the Dirichlet algebra B = C(X) ∩F with values <strong>in</strong> the<br />

symmetric Hilbert module H.<br />

Let P ∈ Proj (H) be the projection onto the closure Im∂ of the range of the<br />

derivation ∂,


Dirichlet Forms on Noncommutative Spaces 263<br />

P H = Im∂ (6.4)<br />

We will call F = P − P ⊥ : H→Hthe phase operator associated to the<br />

regular Dirichlet space.<br />

Theorem 6.5. (Fredholm module of a Dirichlet form) Let (E, F) be a regular<br />

Dirichlet form on L2 (X, m), such that<br />

i) the spectrum {0


264 F. Cipriani<br />

where G(x, y) = � ∞<br />

k ak(x)ak(y) istheGreen function or kernel of the<br />

k=1 λ−1<br />

compact operator H −1<br />

D on L2 (X, m) determ<strong>in</strong>ed by E. S<strong>in</strong>ce, by assumption,<br />

the Green function is bounded, we have<br />

�[F, a]� 2 � �<br />

L2 ≤ 8 sup G(x, x) E[a] < +∞<br />

x∈X<br />

for all a ∈B⊂C(X). S<strong>in</strong>ce the form is regular, B is uniformly dense <strong>in</strong> C(X),<br />

[F, a] is norm cont<strong>in</strong>uous with respect to a ∈ C(X) and the space of compact<br />

operators is norm closed, we have that [F, a] iscompactforalla ∈ C(X).<br />

Even if the boundedness assumption on the Green function restricts the application<br />

of the above construction to situations <strong>in</strong> which the spectral dimension<br />

is low, it applies, however, to the class of Dirichlet forms constructed by<br />

J. Kigami [Ki] on self-similar fractal spaces, associated to regular harmonic<br />

structures (see [CS4]).<br />

Remark 6.6. The above result <strong>in</strong>dicates a direct connection between the summability<br />

properties of the Fredholm module and two of the ma<strong>in</strong> objects of<br />

potential theory, namely, the Dirichlet form E and the Green function G.<br />

An immediate consequence of the above construction is the follow<strong>in</strong>g <strong>in</strong>dex<br />

theorem [At], [Co2 Chapter IV Proposition 2].<br />

Corollary 6.7. Let (Hq ,Fq) be the Fredholm module over Mq(C(X)) =<br />

C(X) ⊗ Mq(C) given by<br />

Hq := H⊗C q , Fq := F ⊗ 1 , πq := π ⊗ id , q ∈ N ∗ .<br />

Def<strong>in</strong>e, on the Hilbert space Hq, the projection operator<br />

Eq := 1+Fq<br />

.<br />

2<br />

Then for every <strong>in</strong>vertible element u ∈ GLq(A) <strong>in</strong> Mq(A), the operator<br />

Eqπq(u)Eq : EqHq −→ EqHq<br />

is a a Fredholm operator. Its <strong>in</strong>dex gives rise to an additive map<br />

φ : K1(A) −→ Z φ([u]) := Index(Eqπq(u)Eq)<br />

on the K-theory group K1(C(X)) = K 1 (X) of C(X).<br />

Notice that, the Fredholm modules (Hq ,Fq) are those associated by<br />

Theorem 6.5 to the family of Dirichlet forms on Mq(C(X)) determ<strong>in</strong>ed by E<br />

(see Def<strong>in</strong>ition 4.1).<br />

We conclude this section notic<strong>in</strong>g that the Fredholm module (H ,F)may<br />

determ<strong>in</strong>e topological <strong>in</strong>variants of X (see [Co2 Chapter IV]).


Dirichlet Forms on Noncommutative Spaces 265<br />

In fact, as the Steenrod K-homology K∗(X) of the compact space X is isomorphic<br />

to the K-homology of K ∗ (C(X)) of the algebra C(X), a specific element<br />

of it is determ<strong>in</strong>ed by the equivalence class of the Fredholm module (H ,F).<br />

Moreover, the Chern character,ch∗(H ,F), of the f<strong>in</strong>itely summable Fredholm<br />

module (H ,F)istheperiodic cyclic cohomology class of the follow<strong>in</strong>g cyclic<br />

cocycles for n ≥ 2,<br />

A ⊗···⊗A ∋ (a 0 ,a 1 ,... ,a n ) ↦→ λnTr ′ (a 0 [F, a 1 ] ...[F, a n ]) ,<br />

where λn := √ 2i(−1) n(n+1)<br />

2 Γ ( n<br />

2 +1)andTr′ (B) := 1<br />

2<br />

7 Appendix<br />

7.1 Topologies on Operator Algebras<br />

Tr(F (FB + BF)).<br />

In this subsection we briefly recall some basic def<strong>in</strong>itions concern<strong>in</strong>g operator<br />

algebras and collect the relationships among their topologies.<br />

AC∗-algebra A is an <strong>in</strong>volutive Banach algebra <strong>in</strong> which the norm and the<br />

<strong>in</strong>volution are related by<br />

�a ∗ a� = �a� 2<br />

a ∈ A.<br />

The algebra C0(X) of cont<strong>in</strong>uous functions, vanish<strong>in</strong>g at <strong>in</strong>f<strong>in</strong>ity on a locally<br />

compact Hausdorff space X, endowed with the uniform norm and the<br />

<strong>in</strong>volution given by the po<strong>in</strong>twise complex conjugation, is a commutative<br />

C ∗ -algebra. A celebrated theorem of I. M. Gelfand (see [Dix1]) ensures that<br />

all commutative C ∗ -algebras are of this form for a suitable space X (to be<br />

identified with the space of all closed maximal ideals of A).<br />

The simplest examples of noncommutative C ∗ -algebras are the full matrix<br />

algebras Mn(C), n ≥ 2. F<strong>in</strong>ite dimensional C ∗ -algebras are direct sums of<br />

full matrix algebras. Examples of <strong>in</strong>f<strong>in</strong>ite dimensional, noncommutative C ∗ -<br />

algebras are the algebra k(H) of all compact operators on an <strong>in</strong>f<strong>in</strong>ite dimensional<br />

Hilbert space H and the algebra B(H) of all bounded operators. In<br />

all these cases the norm is the usual one and the <strong>in</strong>volution is given by the<br />

operation of tak<strong>in</strong>g the operator adjo<strong>in</strong>t. By another fundamental theorem of<br />

I. M. Gelfand (see [Dix1]), any C ∗ -algebra A may be represented as a norm<br />

closed C ∗ -subalgebra of B(H). In particular this implies that the norm <strong>in</strong> a<br />

C ∗ -algebra is uniquely determ<strong>in</strong>ed by the algebraic structure.<br />

Among the C ∗ -algebras, the von Neumann ones M may be def<strong>in</strong>ed as those<br />

which, as Banach spaces, admit preduals <strong>in</strong> the sense that M =(M∗) ∗ for<br />

some Banach space M∗.


266 F. Cipriani<br />

The algebra L ∞ (X, m) of essentially bounded measurable functions on a<br />

measured space (X, m) is a commutative von Neumann algebra and all commutative<br />

von Neumann algebras arise <strong>in</strong> this way.<br />

The algebra B(H) of all bounded operators on a Hilbert space H is a von<br />

Neumann algebra (the type I∞ factor) be<strong>in</strong>g the dual of the Banach space<br />

of all trace class operators on H.<br />

By a fundamental theorem of J. von Neumann (see [Dix2]), for any given<br />

selfadjo<strong>in</strong>t subset S ⊂ B(H) theset<br />

S ′ := {x ∈ B(H) :xy = yx, y ∈ S}<br />

of all bounded operators on H which commute with all operators of S, isa<br />

von Neumann subalgebra of B(H) and moreover any von Neumann algebra<br />

M ⊆ B(H) co<strong>in</strong>cides with its own double commutant<br />

M = M ′′ .<br />

Von Neumann algebras M ⊆ B(H) whose <strong>in</strong>tersection with its own commutant<br />

M ′ , reduces to the one dimensional algebra of scalar multiples of the<br />

identity operator<br />

M ∩ M ′ = C · 1H ,<br />

are called factors and may be considered <strong>in</strong> a precise sense as build<strong>in</strong>g blocks<br />

for general von Neumann algebras.<br />

On a C ∗ -algebra A, one may consider the norm topology, also referred as the<br />

uniform topology and denoted by τu, aswellastheweak topology σ(A,A ∗ ),<br />

<strong>in</strong>duced by its dual space A ∗ .<br />

As von Neumann algebras M are characterized as C∗-algebras which as Banach<br />

spaces, admit preduals M∗ (<strong>in</strong> the sense that M =(M∗) ∗ )besidethe<br />

two topologies above, they carry the weak∗-topology σ(M , M∗) too.<br />

Let M act on the Hilbert space H so that M ⊆ B(H). Other useful topologies<br />

on M, are then the follow<strong>in</strong>g:<br />

• the strong topology is the locally convex topology def<strong>in</strong>ed by the family of<br />

sem<strong>in</strong>orms<br />

M ∋ a →�aξ� ,<br />

parametrized by the vectors ξ ∈H;<br />

• the strong∗ topology is the locally convex topology def<strong>in</strong>ed by the family<br />

of sem<strong>in</strong>orms<br />

M ∋ a →�aξ� + �a ∗ ξ� ,<br />

parameterized by the vectors ξ ∈H;<br />

• σ-strong topology is the locally convex topology def<strong>in</strong>ed by the family of<br />

sem<strong>in</strong>orms


Dirichlet Forms on Noncommutative Spaces 267<br />

��<br />

M ∋ a → �aξn� 2� 1/2<br />

,<br />

parametrized by sequences {ξn} ⊂Hsuch that �<br />

n �ξ�2 < +∞;<br />

• σ-strong ∗ topology is the locally convex topology def<strong>in</strong>ed by the family of<br />

sem<strong>in</strong>orms<br />

n<br />

n<br />

��<br />

M ∋ a → �aξn� 2 + �<br />

�a ∗ ξn� 2� 1/2<br />

,<br />

parametrized by sequences {ξn} ⊂Hsuch that �<br />

n �ξ�2 < +∞;<br />

• σ-weak topology is the locally convex topology def<strong>in</strong>ed by the family of<br />

sem<strong>in</strong>orms<br />

M ∋ a → �<br />

|(ξn|aηn)| ,<br />

n<br />

parametrized by sequences {ξn} , {ηn} ⊂Hsuch that �<br />

n �ξ�2 �<br />

< +∞,<br />

n �η�2 < +∞.<br />

The σ-weak topology co<strong>in</strong>cides with the weak∗-topology σ(M, M∗) (see[BR1<br />

Proposition 2.4.3]), so that bounded sets are relatively compact by the<br />

Banach-Alaoglu Theorem. The relation among the various topologies are subsummed<br />

as follows:<br />

uniform


268 F. Cipriani<br />

Φt ◦ Φs = Φt+s and Φ0 = IA , s,t ≥ 0 .<br />

In case �Φt(a)� = �a�, for all a ∈ A and t ∈ R or t ≥ 0, Φ ia called a group or<br />

semigroup of isometries. If,however,�Φt(a)� ≤�a�, for all a ∈ A and t ∈ R<br />

or t ≥ 0, Φ is called a contraction group or semigroup.<br />

A one-parameter group or semigroup on a C∗-algebra A is said to be<br />

• norm or uniformly cont<strong>in</strong>uous if<br />

• strongly cont<strong>in</strong>uous if<br />

lim<br />

t→0 �Φt − IA� B(A) =0;<br />

lim<br />

t→0 �Φt(a) − a�A =0 ∀ a ∈ A ;<br />

• weakly cont<strong>in</strong>uous if<br />

lim<br />

t→0 〈η | Φt(a) − a〉 =0 ∀a ∈ A, ∀ η ∈ A ∗ .<br />

A one-parameter group or semigroup on a von Neumann algebra M is said<br />

• weakly∗ cont<strong>in</strong>uous if the maps Φt : M → M, areweakly∗cont<strong>in</strong>uous and<br />

lim<br />

t→0 〈η | Φt(a) − a〉 =0 ∀a ∈ M , ∀ η ∈ M∗ .<br />

By a general result, see [BR1 Corollary 3.1.8, Proposition 3.1.3], strong and<br />

weak cont<strong>in</strong>uity of semigroups are equivalent properties, and both of them<br />

imply the bound<br />

�Φt� ≤M · e βt<br />

t ≥ 0 .<br />

for suitable constants M>0andβ∈R. Strongly cont<strong>in</strong>uous semigroups are also called C0-semigroups, weakly∗ cont<strong>in</strong>uous<br />

semigroups are also called C∗ 0 or σ(M, M∗)-cont<strong>in</strong>uous semigroups<br />

and weakly cont<strong>in</strong>uous semigroups are also called σ(A, A∗ )-cont<strong>in</strong>uous semigroups.<br />

The generator of a strongly (resp. weakly) cont<strong>in</strong>uous semigroup {Φt : t ≥ 0}<br />

on a C∗-algebra A, is the operator (L,D(L)) on A def<strong>in</strong>ed as follows: an<br />

element a ∈ A is <strong>in</strong> the doma<strong>in</strong> D(L) if the limit<br />

a − Φt(a)<br />

lim<br />

t→0 t<br />

exists <strong>in</strong> the strong (resp. weak) topology. The action L(a) of the operator<br />

L on the element a of D(L) is def<strong>in</strong>ed as the limit above. The generator<br />

of a weakly ∗ cont<strong>in</strong>uous semigroup on a von Neumann algebra M is def<strong>in</strong>ed


Dirichlet Forms on Noncommutative Spaces 269<br />

similarly, the only difference be<strong>in</strong>g that the limit above is understood to exists<br />

with respect to the weak∗ topology.<br />

It can be verified that the generator (L,D(L)) of a strongly (resp. weakly)<br />

cont<strong>in</strong>uous semigroup is a norm (resp. weakly) densely def<strong>in</strong>ed operator and<br />

a norm-norm (resp. σ(A, A∗ )-σ(A, A∗ )) closed operator on a C∗-algebra A.<br />

Similarly, the generator of a weakly∗ cont<strong>in</strong>uous semigroup, is a weakly∗ densely def<strong>in</strong>ed σ(M, M∗)-σ(M, M∗) closedoperatoronthevonNeumann<br />

algebra M. See [BR1 Proposition 3.1.6]. Moreover, the semigroup is uniformly<br />

cont<strong>in</strong>uous if and only if its generator is a bounded operator. In this case,<br />

one has the representation<br />

Φt =<br />

∞�<br />

n=0<br />

t n<br />

n! Ln<br />

t ≥ 0 ,<br />

where the exponential series converges <strong>in</strong> the norm of the Banach space of<br />

all bounded operators on A.<br />

Theorem 7.1. (St<strong>in</strong>espr<strong>in</strong>g Theorem [Sti]). A l<strong>in</strong>ear map Φ : A → B(h),<br />

from a unital C ∗ -algebra A, <strong>in</strong>totheC ∗ -algebra of all bounded operators on<br />

aHilbertspaceH, is completely positive if and only if it has the form<br />

Φ(a) =V ∗ π(a)V a ∈ A, (7.1)<br />

for some representation π : A → B(K) and some bounded l<strong>in</strong>ear operator<br />

V : H→K.IfA and H are separable, K can be taken to be separable. If A is<br />

a von Neumann algebra and Φ is normal, then π can be taken to be normal.<br />

7.3 Analytic Elements<br />

Let α:={αt : t ∈ R} be a cont<strong>in</strong>uous group of isometries of a C ∗ -algebra A.<br />

An element a ∈ A is said to be α-analytic if there exists a strip Iλ := {z ∈<br />

C : |Imz| 0) and an analytic function f : Iλ → A such that<br />

f(t) =αt(a) t ∈ R .<br />

It may be shown, see [BR1 Chapter 3.1], that an element a ∈ A is α-analytic<br />

if and only if it is analytic for the generator (L, D(L)), <strong>in</strong> the sense that<br />

a ∈ D(L n ) for all n ≥ 1and<br />

∞�<br />

n=0<br />

t n<br />

n! �Ln (a)� < +∞ t ≥ 0 .<br />

The set of entire analytic elements Aα is a norm dense, <strong>in</strong>variant ∗ -subalgebra<br />

of A, <strong>in</strong> the case of a strongly cont<strong>in</strong>uous group of isometries of a C ∗ -algebra


270 F. Cipriani<br />

A and, a σ(M, M∗)-dense, <strong>in</strong>variant ∗ -subalgebra of M, <strong>in</strong>thecaseofa<br />

σ(M, M∗)-cont<strong>in</strong>uous group of isometries of a von Neumann algebra M.<br />

In both cases, these extensions form a group of automorphisms of the algebra<br />

Aα, denotedby{αz : z ∈ C} and <strong>in</strong>dexed by complex numbers.<br />

In particular, if {σ ω t : t ∈ R} is the modular automorphisms group of a von<br />

Neumann algebra M, associated to a normal state ω, andx ∈ Mσ is an entire<br />

analytic element, one has the identity<br />

σ ω z (x) =∆z ω x∆−z<br />

ω , x ∈ Mα z ∈ C ,<br />

where ∆ω denotes the modular operator of ω. Both sides of the above identity<br />

are understood as bounded quadratic forms<br />

(ξ| σ ω z<br />

(x)η) =(∆−z ω ξ| x∆−z ω η) , x ∈ Mα z ∈ C ,<br />

on the dense subspace of H of entire analytic elements of the unitary group<br />

R ∋ t → ∆it ω on H. Infact,forfixedx∈Mα, both sides are entire analytic<br />

functions, co<strong>in</strong>cid<strong>in</strong>g for real z. In this way, one may also show that<br />

σ ω t+is(x) =σ ω t (σ ω is(x)) = σ ω is(σ ω t (x)) t + is ∈ Iλ ,<br />

for all x ∈ M such that the map R ∋ t → σω t (x) ∈ M admits an analytic<br />

extension to a doma<strong>in</strong> conta<strong>in</strong><strong>in</strong>g the strip Iλ.<br />

8 List of Examples<br />

• Self-polarity <strong>in</strong> classical measure theory<br />

• Jordan decomposition <strong>in</strong> commutative<br />

Example 2.1<br />

von Neumann algebras<br />

• Jordan decomposition <strong>in</strong> type I and type II<br />

Example 2.4<br />

von Neumann algebras Example 2.5<br />

• Modular automorphisms groups on type I factors Example 2.9<br />

• Modular automorphisms groups on Power’s factors Example 2.10<br />

• Standard positive cone of f<strong>in</strong>ite von Neumann algebras Example 2.12<br />

• Hilbert-Schmidt standard form on type I factors Example 2.17<br />

• Generalized Hilbert-Schmidt standard form Example 2.18<br />

• Modular operators on type I factors<br />

• Standard forms of group von Neumann algebras<br />

Example 2.19<br />

and Fourier transform Example 2.20<br />

• Symmetric embedd<strong>in</strong>g <strong>in</strong> type I factors Example 2.23


Dirichlet Forms on Noncommutative Spaces 271<br />

• Structure of norm cont<strong>in</strong>uous,<br />

completely positive semigroups Example 2.26<br />

• KMS-states on CAR C∗-algebras Example 2.29<br />

• KMS-states on CCR C∗-algebras • KMS-symmetric, norm cont<strong>in</strong>uous,<br />

Example 2.30<br />

completely Markovian semigroups<br />

• Markovianity of a Schröd<strong>in</strong>ger semigroups<br />

Example 2.32<br />

with respect to a ground state<br />

• Stability of Markovianity under<br />

Example 2.42<br />

subord<strong>in</strong>ation of semigroups<br />

• Faces of positive cones of commutative<br />

Example 2.43<br />

von Neumann algebras Example 2.47<br />

• Bounded Dirichlet forms Example 2.56<br />

• An unbounded Dirichlet form<br />

• Quantum Ornste<strong>in</strong>-Uhlenbeck semigroup,<br />

Example 2.57<br />

Quantum Brownian motion, Birth-and-Death process Example 2.58<br />

• Dirichlet forms and quasi-free states on CAR algebras Example 2.61<br />

• Dirichlet forms and quasi-free states on CCR algebras Example 2.62<br />

• Is<strong>in</strong>g and Heisenberg models Example 3.2<br />

• Gradient and square <strong>in</strong>tegrable vector fields.<br />

• Derivations associated to mart<strong>in</strong>gales<br />

Example 4.3<br />

and stochastic calculus<br />

• Dirichlet forms on commutative C<br />

Example 4.4<br />

∗-algebras: Beurl<strong>in</strong>g-Deny-Le Jan decomposition revisited I<br />

• Dirichlet forms on group C<br />

Example 4.19<br />

∗-algebras associated to functions of negative type Example 4.20<br />

• Clifford Dirichlet form of free fermion systems Example 4.21<br />

• Heat semigroup on noncommutative torus<br />

• Noncommutative Dirichlet forms<br />

Example 4.22<br />

and topology of Riemann surfaces Example 5.2<br />

• Dirichlet forms on hypersurfaces Example 5.3<br />

• Dirichlet forms and topology of Cartan’s symmetric spaces<br />

• Structure of bounded completely Dirichlet forms<br />

Example 5.4<br />

on amenable C∗-algebras • Dirichlet forms on commutative C<br />

Example 5.9<br />

∗-algebras: Beurl<strong>in</strong>g-Deny-Le Jan decomposition revisited II Example 5.11


272 F. Cipriani<br />

References<br />

[Abk] W. Abikoff, The uniformization theorem, Am. Math. Montly. 88, n. 8, (1981),<br />

574-592.<br />

[AC] L. Accardi, C. Cecch<strong>in</strong>i, Conditional expectations <strong>in</strong> von Neuman algebras and<br />

a theorem of Takesaki J. Funct. Anal. 45 (1982), 245-273.<br />

[AFLe] L. Accardi, A. Frigerio, J.T. Lewis, Quantum stochastic processes, Publ. Res.<br />

Inst. Math. Sci. 18, n. 1, (1982), 97–133.<br />

[AFLu] L. Accardi, A. Frigerio, Y.G. Lu, The weak coupl<strong>in</strong>g limit as a quantum functional<br />

central limit., Comm. Math. Phys. 131, n. 3, (1990), 537570.<br />

[AHK1] S. Albeverio, R. Hoegh-Krohn, Dirichlet forms and Markovian semigroups on<br />

C ∗ –algebras, Comm. Math. Phys. 56 (1977), 173-187.<br />

[AHK2] S. Albeverio, R. Hoegh-Krohn, Frobenius theory for positive maps on von Neumann<br />

algebras, Comm. Math. Phys. 64 (1978), 83-94.<br />

[Al] R. Alicki, On the detailed balance condition for non-Hamiltonian systems, Rep.<br />

Math. Phys. 10 (1976), 249-258.<br />

[Ara] H. Araki, Some properties of modular conjugation operator of von Neumann<br />

algebras and a non-commutative Radon-Nikodym theorem with a cha<strong>in</strong> rule,<br />

Pacific J. Math. 50 (1974), 309-354.<br />

[Arv] W. Arveson, “An <strong>in</strong>vitation to C ∗ -algebra”, Graduate Text <strong>in</strong> Mathematics 39,<br />

x + 106 pages, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, Heidelberg, New York, 1976.<br />

[At] M.F. Atiyah, Global theory of elliptic operators, Proc. Internat. Conf. on<br />

Proc. Internat. Conf. on Functional Analysis and Related Topics (Tokyo, 1969)<br />

(1970), 21–30 Univ. of Tokyo Press, Tokyo.<br />

[AtS] M.F. Atiyah, G. B. S<strong>in</strong>ger, The Index of elliptic operators on compact manifolds,<br />

Bull. A.M.S. 69 (1963), 422-433.<br />

[BKP1] C. Bahn, C.K. Ko, Y.M. Park, Dirichlet forms and symmetric Markovian semigroups<br />

on Z2-graded von Neumann algebras Rev. Math. Phys. 15 (2003), no. 8,<br />

823-845.<br />

[BKP2] C. Bahn, C.K. Ko, Y.M. Park, Dirichlet forms and symmetric Markovian semigroups<br />

on CCR Algebras with quasi-free states Rev. Math. Phys. 44, (2003),<br />

723–753.<br />

[B] A. Barchielli, Cont<strong>in</strong>ual measurements <strong>in</strong> quantum mechanics and quantum<br />

stochastic calculus, Lectures Notes Math. 1882 (2006), 207-292.<br />

[BD1] A. Beurl<strong>in</strong>g and J. Deny, Espaces de Dirichlet I: le cas élémentaire, Acta Math.<br />

99 (1958), 203-224.<br />

[BD2] A. Beurl<strong>in</strong>g and J. Deny, Dirichlet spaces, Proc. Nat. Acad. Sci. 45 (1959),<br />

208-215.<br />

[Bi] P. Biane, Logarithmic Sobolev Inequalities, Matrix Models and Free Entropy,<br />

Acta Math. S<strong>in</strong>ica, English Series. 19 (2003), 497-506.<br />

[Boz] M. Bozejko, Positive def<strong>in</strong>ite functions on the free group and the noncommutative<br />

Riesz product, Bollett<strong>in</strong>o U.M.I. 5-A (1986), 13-21.<br />

[BR1] Bratteli O., Rob<strong>in</strong>son D.W., “Operator algebras and Quantum Statistical<br />

Mechanics 1”, Second edition, 505 pages, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, Heidelberg,<br />

New York, 1987.<br />

[BR2] Bratteli O., Rob<strong>in</strong>son D.W., “Operator algebras and Quantum Statistical<br />

Mechanics 2”, Second edition, 518 pages, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, Heidelberg,<br />

New York, 1997.<br />

[Cip1] F. Cipriani, Dirichlet forms and Markovian semigroups on standard forms of von<br />

Neumann algebras, J. Funct. Anal. 147 (1997), 259-300.<br />

[Cip2] F. Cipriani, The variational approach to the Dirichlet problem <strong>in</strong> C ∗ –algebras,<br />

Banach Center Publications 43 (1998), 259-300.<br />

[Cip3] F. Cipriani, Perron theory for positive maps and semigroups on von Neumann algebras,<br />

CMS Conf. Proc., Amer. Math. Soc., Providence, RI 29 (2000), 115-123.


Dirichlet Forms on Noncommutative Spaces 273<br />

[Cip4] F. Cipriani, Dirichlet forms as Banach algebras and applications, Pacific J.<br />

Math. 223 (2006), no. 2, 229-249.<br />

[CFL] F. Cipriani, F. Fagnola, J.M. L<strong>in</strong>dsay, Spectral analysis and Feller property<br />

for quantum Ornste<strong>in</strong>–Uhlenbeck semigroups, Comm. Math. Phys. 210 (2000),<br />

85-105.<br />

[CS1] F. Cipriani, J.-L. Sauvageot, Derivations as square roots of Dirichlet forms, J.<br />

Funct. Anal. 201 (2003), no. 1, 78–120.<br />

[CS2] F. Cipriani, J.-L. Sauvageot, Noncommutative potential theory and the sign of<br />

the curvature operator <strong>in</strong> Riemannian geometry, Geom. Funct. Anal. 13 (2003),<br />

no. 3, 521–545.<br />

[CS3] F. Cipriani, J.-L. Sauvageot, Strong solutions to the Dirichlet problem for differential<br />

forms: a quantum dynamical semigroup approach, Contemp. Math, Amer.<br />

Math. Soc., Providence, RI 335 (2003), 109-117.<br />

[CS4] F. Cipriani, J.-L. Sauvageot, Fredholm modules on p.c.f. self-similar fractals and<br />

their conformal geometry, arXiv:0707.0840v 1 [math.FA] (5 Jul 2007), 16 pages.<br />

[Co1] A. Connes, Caracterisation des espaces vectoriels ordonnés sous-jacents aux<br />

algbres de von Neumann, Ann. Inst. Fourier (Grenoble) 24 (1974), 121-155.<br />

[Co2] A. Connes, “Noncommutative Geometry”, Academic Press, 1994.<br />

[Co3] A. Connes, On the cohomology of operator algebras, J. Funct. Anal. 28 (1978),<br />

no. 2, 248-253.<br />

[CST] A. Connes, D. Sullivan, N. Teleman, Quasconformal mapp<strong>in</strong>gs, operators on<br />

Hilbert space, and local formulae for characteristic classes, Prepr<strong>in</strong>t I.H.E.S.<br />

M/38 (1993).<br />

[CE] E. Christensen, D.E. Evans, Cohomology of operator algebras and quantum<br />

dynamical semigroups, J. London. Math. Soc. 20 (1979), 358-368.<br />

[D1] E.B. Davies, (1976), “Quantum <strong>Theory</strong> of Open Systems”, 171 pages, Academic<br />

Press, London U.K..<br />

[D2] E.B. Davies, Analysis on graphs and noncommutative geometry, J. Funct. Anal.<br />

111 (1993), 398-430.<br />

[DL1] E.B. Davies, J.M. L<strong>in</strong>dsay, Non–commutative symmetric Markov semigroups,<br />

Math. Z. 210 (1992), 379-411.<br />

[DL2] E.B. Davies, J.M. L<strong>in</strong>dsay, Superderivations and symmetric Markov semigroups,<br />

Comm. Math. Phys. 157 (1993), 359-370.<br />

[DR1] E.B. Davies, O.S. Rothaus, Markov semigroups on C ∗ –bundles, J. Funct. Anal.<br />

85 (1989), 264-286.<br />

[DR2] E.B. Davies, O.S. Rothaus, A BLW <strong>in</strong>equality for vector bundles and applications<br />

to spectral bounds, J. Funct. Anal. 86 (1989), 390-410.<br />

[Del] G.F. Dell’Antonio, Structure of the algebras of some free systems, Comm. Math.<br />

Phys. 9 (1968), 81-117.<br />

[Den] J. Deny, Méthodes hilbertien en thorie du potentiel, <strong>Potential</strong> <strong>Theory</strong> (C.I.M.E.,<br />

ICiclo,Stresa), Ed. Cremonese Roma, 1970. 85, 121-201.<br />

[Dir] P. A. M. Dirac, The quantum theory of the electron, Pro. Roy. Soc. of London<br />

A 117 (1928), 610-624.<br />

[Do] J.L. Doob, “<strong>Classical</strong> potential theory and its probabilistic counterpart”,<br />

Spr<strong>in</strong>ger-Verlag, New York, 1984.<br />

[Dix1] J. Dixmier, “Les C ∗ –algèbres et leurs représentations”, Gauthier–Villars, Paris,<br />

1969.<br />

[Dix2] J. Dixmier, “Les algébres d’operateurs dans les espaces hilbertienne (algébres de<br />

von Neumann)”, Gauthier–Villars, Paris, 1969.<br />

[F] W. Faris, Invariant cones and uniqueness of the ground state for Fermion systems,<br />

J. Math. Phys. 13 (1972), 1285-1290.<br />

[FU] F. Fagnola, V. Umanita’, Generators of detailed balance quantum Markov semigroups,<br />

Inf<strong>in</strong>. Dimens. Anal. Quantum Probab. Relat. Top. 10 (2007), 335-363.<br />

[FOT] M. Fukushima, Y. Oshima, M. Takeda, “Dirichlet Forms and Symmetric Markov<br />

Processes”, de Gruyter Studies <strong>in</strong> Mathematics, 1994.


274 F. Cipriani<br />

[GL1] S. Goldste<strong>in</strong>, J.M. L<strong>in</strong>dsay, Beurl<strong>in</strong>g–Deny conditions for KMS–symmetric dynamical<br />

semigroups, C. R. Acad. Sci. Paris, Ser. I 317 (1993), 1053-1057.<br />

[GL2] S. Goldste<strong>in</strong>, J.M. L<strong>in</strong>dsay, KMS–symmetric Markov semigroups, Math. Z. 219<br />

(1995), 591-608.<br />

[GL3] S. Goldste<strong>in</strong>, J.M. L<strong>in</strong>dsay, Markov semigroup KMS–symmetric for a weight,<br />

Math. Ann. 313 (1999), 39-67.<br />

[GKS] V. Gor<strong>in</strong>i, A. Kossakowski, E.C.G. Sudarshan, Completely positive dynamical<br />

semigroups of N-level systems, J. Math. Phys. 17 (1976), 821-825.<br />

[GS] D.Goswami, K.B. S<strong>in</strong>ha, “Quantum stochastic processes and noncommutative<br />

geometry”, Cambridge Tracts <strong>in</strong> Mathematics 169, 290 pages, Cambridge University<br />

press, 2007.<br />

[G1] L. Gross, Existence and uniqueness of physical ground states, J. Funct. Anal.<br />

10 (1972), 59-109.<br />

[G2] L. Gross, Hypercontractivity and logarithmic Sobolev <strong>in</strong>equalities for the<br />

Clifford–Dirichlet form, Duke Math. J. 42 (1975), 383-396.<br />

[GIS] D.Guido,T.Isola,S.Scarlatti, Non–symmetricDirichletformsonsemif<strong>in</strong>ite<br />

von Neumann algebras, J. Funct. Anal. 135 (1996), 50-75.<br />

[H1] U. Haagerup, Standard forms of von Neumann algebras, Math. Scand. 37 (1975),<br />

271-283.<br />

[H2] U. Haagerup, L p -spaces associated with an arbitrary von Neumann algebra,<br />

Algebre d’opérateurs et leur application en Physique Mathematique. Colloques<br />

Internationux du CNRS 274, Ed. du CNRS, Paris 1979 , 175-184.<br />

[H3] U. Haagerup, All nuclear C ∗ -algebras are amenable, Invent. Math. 74 (1983),<br />

no. 2, 305-319.<br />

[HHW] R. Haag, N.M. Hugenoltz, M. W<strong>in</strong>n<strong>in</strong>k, On the equilibrium states <strong>in</strong> quantum<br />

statistical mechanics, Comm. Math. Phys. 5 (1967), 215-236.<br />

[Io] B. Iochum, “Cones autopolaires et algebres de Jordan”, Lecture Notes <strong>in</strong> Mathematics,<br />

vol. 1049, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 1984.<br />

[Ki] J. Kigami, “Analysis on Fractals”, Cambridge Tracts <strong>in</strong> Mathematics vol. 143,<br />

Cabridge: Cambridge University Press, 2001.<br />

[KP] C.K. Ko, Y.M. Park, Construction of a family of quantum Ornste<strong>in</strong>-Uhlenbeck<br />

semigroups J. Math. Phys. 45, (2004), 609–627.<br />

[Ko] H. Kosaki, Application of the complex <strong>in</strong>terpolation method to a von Neumann<br />

algebra: non-commutative L p -spaces, J. Funct. Anal. 56 (1984), 29-78.<br />

[KFGV] A. Kossakowski, A. Frigerio, V. Gor<strong>in</strong>i, M. Verri, Quantum detailed balance and<br />

the KMS condition, Comm. Math. Phys. 57 (1977), 97-110.<br />

[Kub] R. Kubo, Statistical-mechanical theory of irreversible processes. I. General theory<br />

and simple applications to magnetic and conduction problems, J. Phys. Soc.<br />

Japan 12 (1957), 570-586.<br />

[LM] H.B. Lawson JR., M.-L. Michelson, “Sp<strong>in</strong> Geometry”, Pr<strong>in</strong>ceton University<br />

Press, Pr<strong>in</strong>ceton New Jersey, 1989.<br />

[LJ] Y, Le Jan, Mesures associés a une forme de Dirichlet. Applications., Bull. Soc.<br />

Math. France 106 (1978), 61-112.<br />

[Lig] T.M. Liggett, “Interact<strong>in</strong>g Particle Systems”, Grund. math. Wissen., 276.<br />

Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 1985.<br />

[L<strong>in</strong>] G, L<strong>in</strong>dblad, On the generators of quantum dynamical semigroups, Comm.<br />

Math. Phys. 48 (1976), 119-130.<br />

[LR] E.H. Lieb, D.W. Rob<strong>in</strong>son, The f<strong>in</strong>ite group velocity of quantum sp<strong>in</strong> systems,<br />

Comm. Math. Phys. 28 (1972), 251-257.<br />

[MZ1] A. Majewski, B. Zegarl<strong>in</strong>ski, Quantum stochastic dynamics. I. Sp<strong>in</strong> systems on<br />

a lattice, Math. Phys. Electron. J. 1 (1995), Paper 2, 1-37.<br />

[MZ2] A. Majewski, B. Zegarl<strong>in</strong>ski, On Quantum stochastic dynamics on noncommutative<br />

Lp–spaces, Lett. Math. Phys. 36 (1996), 337-349.<br />

[MOZ] A. Majewski, R. Olkiewicz, B. Zegarl<strong>in</strong>ski, Dissipative dynamics for quantum<br />

sp<strong>in</strong> systems on a lattice, J. Phys. A 31 (1998), no. 8, 2045-2056.


Dirichlet Forms on Noncommutative Spaces 275<br />

[Mat1] T. Matsui, Markov semigroups on UHF algebras, Rev. Math. Phys. 5 (1993),<br />

no. 3, 587-600.<br />

[Mat2] T. Matsui, Markov semigroups which describe the time evolution of some higher<br />

sp<strong>in</strong> quantum models, J. Funct. Anal. 116 (1993), no. 1, 179-198.<br />

[Mat3] T. Matsui, Quantum statistical mechanics and Feller semigroups, Quantum<br />

probability communications QP-PQ X 31 (1998), 101-123.<br />

[Mok] G. Mokobodzki, Fermabilité des formes de Dirichlet et <strong>in</strong>égalité detypePo<strong>in</strong>caré,<br />

Pot. Anal. 4 (1995), 409-413.<br />

[MS] D.C. Mart<strong>in</strong>, J. Schw<strong>in</strong>ger, <strong>Theory</strong> of many-particle systems. I., J. Phys. Rev.<br />

115 (1959), 1342-1373.<br />

[MvN] F.J. Murray, J. von Neumann, On r<strong>in</strong>gs of operators, Ann. Math. 37 (1936),<br />

116-229. On r<strong>in</strong>gs of operators II, Trans. Amer. Math. Soc. 41 (1937), 208-248.<br />

On r<strong>in</strong>gs of operators IV, Ann. Math. 44 (1943), 716-808.<br />

[Ne] E. Nelson, Notes on non-commutative <strong>in</strong>tegration, Ann. Math. 15 (1974),<br />

103-116.<br />

[OP] M. Ohya, D. Petz, “Quantum Entropy and its use”, viii + 335 pages, Spr<strong>in</strong>ger-<br />

Verlag, Berl<strong>in</strong>, Heidelberg, New York, 1993.<br />

[P1] Y.M. Park, Construction of Dirichlet forms on standard forms of von Neumann<br />

algebras Inf<strong>in</strong>ite Dim. Anal., Quantum. Prob. and Related Topics 3 (2000),<br />

1-14.<br />

[P2] Y.M. Park, Ergodic property of Markovian semigroups on standard forms of von<br />

Neumann algebras J. Math. Phys. 46 (2005), 113507.<br />

[P3] Y.M. Park, Remarks on the structure of Dirichlet forms on standard forms of<br />

von Neumann algebras Inf<strong>in</strong>. Dimens. Anal. Quantum Probab. Relat. 8 no. 2<br />

(2005), 179-197.<br />

[Par] K. R. Parthasarathy, “An <strong>in</strong>troduction to quantum stochastic calculus”, Monographs<br />

<strong>in</strong> Mathematics, 85. Birkhuser Verlag, Basel, 1992.<br />

[Ped] G. Pedersen, “C ∗ -algebras and their authomorphisms groups”, London Mathematical<br />

Society Monographs, 14. Academic Press, Inc., London-New York, 1979.<br />

[Pen] R.C. Penney, Self-dual cones <strong>in</strong> Hilbert space., J. Funct. Anal. 21 (1976), 305-315.<br />

[Pet] P. Petersen, “Riemannian Geometry”, Graduate Text <strong>in</strong> Mathematics 171, xii<br />

+ 432 pages, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, Heidelberg, New York, 1998.<br />

[Petz] D. Petz, A dual <strong>in</strong> von Neumann algebra Quart. J. Math. Oxford 35 (1984),<br />

475-483.<br />

[RS] M. Reed, B. Simon, “Methods of modern mathematical physics. II. Fourier<br />

Analysis, Self-adjo<strong>in</strong>tness”, Academic Press, xi+361 pages, New York - London,<br />

1975.<br />

[Ric] C.E. Rickart, “General <strong>Theory</strong> of Banach Algebras”, D. van Nostrand Company<br />

Inc. Pr<strong>in</strong>ceton New Jersey, 1960.<br />

[S1] J.-L. Sauvageot, Sur le produit tensoriel relatif d’espaces de Hilbert, J. Operator<br />

<strong>Theory</strong>. 9 (1983), 237-252.<br />

[S2] J.-L. Sauvageot, Tangent bimodule and locality for dissipative operators on C ∗ –<br />

algebras, Quantum <strong>Probability</strong> and Applications IV, Lecture Notes <strong>in</strong> Math.<br />

1396 (1989), 322-338.<br />

[S3] J.-L. Sauvageot, Quantum differential forms, differential calculus and semigroups,<br />

Quantum <strong>Probability</strong> and Applications V, Lecture Notes <strong>in</strong> Math. 1442<br />

(1990), 334-346.<br />

[S4] J.-L. Sauvageot, Semi–groupe de la chaleur transverse sur la C ∗ –algèbre d’un<br />

feulleitage riemannien, C.R. Acad. Sci. Paris Sér. I Math. 310 (1990), 531-536.<br />

[S5] J.-L. Sauvageot, Le probleme de Dirichlet dans les C ∗ –algèbres, J. Funct. Anal.<br />

101 (1991), 50-73.<br />

[S6] J.-L. Sauvageot, From classical geometry to quantum stochastic flows: an example,<br />

Quantum probability and related topics, QP-PQ, VII, 299-315, World Sci.<br />

Publ., River Edge, NJ, 1992.


276 F. Cipriani<br />

[S7] J.-L. Sauvageot, Semi–groupe de la chaleur transverse sur la C ∗ –algèbre d’un<br />

feulleitage riemannien, J. Funct. Anal. 142 (1996), 511-538.<br />

[S8] J.-L. Sauvageot, Strong Feller semigroups on C ∗ -algebras, J. Op. Th. 42 (1999),<br />

83-102.<br />

[Se] E.A. Segal, A non-commutative extension of abstract <strong>in</strong>tegration, Ann. of Math.<br />

57 (1953), 401-457.<br />

[SU] R. Schrader, D. A. Uhlenbrock, Markov structures on Clifford algebras, Jour.<br />

Funct. Anal. 18 (1975), 369-413.<br />

[Sew] G. Sewell, “Quantum Mechanics and its Emergent Macrophysics”, Pr<strong>in</strong>ceton<br />

Universisty Press, 292 pages, Pr<strong>in</strong>ceton and Oxford, 2002.<br />

[Sti] W.F. St<strong>in</strong>espr<strong>in</strong>g, Positive functions on C ∗ -algebras, Proc. Amer. Math. Soc. 6<br />

(1975), 211-216.<br />

[T1] M. Takesaki, “Structure of factors and automorphism groups”, CBMS Regional<br />

Conference Series <strong>in</strong> Mathematics, American Mathematical Society, Providence,<br />

R.I. 51 (1983).<br />

[T2] M. Takesaki, “<strong>Theory</strong> of Operator Algebras I”, Encyclopedia of Mathematical<br />

Physics, 415 pages, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, Heidelberg, New York, 2000.<br />

[Te] M. Terp, Interpolation spaces between a von Neumann algebra and its predual,<br />

J. Operator <strong>Theory</strong> 8 (1982), 327-360.<br />

[V] A. Van Daele, A new approach to the Tomita-Takesaki theory of generalized<br />

Hilbert algebras, J. Funct. Anal. 15 (1974), 379-393.<br />

[Voi1] D.V. Voiculescu, Lectures on Free <strong>Probability</strong> theory., Lecture Notes <strong>in</strong> Math.<br />

1738 (2000), 279-349.<br />

[Voi2] D.V. Voiculescu, The analogues of entropy and of Fisher’s <strong>in</strong>formation measure<br />

<strong>in</strong> free probability theory, Invent. Math. 132 (1998), 189-227.


Applications of Quantum Stochastic<br />

Processes <strong>in</strong> Quantum Optics<br />

Luc Bouten<br />

Abstract These lecture notes provide an <strong>in</strong>troduction to quantum filter<strong>in</strong>g<br />

and its applications <strong>in</strong> quantum optics. We start with a brief <strong>in</strong>troduction to<br />

quantum probability, focus<strong>in</strong>g on the spectral theorem. Then we <strong>in</strong>troduce<br />

the conditional expectation and quantum stochastic calculus. In the last part<br />

of the notes we discuss the filter<strong>in</strong>g problem.<br />

1 Quantum <strong>Probability</strong><br />

In quantum theory observables are represented by selfadjo<strong>in</strong>t operators on a<br />

Hilbert space. When an observable is be<strong>in</strong>g measured, we randomly obta<strong>in</strong><br />

a measurement result. That is, the result of the measurement is given by a<br />

random variable on some classical probability space. In this section we will<br />

<strong>in</strong>vestigate how we can pass from a selfadjo<strong>in</strong>t operator to a classical random<br />

variable. Along the way we will setup a generalised theory of probability,<br />

called noncommutative or quantum probability, that is rich enough to conta<strong>in</strong><br />

quantum mechanics.<br />

1.1 The Spectral Theorem<br />

In these notes we take as a part of its def<strong>in</strong>ition that a Hilbert space is<br />

separable. The follow<strong>in</strong>g theorem shows that a self-adjo<strong>in</strong>t operator S on a<br />

L. Bouten<br />

Caltech Physical Measurement and Control, MC 266-33,<br />

1200 E California Blvd, Pasadena, CA 91125, USA<br />

e-mail: bouten@caltech.edu<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 277<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


278 L. Bouten<br />

Hilbert space can be identified with a random variable h on a measure space<br />

(Ω,Σ,µ). It is the heart of quantum mechanics.<br />

Theorem 1.1. (Spectral Theorem) Let H be a Hilbert space and let S :<br />

H→Hbe a bounded selfadjo<strong>in</strong>t operator. Then there exist a measure space<br />

(Ω,Σ,µ), aunitaryU : H→L 2 (Ω,Σ,µ) and a bounded real valued Σmeasurable<br />

function h on Ω such that<br />

S = U ∗ MhU,<br />

where the multiplication operator Mh : L 2 (Ω,Σ,µ) → L 2 (Ω,Σ,µ) is given by<br />

for all ω ∈ Ω.<br />

(Mhf)(ω) =h(ω)f(ω),<br />

Proof. See Reed and Simon volume I [23].<br />

Suppose that the dimension n of H is f<strong>in</strong>ite. S<strong>in</strong>ce S is self-adjo<strong>in</strong>t, we<br />

can always f<strong>in</strong>d an orthonormal basis (e1,...,en) such that S is diagonal,<br />

i.e. S = diag(h1,...,hn). Note that the real numbers hi and hj are not<br />

necessarily different numbers, some eigenvalues might be degenerate. Let us<br />

def<strong>in</strong>e Ω = {1,...,n}. Moreover, def<strong>in</strong>e h : Ω → R by h(ω) =hω for all<br />

ω ∈ Ω. LetΣ = P(Ω) be the sigma-algebra of all subsets of Ω. Wetake<br />

for µ simply the count<strong>in</strong>g measure. Def<strong>in</strong>e U : H→L 2 (Ω,Σ,µ) by l<strong>in</strong>ear<br />

extension of the map<br />

Ueω = χ {ω}, ω ∈ Ω,<br />

where χ {ω}(ω ′ )is1ifω ′ = ω and 0 otherwise. Note that U is unitary and<br />

that S = U ∗ MhU. That means, we have now proved the spectral theorem<br />

<strong>in</strong> f<strong>in</strong>ite dimension. We now also see that the ma<strong>in</strong> idea of the theorem is to<br />

diagonalise an operator. The spectral theorem gives a precize mean<strong>in</strong>g to the<br />

notion of diagonalisation also <strong>in</strong> the <strong>in</strong>f<strong>in</strong>ite dimensional case.<br />

Let X : R → R be the map given by X(x) =x and let p = �l m<br />

m=0<br />

αmX<br />

be a polynomial. Suppose S is a self-adjo<strong>in</strong>t operator on a Hilbert space<br />

H. The spectral theorem provides a measure space (Ω,Σ,µ), a unitary U :<br />

H→L2 (Ω,Σ,µ) andaΣ-measurable real valued function h such that S =<br />

U ∗MhU. It is easy to see that<br />

l�<br />

m=0<br />

αmS m = U ∗ Mp◦hU. (1.1)<br />

This shows that we can def<strong>in</strong>e the polynomial p evaluated at S, denotedp(S),<br />

<strong>in</strong> two equivalent ways, i.e. either by the left hand side or the right hand of<br />

Equation (1.1). The spectrum of an operator T : H→His def<strong>in</strong>ed as the set<br />

σ(T )= � λ ∈ C; T − λI does not have a bounded <strong>in</strong>verse � . (1.2)


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 279<br />

In f<strong>in</strong>ite dimensions the spectrum of an operator is just the set of its eigenvalues.<br />

The spectrum of a bounded operator T is compact. For the operator<br />

S, diagonalised by the spectral theorem as S = U ∗MhU, itisnothardto<br />

see that h(ω) ∈ σ(S) forµ-almost all ω ∈ Ω and that σ(S) is the smallest<br />

compact subset of C with this property. We denote the range of h by Ran(h)<br />

and are guided by the idea Ran(h) =σ(S).<br />

Def<strong>in</strong>ition 1.2. (Functional calculus) Let S : H → H be a bounded,<br />

self-adjo<strong>in</strong>t operator and let j : σ(S) → C be a bounded Borel measurable<br />

function. The spectral theorem provides a measure space (Ω,Σ,µ), a unitary<br />

U : H→L2 (Ω,Σ,µ) andaΣ-measurable real valued function h such that<br />

S = U ∗MhU. Def<strong>in</strong>e the bounded l<strong>in</strong>ear operator j(S) :H→Hby<br />

j(S) =U ∗ Mj◦hU.<br />

For polynomials p it is clear from Equation (1.1) that this def<strong>in</strong>ition does<br />

not depend on which ‘diagonalisation’ of S we choose (the spectral theorem<br />

shows existence, not uniqueness of (Ω,Σ,µ), U and h). With polynomials<br />

we can uniformly approximate cont<strong>in</strong>uous functions arbitrarily well (σ(S) is<br />

compact) and with cont<strong>in</strong>uous functions we can approximate bounded Borel<br />

measurable functions arbitrarily well po<strong>in</strong>twise. In this way it follows that<br />

the def<strong>in</strong>ition of j(S) does not depend on the diagonalisation for all bounded<br />

measurable functions j. See Reed and Simon [23] for further <strong>in</strong>formation.<br />

Suppose the bounded self-adjo<strong>in</strong>t operator S represents an observable of<br />

some physical system. Suppose the system is <strong>in</strong> a state given by a vector v<br />

of unit length <strong>in</strong> the Hilbert space H. We now discuss some of the basics of<br />

quantum mechanics <strong>in</strong> this sett<strong>in</strong>g. When it is measured, the observable S<br />

can only take values <strong>in</strong> the spectrum σ(S). Let B be a Borel subset of σ(S).<br />

The set B corresponds to the event ‘S takes a value <strong>in</strong> B’. The probability<br />

of this event is given by<br />

P � S takes a value <strong>in</strong> B � = � v, χB(S)v � ,<br />

where χB is the characteristic function of the Borel set B, i.e. χB(s) =1if<br />

s ∈ B and zero otherwise. The spectral theorem provides a measure space<br />

(Ω,Σ,µ), a unitary U : H→L 2 (Ω,Σ,µ) andaΣ-measurable real valued<br />

function h such that S = U ∗ MhU. By def<strong>in</strong>ition χB(S) =U ∗ MχB◦hU =<br />

U ∗ Mχ h −1 (B) U.Notethats<strong>in</strong>ceh is measurable, the set h −1 (B) isanelement<br />

of Σ. In fact, we can assign probabilities to all Q ∈ Σ by<br />

P(Q) = � v, U ∗ MχQUv � .<br />

Summariz<strong>in</strong>g, we can represent S as a random variable h on the probability<br />

space (Ω,Σ,P). The spectral theorem transforms quantum mechanics to<br />

classical probability theory.<br />

We will now extend the spectral theorem so that we can simultaneously diagonalise<br />

a whole class of commut<strong>in</strong>g bounded normal operators. An operator


280 L. Bouten<br />

S : H→His called normal if it commutes with its adjo<strong>in</strong>t, i.e. SS ∗ = S ∗ S.<br />

Before stat<strong>in</strong>g the theorem, we need to <strong>in</strong>troduce some mathematical objects.<br />

Let H be a Hilbert space. Denote by B(H) the algebra of all bounded<br />

operators on H and let S be a subset of B(H). We call the set<br />

S ′ = � R ∈B(H); RS = SR ∀S ∈S � ,<br />

the commutant of S <strong>in</strong> B(H). A von Neumann algebra A on H is a<br />

∗ ′′ -subalgebra of B(H) that equals its double commutant, i.e. A = A. It<br />

is a consequence of von Neumann’s double commutant theorem (see e.g. [17])<br />

that a von Neumann algebra is closed <strong>in</strong> the weak operator topology. It immediately<br />

follows from A ′′ = A that the identity I ∈B(H) isanelementof<br />

the von Neumann algebra A. Astate is a l<strong>in</strong>ear map ρ : A→C such that ρ<br />

is positive <strong>in</strong> the sense that ρ(S∗S) ≥ 0 for all S ∈A,andsuchthatρis normalised<br />

ρ(I) = 1. A state is called normal if it is weak operator cont<strong>in</strong>uous<br />

on the unit ball of A.<br />

Theorem 1.3. (Spectral Theorem) Let C be a commutative von Neumann<br />

algebra and let ρ be a normal state on C. There exist a measure space (Ω,Σ,µ)<br />

and a ∗-isomorphism ι from C to L∞ (Ω,Σ,µ), the space of all bounded measurable<br />

functions on Ω. Furthermore, there exists a probability measure P on<br />

Σ such that<br />

for all C ∈C.<br />

�<br />

ρ(C) =<br />

Ω<br />

ι(C)(ω)P(dω),<br />

Theorem 1.3 can be obta<strong>in</strong>ed from Theorem 1.1 by us<strong>in</strong>g that under mild<br />

conditions (which are fulfilled s<strong>in</strong>ce our von Neumann algebra acts on a separable<br />

Hilbert space) a commutative von Neumann algebra C is generated by<br />

a s<strong>in</strong>gle bounded selfadjo<strong>in</strong>t operator S [27].<br />

Given a probability space (Ω,Σ,P), we can study the commutative von<br />

Neumann algebra C := L ∞ (Ω,Σ,P), act<strong>in</strong>g on the Hilbert space L 2 (Ω,Σ,P)<br />

by po<strong>in</strong>twise multiplication equipped with the normal state ρ given by expectation<br />

with respect to the measure P. The pair (C,ρ) faithfully encodes the<br />

probability space (Ω,Σ,P) [19]. Indeed, the sigma-algebra Σ can be reconstructed<br />

(up to equivalence of sets with P-null symmetric difference, a po<strong>in</strong>t<br />

on which we will not dwell here) as the set of projections <strong>in</strong> C, i.e. the set of<br />

characteristic functions of sets <strong>in</strong> Σ, and the probability measure is given by<br />

act<strong>in</strong>g with the state ρ on this set of projections. We conclude that study<strong>in</strong>g<br />

commutative von Neumann algebras equipped with normal states is the same<br />

as study<strong>in</strong>g probability spaces. This motivates the follow<strong>in</strong>g def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 1.4. A non-commutative or quantum probability space is a pair<br />

(A,ρ)whereA is a von Neumann algebra on some Hilbert space H and ρ is<br />

a normal state.<br />

Note that we do not require the state ρ to be faithful. We use quantum<br />

probability spaces to model experiments <strong>in</strong>volv<strong>in</strong>g quantum mechanics. The


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 281<br />

firstth<strong>in</strong>gtodoistosetamodel,say(A,ρ). It is a feature of quantum<br />

mechanics that <strong>in</strong> one s<strong>in</strong>gle realization of an experiment only commut<strong>in</strong>g<br />

observables can be measured. That is, the experiment is determ<strong>in</strong>ed by a<br />

commutative von Neumann subalgebra C of A. The pair (C,ρ|C) isequivalent<br />

to a classical probability model (Ω,Σ,P) via the spectral theorem. The<br />

operators <strong>in</strong> C are mapped to random variables ι(C) on(Ω,Σ,P) and represent<br />

the stochastic measurement results of the experiment. The pair (A,ρ)<br />

describes the collection of all the experiments (i.e. the collection of commutative<br />

subalgebras of A) and their statistics <strong>in</strong> different realizations of the<br />

experiment <strong>in</strong> a concise way.<br />

1.2 Unbounded Operators<br />

Up to now we have only considered bounded operators. In general, however,<br />

an observable is given by a selfadjo<strong>in</strong>t, not necessarily bounded, operator<br />

on some dense doma<strong>in</strong> of the Hilbert space H. In these notes we will not<br />

put emphasis on the problems one encounters when deal<strong>in</strong>g with unbounded<br />

operators. When add<strong>in</strong>g or multiply<strong>in</strong>g unbounded operators, it is important<br />

to keep track of the doma<strong>in</strong>s <strong>in</strong>volved. In this subsection we discuss some<br />

techniques for relat<strong>in</strong>g an unbounded operator to a von Neumann algebra.<br />

Def<strong>in</strong>ition 1.5. Let T be a densely def<strong>in</strong>ed l<strong>in</strong>ear operator on some Hilbert<br />

space H. Denote the doma<strong>in</strong> of T by Dom(T ). The operator T is called<br />

closed if its graph {(x, T x); x ∈ Dom(T )} is a closed subset of H×H. Def<strong>in</strong>e<br />

Dom(T ∗ )asthesetofx ∈Hfor which there is a y ∈Hsuch that<br />

〈Tz,x〉 = 〈z,y〉, ∀z ∈ Dom(T ).<br />

For each such x ∈ Dom(T ∗ ), we def<strong>in</strong>e T ∗ x = y. WecallT ∗ the adjo<strong>in</strong>t of T .<br />

The operator T is called symmetric if T ⊂ T ∗ , i.e. Dom(T ) ⊂ Dom(T ∗ )and<br />

Tz = T ∗ z for all z ∈ Dom(T ). Equivalently, T is symmetric if and only if<br />

〈Tz,y〉 = 〈z,Ty〉 for all z,y ∈ Dom(T). The operator T is called selfadjo<strong>in</strong>t<br />

if T = T ∗ , i.e. if T is symmetric and Dom(T )=Dom(T ∗ ).<br />

Theorem 1.6. (basic criterion for selfadjo<strong>in</strong>tness) Let T be a symmetric<br />

operator on a Hilbert space H. The follow<strong>in</strong>g three statements are equivalent<br />

1. T is selfadjo<strong>in</strong>t.<br />

2. T is closed, Ker(T ∗ + iI) ={0} and Ker(T ∗ − iI) ={0}.<br />

3. Ran(T + iI) =H and Ran(T − iI) =H.<br />

The proof of the above theorem can be found <strong>in</strong> [23]. Theorem 1.6 shows<br />

that T + iI is an <strong>in</strong>jective operator with range H. This means (T + iI) −1<br />

is well-def<strong>in</strong>ed with doma<strong>in</strong> H. Moreover,s<strong>in</strong>ceT + iI is closed, (T + iI) −1 is<br />

also closed. The closed graph theorem (see [23]) then asserts that (T + iI) −1


282 L. Bouten<br />

is bounded. A similar argument shows that T − iI is <strong>in</strong>vertible and that<br />

(T − iI) −1 is bounded. Let λ be <strong>in</strong> the complement of the spectrum of T (i.e.<br />

T − λI has a bounded <strong>in</strong>verse). Def<strong>in</strong>e the resolvent of T at λ as Rλ(T )=<br />

(T − λI) −1 (See also Section 3.2 of the lectures by N. Privault). Note that<br />

we have the follow<strong>in</strong>g identity<br />

Rλ(T ) − Rµ(T )=Rλ(T )(T − µI)Rµ(T ) − Rλ(T )(T − λI)Rµ(T )<br />

=(λ− µ)Rλ(T )Rµ(T ).<br />

Add<strong>in</strong>g this same identity with the roles of µ and λ <strong>in</strong>terchanged shows that<br />

Rλ(T )andRµ(T ) commute. Therefore (T + iI) −1 and (T − iI) −1 commute.<br />

The equality<br />

� (T −iI)x, (T +iI) −1 (T +iI)y � = � (T −iI) −1 (T −iI)x, (T +iI)y � , x,y ∈H,<br />

and the fact that Ran(T ± iI) =H shows that (T + iI) −1∗ =(T − iI) −1 .<br />

Therefore (T + iI) −1 is normal and therefore the von Neumann algebra generated<br />

by (T + iI) −1 is commutative.<br />

Def<strong>in</strong>ition 1.7. Let A be a von Neumann algebra on some Hilbert space<br />

H. LetT be a selfadjo<strong>in</strong>t operator on a dense doma<strong>in</strong> of H. T is said to be<br />

affiliated to A if (T − iI) −1 is an element of A. WedenotethisasTηA.<br />

Note that T is affiliated to the commutative von Neumann algebra generated<br />

by (T + iI) −1 . Suppose that C is a commutative von Neumann algebra<br />

on some Hilbert space H and suppose that T is a selfadjo<strong>in</strong>t operator affiliated<br />

to C. Letρ be a normal state on C. The spectral theorem (Theorem 1.3)<br />

provides a measure space (Ω,Σ,µ), a ∗ -isomorphism from C→L ∞ (Ω,Σ,µ)<br />

and a probability measure P on Σ such that ρ(C) =EP(ι(C)) for all C ∈C.<br />

That is, ι allows us to represent the elements <strong>in</strong> C as classical random variables<br />

on the probability space (Ω,Σ,P). S<strong>in</strong>ce (T + iI) −1 is an element of C<br />

we can now also represent T on (Ω,Σ,P) by<br />

ι(T )(ω) =<br />

1<br />

ι � (T + iI) −1� − i, ω ∈ Ω. (1.3)<br />

(ω)<br />

Note that ι(T ) is not bounded. Nevertheless, we have succeeded <strong>in</strong> represent<strong>in</strong>g<br />

T as a classical random variable on (Ω,Σ,P).<br />

The follow<strong>in</strong>g theorem shows there is a one-one correspondence between<br />

strongly cont<strong>in</strong>uous one-parameter groups of unitaries and selfadjo<strong>in</strong>t operators.<br />

The theorem is a classic result and can be found e.g. <strong>in</strong> [17] or [23].<br />

Theorem 1.8. (Stone’s theorem) Let A be a von Neumann algebra on<br />

some Hilbert space H and let {Ut}t∈R ⊂Abe a group of unitaries that is<br />

cont<strong>in</strong>uous <strong>in</strong> the strong operator topology. There exists a unique selfadjo<strong>in</strong>t<br />

operator S affiliated to A such that Ut =exp(itS) for all t ∈ R. The operator<br />

S is called the Stone generator of {Ut}t∈R.


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 283<br />

1.3 Wiener and Poisson Processes<br />

We now <strong>in</strong>troduce the quantum probability space with which we will model<br />

the quantized electromagnetic field. We will see that the model is rich enough<br />

to support an entire family of Wiener and Poisson processes. These different<br />

processes do not all commute with each other. In this sense quantum<br />

probability is richer than the classical theory.<br />

Def<strong>in</strong>ition 1.9. Let H be a Hilbert space. The bosonic or symmetric Fock<br />

space over H is the Hilbert space<br />

F(H) =C ⊕<br />

∞�<br />

H ⊗sn ,<br />

n=1<br />

where ⊗s denotes the symmetric tensor product.<br />

The Fock space F(H) describes a field of bosons. The different levels <strong>in</strong> the<br />

Fock space correspond to different numbers of photons present <strong>in</strong> the field.<br />

Note that it is possible to have superpositions between states with different<br />

number of particles. For f ∈Hwe def<strong>in</strong>e the exponential vector e(f) by<br />

e(f) =1⊕<br />

∞�<br />

n=1<br />

1<br />

√ n! f ⊗n .<br />

The vector Φ = e(0) = 1 ⊕ 0 ⊕ 0 ⊕ ... is called the vacuum vector. Theset<br />

of all exponential vectors is l<strong>in</strong>early <strong>in</strong>dependent and the l<strong>in</strong>ear span of all<br />

exponential vectors D is a dense subset of F(H) (see e.g. [21]). On the dense<br />

doma<strong>in</strong> D we def<strong>in</strong>e for all f ∈Han operator W (f) by<br />

W (f)e(g) =e<br />

1 −〈f,g〉− 2<br />

||f|| 2<br />

e(f + g), g ∈H. (1.4)<br />

W (f) is an isometric map D→D. Therefore, W (f) extends uniquely to a<br />

unitary operator on F(H). The extension of W (f) toF(H) is also denoted<br />

by W (f). The operators W (f) are called Weyl operators. Itcanbeshown<br />

(e.g. [21]) that the Weyl operators W (f), f ∈Hgenerate the algebra W =<br />

B(F(H)) of all bounded operators on F(H). From the def<strong>in</strong>ition <strong>in</strong> (1.4) it<br />

is easy to see that they satisfy the follow<strong>in</strong>g Weyl relations<br />

1. W(f) ∗ = W (−f), f ∈H,<br />

2. W(f)W (g) =e −iIm〈f,g〉 W (f + g), f,g ∈H.<br />

(1.5)<br />

For a fixed f <strong>in</strong> H, the family {W (sf)}s∈R forms a one-parameter group<br />

of unitaries. This one-parameter group is <strong>in</strong> fact cont<strong>in</strong>uous with respect to<br />

the strong operator topology [22]. Denote its Stone generator by B(f), i.e.<br />

we have W (sf) =exp(isB(f)). The operators B(f) are called field operators.<br />

Let H = L 2 (R + ) and fix α ∈ [0,π). It immediately follows from the Weyl


284 L. Bouten<br />

relations Equation (1.5) that all the operators <strong>in</strong> the set S = {W (sf); f =<br />

e iα χ [0,t], t ≥ 0, s ∈ R} commute. Here χ [0,t] is the <strong>in</strong>dicator function of the<br />

<strong>in</strong>terval [0,t], i.e. the function that is 1 on [0,t] and 0 elsewhere. Let C α be<br />

the commutative von Neumann algebra generated by S. Def<strong>in</strong>e for all t ≥ 0<br />

the follow<strong>in</strong>g selfadjo<strong>in</strong>t operators affiliated to C α<br />

B α t := B(eiα χ [0,t]).<br />

Let the vacuum state φ be given by φ = 〈Φ, · Φ〉. The spectral theorem<br />

(Theorem 1.3) provides a measure space (Ω α ,Σ α ,µ α ), a ∗ -isomorphism<br />

ι : C α → L ∞ (Ω α ,Σ α ,µ α ), and a probability measure P α on Σ α such that<br />

φ(C) =EP α(ι(C)) for all C ∈Cα . Us<strong>in</strong>g the def<strong>in</strong>ition <strong>in</strong> Equation (1.3)<br />

we can represent the operators B α t for t ≥ 0 as random variables ι(B α t )on<br />

(Ωα ,Σα , Pα ). Note that <strong>in</strong> the above discussion the parameter α was fixed.<br />

The operators Bα t and Bα′ s for α �= α ′ do not commute, i.e. they can not be<br />

represented as random variables on the same classical probability space.<br />

We will now study the process ι(Bα t ),t≥ 0on(Ωα ,Σα , Pα ). We calculate<br />

the characteristic function of an <strong>in</strong>crement ι(Bα t ) − ι(Bα s )(wheret≥s) EPα �<br />

e ik<br />

�<br />

α<br />

ι(Bt )−ι(Bα s )<br />

��<br />

= EPα � �<br />

ι e ik<br />

�<br />

α<br />

Bt −Bα<br />

��� �<br />

s = φ W � �<br />

kχ [s,t]<br />

�<br />

�<br />

1 −<br />

= Φ, e 2 k2 �<br />

1<br />

(t−s) −<br />

e(kχ[s,t]) = e 2 k2 (t−s)<br />

, k ∈ R.<br />

This shows that the <strong>in</strong>crement has a mean zero Gaussian distribution with<br />

variance t − s (see e.g. [28]). Furthermore, a similar calculation shows that<br />

the jo<strong>in</strong>t characteristic function of two <strong>in</strong>crements is given by (where t1 ≥<br />

s1 ≥ t2 ≥ s2)<br />

EP α<br />

EP α<br />

�<br />

e ik1<br />

�<br />

e ik1<br />

�<br />

α<br />

ι(Bt1 )−ι(B α � �<br />

α<br />

s ) +ik2 ι(B<br />

1 t2 )−ι(B α ��<br />

s )<br />

1 −<br />

2 = e 2 (k2 1 (t1−s1)+k2 2 (t2−s2)) =<br />

�<br />

α<br />

ι(Bt )−ι(B<br />

1 α<br />

��<br />

s )<br />

1 EPα �<br />

e ik2<br />

�<br />

α<br />

ι(Bt )−ι(B<br />

2 α<br />

��<br />

s )<br />

2<br />

, k1,k2 ∈ R.<br />

S<strong>in</strong>ce the jo<strong>in</strong>t characteristic function is multiplicative, the <strong>in</strong>crements are<br />

<strong>in</strong>dependent [28]. That is, the process ι(B α t ) has <strong>in</strong>dependent mean zero<br />

Gaussian <strong>in</strong>crements with variance the length of the <strong>in</strong>crement. This means<br />

)isaWiener process. Note that for every α ∈ [0,π) we have now con-<br />

ι(Bα t<br />

structed a Wiener process on the Fock space. Note that for different values<br />

of α these processes do not commute with each other. The idea to simultaneously<br />

diagonalise the fields {Bα t }t≥0 is implicit <strong>in</strong> some of the earliest work<br />

on quantum field theory. However, Segal [24] <strong>in</strong> the 1950s was the first to<br />

emphasize the connection with probability theory.<br />

We will now construct a Poisson process on the Fock space. The second<br />

quantization of an element A ∈B(H) such that �A� ≤1 (i.e. a contraction)<br />

is def<strong>in</strong>ed by<br />

Γ (A) =I ⊕<br />

∞�<br />

A ⊗n . (1.6)<br />

n=1


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 285<br />

For all contractions A, B ∈B(H) we immediately have Γ (A)Γ (B) =Γ (AB).<br />

This means that if A and B commute, then Γ (A) andΓ (B) also commute.<br />

Let S be a selfadjo<strong>in</strong>t element of B(H). The selfadjo<strong>in</strong>t operator S generates<br />

a one-parameter group Us =exp(isS) of unitaries <strong>in</strong> B(H). After second<br />

quantization this leads to a one-parameter group Γ (exp(isS)) of unitaries<br />

<strong>in</strong> W = B(F(H)). Denote the Stone generator of Γ (exp(itS)) by Λ(S), i.e.<br />

Γ (exp(isS)) = exp(isΛ(S)). Note that Λ(S) is affiliated to the commutative<br />

von Neumann algebra generated by {Γ (Us)}s∈R.<br />

Let H = L 2 (R + ). For t ≥ 0, let Pt be the projection given by Pt = Mχ [0,t] . Pt<br />

is selfadjo<strong>in</strong>t and therefore generates a one-parameter group U t s =exp(isPt).<br />

Let C be the commutative algebra generated by all U t s for s ∈ R,t ≥ 0.<br />

Introduce the shorthand Λt for the Stone generators Λ(Pt). Note that for all<br />

t ≥ 0, Λt is affiliated to C. Onthenth layer of the symmetric Fock space<br />

Γ (exp(isPt)) acts as exp(isPt) ⊗n . Differentiation with respect to s shows<br />

that on the nth layer of the symmetric Fock space Λt = Pt ⊗ I n−1 + I ⊗ Pt ⊗<br />

I n−2 + ...+ I n−1 ⊗ Pt. This shows that Λt counts how many particles (i.e.<br />

photons) are present <strong>in</strong> the <strong>in</strong>terval [0,t].<br />

For f ∈ L 2 (R) we def<strong>in</strong>e a coherent vector ψ(f) as the exponential vector<br />

normalised to unit length, i.e.<br />

ψ(f) =e<br />

1 − 2<br />

||f|| 2<br />

e(f) =W (f)e(0) = W (f)Φ.<br />

We can now def<strong>in</strong>e a coherent state on W = B(F(L2 (R + ))) by ρ(A) =<br />

〈ψ(f),Aψ(f)〉. The spectral theorem (Theorem 1.3) provides a measure space<br />

(Ω,Σ,µ), a ∗-isomorphism ι : C→L∞ (Ω,Σ,µ) and a probability measure<br />

P such that EP(ι(C)) = ρ(C) for all C ∈C. S<strong>in</strong>ce the selfadjo<strong>in</strong>t operators<br />

Λt are affiliated with C, we can represent them as random variables ι(Λt) on<br />

(Ω,Σ,P).<br />

For t1 ≥ s1 ≥ t2 ≥ s2 the jo<strong>in</strong>t characteristic function of the <strong>in</strong>crements<br />

ι(Λt1) − ι(Λs1) andι(Λt2) − ι(Λs2) isgivenby<br />

�<br />

e ik1<br />

� � � ��<br />

ι(Λt1 )−ι(Λs1 ) +ik2 ι(Λt2 )−ι(Λs2 )<br />

=<br />

EP<br />

� Pt1 −Ps 1<br />

� �<br />

ψ(f),Γ e ik1<br />

−||f|| 2� �<br />

e e(f),e e ik1<br />

� � �<br />

e<br />

f,<br />

e ik 1<br />

� Pt1 −Ps 1<br />

�� �<br />

Γ e ik2<br />

�<br />

e ik2<br />

� Pt1 −Ps 1<br />

e ik 2<br />

� Pt2 −Ps 2<br />

�<br />

−1<br />

� Pt2 −Ps 2<br />

� Pt2 −Ps 2<br />

�<br />

f<br />

�<br />

=<br />

�� �<br />

ψ(f) =<br />

� ��<br />

f =<br />

� t1<br />

(e s e 1 ik1 −1)|f| 2 � t2<br />

dλ (e s e 2 ik2 −1)|f| 2 dλ<br />

, k1,k2 ∈ R.<br />

This shows that {ι(Λt)}t≥0 is a process with <strong>in</strong>dependent <strong>in</strong>crements. Moreover,<br />

it shows that ι(Λt) t≥0 is a Poisson process with <strong>in</strong>tensity measure |f| 2 dλ,<br />

where λ stands for the Lebesgue measure. S<strong>in</strong>ce ψ(f) =W (f)Φ, we can also


286 L. Bouten<br />

work with respect to the vacuum by sandwich<strong>in</strong>g Λt by W (f). We conclude<br />

that <strong>in</strong> the vacuum the process W (f) ∗ ΛtW (f) is a Poisson process with <strong>in</strong>tensity<br />

measure |f| 2 dλ.<br />

2 Conditional Expectations<br />

In this section we start with a discussion to illustrate how the classical conditional<br />

expectation can be transferred to quantum models us<strong>in</strong>g the spectral<br />

theorem. In push<strong>in</strong>g the classical concept as far as possible, we arrive at a<br />

def<strong>in</strong>ition for the quantum conditional expectation. In Section 4 we state the<br />

quantum filter<strong>in</strong>g problem <strong>in</strong> terms of the newly def<strong>in</strong>ed quantum conditional<br />

expectation.<br />

2.1 Towards a Def<strong>in</strong>ition<br />

Let B be a von Neumann algebra on a Hilbert space H and let P be a normal<br />

state on B. LetX and Y be two selfadjo<strong>in</strong>t commut<strong>in</strong>g elements of B. Their<br />

expectations are given by P(X) andP(Y ), respectively. In this subsection we<br />

show how to def<strong>in</strong>e the conditional expectation P(X|Y )ofX given Y .<br />

Let us first recall the classical def<strong>in</strong>ition. Suppose that F and G are random<br />

variables on a probability space (Ω,Σ,P). Let us suppose for convenience<br />

that Ω is a f<strong>in</strong>ite set. S<strong>in</strong>ce G is a function on a f<strong>in</strong>ite set, its range Ran(G)<br />

is also a f<strong>in</strong>ite set. The classical conditional expectation EP(F |G) ofF given<br />

G is the random variable (i.e. a measurable function from Ω to C) givenby<br />

EP(F |G)(ω) =<br />

�<br />

g∈Ran(G)<br />

EP(Fχ [G=g])<br />

P([G = g]) χ [G=g](ω), ω ∈ Ω, (2.1)<br />

where [G = g] istheset{ω∈Ω; G(w) =g} and χ [G=g] is the <strong>in</strong>dicator<br />

function of that set. Note that EP(F |G) isnotjustΣ-measurable, but even<br />

σ(G)-measurable, where σ(G) istheσ-algebra generated by G. IfΩ is not a<br />

f<strong>in</strong>ite set, and might even be cont<strong>in</strong>uous, this last po<strong>in</strong>t is the guid<strong>in</strong>g idea.<br />

The conditional expectation is then def<strong>in</strong>ed as the orthogonal projection from<br />

L2 (Ω,Σ,P) ontoL2 (Ω,σ(G), P) (andthenextendedtoL1 ), see [28].<br />

Instead of a direct def<strong>in</strong>ition as <strong>in</strong> Equation (2.1), or as an orthogonal<br />

projection on an L2-space, we can provide an abstract characterization of<br />

the conditional expectation. Given a σ-subalgebra Σ0 of Σ, wecallaΣ0measurable<br />

random variable EP(F |Σ0) a version of the conditional expectation<br />

of F on Σ0, if for all Σ0-measurable random variables S we have<br />

�<br />

EP(F |Σ0)S � = EP(FS). (2.2)<br />

EP


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 287<br />

In geometric terms this is just a characterisation of the projection discussed<br />

above. Therefore, there exists an object that satisfies this def<strong>in</strong>ition, see [28].<br />

It is easy to see that the direct def<strong>in</strong>ition <strong>in</strong> Equation (2.1) satisfies the abstract<br />

characterisation of Equation (2.2) if we take Σ0 = σ(G). Furthermore,<br />

there is almost surely only one EP(F |Σ0) satisfy<strong>in</strong>g (2.2) (hence the term<strong>in</strong>ology<br />

“. . . a version of the...”). Theproofofthisstatementisthesameas that for the quantum case, which we will give below. Start<strong>in</strong>g from the def<strong>in</strong>ition<br />

of the classical conditional expectation EP(F |Σ0) <strong>in</strong> Equation (2.2) we<br />

can prove that it has the follow<strong>in</strong>g properties see e.g. [28]. It is l<strong>in</strong>ear <strong>in</strong> F ,it<br />

maps positive random variables to positive random variables, it preserves χΩ,<br />

it has the module property EP(FS|Σ0) =EP(F |Σ0)S for all Σ0-measurable<br />

functions S, it satisfies the tower property EP(EP(F |Σ1)|Σ0) =EP(F |Σ0)<br />

whenever Σ0 ⊂ Σ1, and it is the least mean square estimate of F (see also<br />

further below).<br />

Let us return to the commut<strong>in</strong>g selfadjo<strong>in</strong>t operators X and Y <strong>in</strong> the von<br />

Neumann algebra B equipped with the normal state P. The operators X<br />

and Y together generate a commutative von Neumann subalgebra X of B.<br />

The spectral theorem provides a measure space (Ω,Σ,µ), a ∗-isomorphism ι<br />

from X to L∞ (Ω,Σ,µ) and a probability measure P on Σ such that P(C) =<br />

EP(ι(C)) for all C ∈X. We know how to condition ι(X) onι(Y ), that can be<br />

done with the classical conditional expectation EP(ι(X)|ι(Y )). It is natural<br />

to def<strong>in</strong>e the quantum conditional expectation as<br />

P(X|Y )=ι −1� � � �<br />

EP ι(X) �ι(Y ) �<br />

.<br />

We now see that we can only condition those observables X ∈Bthat commute<br />

with Y . That, however, is exactly as it should be. In one realization<br />

of an experiment we can only access commut<strong>in</strong>g observables. Two noncommut<strong>in</strong>g<br />

observables can never both be assigned a numerical value <strong>in</strong> a s<strong>in</strong>gle<br />

realization of the experiment, i.e. there is no need for condition<strong>in</strong>g them<br />

on each other. This idea has been called the nondemolition pr<strong>in</strong>ciple. Note,<br />

however, that two operators X1 and X2 that both commute with Y need not<br />

necessarily commute with each other.<br />

It is now straightforward to def<strong>in</strong>e the conditional expectation of a selfadjo<strong>in</strong>t<br />

operator X on a commutative subalgebra C of B with which X commutes,<br />

i.e. XC = CX for all C ∈C.Inshort,letXbethe commutative<br />

algebra generated by the whole of C and X together. Apply the spectral<br />

theorem to (X , P), that enables the def<strong>in</strong>ition of a ι, and then def<strong>in</strong>e<br />

P(X|C) :=ι −1<br />

�<br />

� �<br />

�<br />

ι(X) �σ � ι(C); C ∈C ���<br />

. (2.3)<br />

EP<br />

In the next subsection we start with the formal def<strong>in</strong>ition of the quantum<br />

conditional expectation.


288 L. Bouten<br />

2.2 The Quantum Conditional Expectation<br />

Def<strong>in</strong>ition 2.1. (Quantum conditional expectation) Let (B, P) be a<br />

quantum probability space. Let C be a commutative von Neumann subalgebra<br />

of B. DenotebyA its relative commutant, i.e. A = C ′ := {A ∈B; AC =<br />

CA, ∀C ∈C}.ThenP(·|C) : A→Cis (a version of) the conditional expectation<br />

from A onto C if<br />

P � P(A|C)C � = P(AC), ∀A ∈A, ∀C ∈C. (2.4)<br />

Note that the conditional expectation P( ·|C) is def<strong>in</strong>ed only on the commutant<br />

A of C! The def<strong>in</strong>ition that we gave here is more restrictive than the<br />

one that is usual <strong>in</strong> quantum probability (see e.g. [26]). There, one also allows<br />

for the condition<strong>in</strong>g on noncommutative subalgebras.<br />

In applications, we start with the quantum probability space (B ⊗W, P).<br />

Here B is the algebra with which we model some system of <strong>in</strong>terest (two-level<br />

atom, a gas of atoms etc.), W = B(F(L2 (R))) is the algebra with which we<br />

model the electromagnetic field and P = ρ ⊗ φ, whereρsome normal state<br />

on B and φ is the vacuum state on the field. A commutative subalgebra C<br />

is then generated by the observations we perform (see further below). Next,<br />

we choose A to be the relative commutant of C. The algebra A consists<br />

of the observables that have not been demolished by our observations, i.e.<br />

it consists of all operators that are still compatible with C. The follow<strong>in</strong>g<br />

lemma establishes existence and uniqueness for the conditional expectation<br />

of Def<strong>in</strong>ition 2.1.<br />

Lemma 2.2. The conditional expectation of Def<strong>in</strong>ition 2.1 exists and is<br />

unique with probability one, i.e. any two versions P and Q of P(A|C) satisfy<br />

�P − Q�P =0,where�X�2 P := P(X∗X). Proof. Existence. For a self-adjo<strong>in</strong>t element A ∈Awe can def<strong>in</strong>e P(A|C) via<br />

Equation (2.3). We now need to check that it satisfies the abstract characterisation<br />

Equation (2.4). That, however, follows easily from the classical<br />

couterpart Equation (2.2). If A ∈Ais not self-adjo<strong>in</strong>t, then we write it as<br />

the sum of two self-adjo<strong>in</strong>t elements<br />

A = (A + A∗ ) − i � i(A − A∗ ) �<br />

,<br />

2<br />

and simply extend P( ·|C) l<strong>in</strong>early. It is easy to see that that obeys<br />

Equation (2.4).<br />

Uniqueness w.p. one. Def<strong>in</strong>e the pre-<strong>in</strong>ner product 〈X, Y 〉 := P(X ∗ Y )onA<br />

(it might have nontrivial kernel if P is not faithful.) Then 〈C, A − P(A|C)〉 =<br />

P(C ∗ A) − P(C ∗ P(A|C)) = 0 for all C ∈Cand A ∈A, i.e. A − P(A|C) is<br />

orthogonal to C. NowletP and Q be two versions of P(A|C). It follows that<br />

〈C, P −Q〉 =0forallC ∈C.ButP −Q ∈C,so〈P −Q, P −Q〉 = �P −Q� 2 P =0.


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 289<br />

The next lemma asserts that, as <strong>in</strong> the classical case, the conditional expectation<br />

is the least mean square estimate.<br />

Lemma 2.3. P(A|C) is the least mean square estimate of A given C, i.e.<br />

Proof. For all K ∈Cwe have<br />

�A − P(A|C)�P ≤�A − K�P, ∀ K ∈C.<br />

�A − K� 2 P = � � A − P(A|C)+P(A|C) − K � � 2<br />

P<br />

= � � A − P(A|C) � � 2<br />

P + � � P(A|C) − K � � 2<br />

P ≥ � � A − P(A|C) � � 2<br />

P ,<br />

where, <strong>in</strong> the next to last step, we used that A − P(A|C) is orthogonal to<br />

P(A|C) − K ∈C.<br />

Remark. We have now highlighted the existence, uniqueness with probability<br />

one, and the mean least squares property of the quantum conditional<br />

expectation. The other elementary properties of classical conditional expectations<br />

and their proofs [28] carry over directly to the noncommutative sett<strong>in</strong>g.<br />

In particular, we have l<strong>in</strong>earity, positivity, preservation of the identity,<br />

the tower property P(P(A|B)|C) =P(A|C) ifC ⊂ B, the module property<br />

P(AB|C) =P(A|C)B for B ∈C, etc. As an example, let us prove l<strong>in</strong>earity.<br />

It suffices to show that Z = αP(A|C) +βP(B|C) satisfies the def<strong>in</strong>ition<br />

of P(αA + βB|C), i.e. P(ZC) =P((αA + βB)C) for all C ∈C. But this is<br />

immediate from the l<strong>in</strong>earity of P and Def<strong>in</strong>ition 2.1.<br />

In Section 4 we will need to relate conditional expectations with respect<br />

to different states to each other. This is done by the follow<strong>in</strong>g Bayes-type<br />

formula.<br />

Lemma 2.4. (Bayes formula [8] [7]) Let (B, P) be a noncommutative probability<br />

space. Let C be a commutative von Neumann subalgebra of B and let<br />

A be its relative commutant, i.e. A = C ′ := {A ∈B; AC = CA, ∀C ∈C}.<br />

Furthermore, let Q be an element <strong>in</strong> A such that Q∗Q is <strong>in</strong>vertible and<br />

P(Q∗Q)=1. We can def<strong>in</strong>e a state on A by Q(A) :=P(Q∗AQ) and we<br />

have<br />

Q � X � �C � = P�Q∗XQ � �<br />

�C P � Q∗Q � �<br />

�<br />

, X ∈A.<br />

C<br />

Proof. Let K be an element of C. For all X ∈A,wecanwrite<br />

�<br />

P P � Q ∗ XQ|C � �<br />

K = P � Q ∗ XKQ � = Q � XK � �<br />

= Q Q � X|C � �<br />

K =<br />

�<br />

P Q ∗ QQ � X|C � � � �<br />

K = P P Q ∗ QQ � X|C � � ��<br />

�<br />

K�C<br />

=<br />

�<br />

P P � Q ∗ Q � � � � �<br />

�C Q X|C K .<br />

Note that the <strong>in</strong>vertibility of P � Q ∗ Q � � C � follows immediately from the <strong>in</strong>vertibility<br />

of Q ∗ Q.


290 L. Bouten<br />

3 Quantum Stochastic Calculus<br />

We have seen <strong>in</strong> Section 1.3 that the quantum probability space (W,φ)conta<strong>in</strong>s<br />

a rich variety of Wiener and Poisson processes. We will now <strong>in</strong>troduce<br />

stochastic <strong>in</strong>tegrals with respect to these processes and study their stochastic<br />

calculus. Quantum stochastic calculus has been <strong>in</strong>troduced by Hudson and<br />

Parthasarathy [16]. Quantum stochastic differential equations describe the <strong>in</strong>teraction<br />

between atoms and the electromagnetic field <strong>in</strong> the weak coupl<strong>in</strong>g<br />

limit.<br />

3.1 The Stochastic Integral<br />

Lemma 3.1. Let H1 and H2 be Hilbert spaces. There exists a unique unitary<br />

isomorphism U : F(H1 ⊕H2) →F(H1) ⊗F(H2) such that<br />

Ue(x ⊕ y) =e(x) ⊗ e(y), x ∈H1, y ∈H2. (3.1)<br />

Proof. By l<strong>in</strong>ear extension of the relation <strong>in</strong> Equation (3.1) we get a surjective<br />

map from D = span{e(x); x ∈ H1⊕H2} to D1 ⊗D2 where<br />

Di =span{e(x); x ∈Hi}. NotethatDis dense <strong>in</strong> F(H1 ⊕H2) andD1⊗D2 is dense <strong>in</strong> F(H1) ⊗F(H2). Moreover, for all x1,y1 ∈H1 and x2,y2 ∈H2 we<br />

have<br />

�<br />

e(x1 ⊕ x2),e(y1 ⊕ y2) � = e 〈x1+x2,y1+y2〉 = e 〈x1,y1〉 e 〈x2,y2〉<br />

= � e(x1),e(y1) �� e(x2),e(y2) � ,<br />

i.e. U : D→D1⊗D2 is isometric and therefore extends uniquely to a unitary.<br />

We will often identify F(H1⊕H2)andF(H1)⊗F(H2) us<strong>in</strong>g the unitary of<br />

Lemma 3.1. For all t ≥ s ≥ 0wedef<strong>in</strong>eF = F(L 2 (R + )), F t] = F(L 2 ([0,t])),<br />

F [s,t] = F(L 2 ([s, t])) and F [t = F(L 2 ([t, ∞)). For tn > ... > t1 > 0wehave<br />

L 2 (R + )=L 2 ([0,t1]) ⊕ L 2 ([t1,t2]) ⊕ ...⊕ L 2 ([tn, ∞)). Therefore, <strong>in</strong> the sense<br />

of Lemma 3.1, we get<br />

F = F t1] ⊗F [t1,t2] ⊗ ...⊗F [tn−1,tn] ⊗F [tn .<br />

Every partition leads to a tensor product splitt<strong>in</strong>g of the symmetric Fock<br />

space. In this sense the symmetric Fock space is a cont<strong>in</strong>uous tensor product.<br />

Let H1 and H2 be Hilbert spaces. Denote H = H1 ⊕H2. It easily follows<br />

from the def<strong>in</strong>ition Equation (1.4) that<br />

W (x1 ⊕ x2) =W (x1) ⊗ W (x2), x1 ∈H1, x2 ∈H2. (3.2)


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 291<br />

This means that the algebra generated by the Weyl operators also splits as<br />

a cont<strong>in</strong>uous tensor product, i.e. for tn > ... > t1 > 0wehave<br />

W = W t1] ⊗W [t1,t2] ⊗ ...⊗W [tn−1,tn] ⊗W [t,<br />

where W = B(L 2 ([0, ∞))), W t] = B(L 2 ([0,t])), W [s,t] = B(L 2 ([s, t])) and<br />

W [t = B(L 2 ([t, ∞))). Furthermore, from the def<strong>in</strong>ition <strong>in</strong> Equation (1.6) it<br />

follows that<br />

Γ (S1 ⊕ S2) =Γ (S1) ⊗ Γ (S2) (3.3)<br />

for all contractions S1 ∈B(H1) andS2 ∈B(H2).<br />

Def<strong>in</strong>ition 3.2. On the dense doma<strong>in</strong> D = span{e(f); f ∈ L 2 (R + )} we<br />

<strong>in</strong>troduce annihilation At and creation At operators by<br />

At = 1<br />

�<br />

�<br />

B(iχ [0,t]) − iB(χ [0,t]) , A<br />

2<br />

∗ t<br />

1<br />

�<br />

�<br />

= B(iχ [0,t])+iB(χ [0,t]) ,<br />

2<br />

where B(f) denotes the Stone generator of W (tf). The restriction of Λt to D<br />

is also denoted by Λt andiscalledthegauge process. The operators At, A ∗ t<br />

and Λt are called the fundamental noises.<br />

It can be shown e.g. [21] that the doma<strong>in</strong> of the Stone generators B(f), f∈<br />

L2 (R + )conta<strong>in</strong>sD, thatAte(f) = 〈χ [0,t],f〉e(f) andthatA∗ t is the adjo<strong>in</strong>t<br />

of At restricted to D. Moreover, from its def<strong>in</strong>ition it follows that<br />

〈e(f),Λte(f)〉 = 〈g, χ [0,t]f〉〈e(g),e(f)〉. In particular this means that AtΦ =<br />

Ate(0) = 0 and ΛtΦ = 0, properties that we will exploit later on. Let Mt<br />

be one of the fundamental noises. It is a consequence of Equations (3.2) and<br />

(3.3) and the def<strong>in</strong>ition of the fundamental noises as (l<strong>in</strong>ear comb<strong>in</strong>ations of)<br />

generators of one-parameter groups that<br />

�<br />

�<br />

(Mt − Ms)e(f) =e(fs]) ⊗ (Mt − Ms)e(f [s,t]) ⊗ e(f [t), (3.4)<br />

where f ∈ L 2 (R + ), f s] = χ [0,s]f, f [s,t] = χ [0,t]f, f [t = χ [t,∞)f and (Mt −<br />

Ms)e(f [s,t]) ∈F [s,t]. In the follow<strong>in</strong>g we will for notational convenience often<br />

omit the tensor product signs between exponential vectors. Let H be a Hilbert<br />

space, called the <strong>in</strong>itial space. We tensor the <strong>in</strong>itial space to F and extend<br />

the operators At, A ∗ t and Λt to H⊗F by ampliation, i.e. by tensor<strong>in</strong>g the<br />

identity to them on H (however, to keep notation light we will not denote<br />

it). We denote the algebra of all bounded operators on H by B.<br />

Def<strong>in</strong>ition 3.3. (Simple quantum stochastic <strong>in</strong>tegral) Let {Ls}0≤s≤t<br />

be an adapted (i.e. Ls ∈B⊗W s] for all 0 ≤ s ≤ t) simple process with<br />

respect to the partition {s0 =0,s1,...,sp = t} <strong>in</strong> the sense that Ls = Lsj<br />

whenever sj ≤ s < sj+1. The stochastic <strong>in</strong>tegral of L with respect to a<br />

fundamental noise M on H⊗D is then def<strong>in</strong>ed as [16, 21]


292 L. Bouten<br />

� t<br />

p−1 � � ��<br />

�<br />

LsdMs xe(f) := Lsj xe(fsj ]) (Msj+1 − Msj )e(f [sj,sj+1]) e(f [sj+1 ),<br />

0<br />

j=0<br />

for all x ∈Hand f ∈ L 2 (R + ).<br />

Let {Ls}0≤s≤t be an adapted process, i.e. Ls ∈B⊗W s] for all 0 ≤ s ≤ t.<br />

We would like to def<strong>in</strong>e the <strong>in</strong>tegral<br />

It =<br />

� t<br />

0<br />

LsdMs,<br />

as a limit of simple <strong>in</strong>tegrals In t = � t<br />

0 Lns dMs where Ln is an approximation of<br />

L by simple adapted processes. In the classical case we use the Itô isometry<br />

to def<strong>in</strong>e the stochastic <strong>in</strong>tegral as a mean square limit of simple processes.<br />

Moreover, one can show that each mean square <strong>in</strong>tegrable process can be<br />

approximated by simple processes. Let us see how far we would get us<strong>in</strong>g<br />

this procedure <strong>in</strong> the quantum case. For simplicity suppose that the <strong>in</strong>itial<br />

space is trivial H = C. Suppose that we are work<strong>in</strong>g with respect to the<br />

vacuum state φ on W. Follow<strong>in</strong>g the classical reason<strong>in</strong>g, we are look<strong>in</strong>g for<br />

an operator It such that 〈(It − In t )Φ, (It − In t )Φ〉 →0asn→∞where In t are<br />

simple <strong>in</strong>tegrals correspond<strong>in</strong>g to simple approximations of L. This however,<br />

would only fix the action of It on the vacuum vector Φ. It rema<strong>in</strong>s unclear<br />

what the doma<strong>in</strong> of It is and what the action of It is on the vectors <strong>in</strong> that<br />

doma<strong>in</strong> that are not the vacuum.<br />

The solution to this problem was given by Hudson and Parthasarathy [16].<br />

They simply fix the doma<strong>in</strong> for a stochastic <strong>in</strong>tegral It to H⊗D (one could<br />

choose a dense doma<strong>in</strong> <strong>in</strong> H, for simplicity we have chosen H). Moreover, they<br />

consider I n t to be an approximation of It if 〈(It −I n t )x⊗ψ, (It −I n t<br />

)x⊗ψ〉 →0<br />

as n →∞for all x ∈Hand ψ ∈D. This limit exists if � t<br />

0 |Ls − L n s |ds →<br />

0asn → ∞ for all x ∈ H, ψ ∈ D and the limit is <strong>in</strong>dependent of the<br />

approximation [16]. Moreover, every adapted square <strong>in</strong>tegrable process L, i.e.<br />

� t<br />

0 |Lsx ⊗ ψ | 2 ds < ∞ for all x ∈H, ψ ∈Dadmits a simple approximation<br />

[16]. This means that for every adapted square <strong>in</strong>tegrable Ls we now have an<br />

unambiguous def<strong>in</strong>ition of the stochastic <strong>in</strong>tegral � t<br />

0 LsdMs, whereMs can<br />

be any of the three fundamental noises. Adapted square <strong>in</strong>tegrable processes<br />

are said to be stochastically <strong>in</strong>tegrable. We use the follow<strong>in</strong>g shorthand for<br />

stochastic <strong>in</strong>tegrals dXt = LtdMt means Xt = X0 + � t<br />

0 LsdMs.<br />

S<strong>in</strong>ce AtΦ = ΛtΦ = 0 it is immediate from the def<strong>in</strong>ition that quantum<br />

stochastic <strong>in</strong>tegrals with respect to At and Λt act<strong>in</strong>g on Φ are zero, or <strong>in</strong>f<strong>in</strong>itesimally<br />

dΛtΦ [t = dAtΦ [t = 0. From this we can immediately conclude<br />

that vacuum expectations of stochastic <strong>in</strong>tegrals with respect to At and Λt<br />

vanish. Furthermore, we have � Φ, � t<br />

0 LsdA∗ sΦ � = ��t 0 L∗sdAsΦ, Φ � = 0, i.e.<br />

are zero as<br />

vacuum expectation of stochastic <strong>in</strong>tegrals with respect to A∗ t<br />

well. Note, however, that dA∗ t Φ [t �= 0.


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 293<br />

To get some more feel<strong>in</strong>g for the def<strong>in</strong>ition of the quantum stochastic <strong>in</strong>tegral<br />

we will now <strong>in</strong>vestigate which quantum stochastic differential equation<br />

is satisfied by the Weyl operators W (f t]) (f ∈ L 2 (R + )). Note that the stochastic<br />

<strong>in</strong>tegrals are def<strong>in</strong>ed on the doma<strong>in</strong> D. Therefore we calculate for g<br />

and h <strong>in</strong> L 2 (R + )<br />

φ(t) := � e(g),W(f t])e(h) � = e −〈f t],h〉− 1<br />

2 ||f t]|| 2� e(g),e(h + f t]) �<br />

which means that<br />

φ(t) − φ(0) =<br />

= e 〈g,f t]〉−〈f t],h〉− 1<br />

2 ||f t]|| 2<br />

e 〈g,h〉 ,<br />

� t<br />

0<br />

�<br />

e(g), d<br />

�<br />

〈g, fs]〉−〈fs],h〉− ds<br />

1<br />

2 |f s] | 2�<br />

�<br />

φ(s)e(h) ds.<br />

Let us turn to the def<strong>in</strong>ition of the stochastic <strong>in</strong>tegral, Def<strong>in</strong>ition 3.3. Let<br />

{0 =s0,s1,...,sp = t} be a partition of [0,t]andchooseLs = f(sj)W (f sj])<br />

for sj ≤ s


294 L. Bouten<br />

Def<strong>in</strong>ition 3.4. Apair(L, L † ) of adapted processes def<strong>in</strong>ed on H⊗D is<br />

called an adjo<strong>in</strong>t pair if<br />

� x⊗e(f),Lty⊗e(f) � = � L †<br />

t x⊗e(f),y⊗e(f)� , x,y ∈H,f,g∈ L 2 (R + ),t≥ 0.<br />

The dagger replaces the adjo<strong>in</strong>t on the doma<strong>in</strong> H⊗D.Itiseasytosee<br />

that (A, A∗ )and(Λ, Λ) are adjo<strong>in</strong>t pairs. Moreover, if (L, L † )isanadjo<strong>in</strong>t<br />

pair, then (X, X † ) is an adjo<strong>in</strong>t pair, where dXt = LdMt and dX †<br />

t = L † dM †<br />

t .<br />

Theorem 3.5. (Quantum Itô rule[16])Let (B,B † ), (C, C † ), (D, D † ),<br />

(E,E † ) be adjo<strong>in</strong>t pairs of stochastically <strong>in</strong>tegrable processes. Let Ft,Gt,Ht<br />

and It be stochastically <strong>in</strong>tegrable processes. Let Xt and Yt be stochastic <strong>in</strong>tegrals<br />

of the form<br />

+ Etdt,<br />

dXt = BtdΛt + CtdAt + DtdA ∗ t<br />

dYt = FtdΛt + GtdAt + HtdA ∗ t + Itdt,<br />

Suppose XtYt is an adapted process def<strong>in</strong>ed on H⊗D and XF,XG,XH,XI,<br />

BY,CY,DY,EY,BF, CF,BH and CH are stochastically <strong>in</strong>tegrable, then<br />

d(XtYt) =Xt dYt +(dXt) Yt + dXt dYt,<br />

where dXtdYt should be evaluated accord<strong>in</strong>g to the quantum Itô table<br />

dAt dΛt dA∗ t dt<br />

dAt 0 dAt dt 0<br />

dΛt 0 dΛt dA∗ t 0<br />

dA∗ t 0 0 0 0<br />

dt 0 0 0 0<br />

i.e. dXtdYt = BtFtdΛt + CtFtdAt + BtHtdA ∗ t + CtHtdt.<br />

Suppose that the product XtYt is an adapted process def<strong>in</strong>ed on H⊗<br />

D. Then we can read off an expression for XtYt � from the matrix elements<br />

†<br />

X t x⊗e(f),Yty⊗e(g) � , which expla<strong>in</strong>s the need for the concept of an adjo<strong>in</strong>t<br />

pair. Writ<strong>in</strong>g out these matrix elements us<strong>in</strong>g the def<strong>in</strong>ition of the stochastic<br />

<strong>in</strong>tegral is the basic <strong>in</strong>gredient of the proof of the quantum Itô rule, see [16]<br />

and [21]. To get some more feel<strong>in</strong>g for the proof of the quantum Itô rulewe<br />

compute for x, y ∈Hand f,g ∈ L2 (R + )<br />

�<br />

xe(f),AtA ∗ t ye(g) � = � xe(f), � A ∗ t At +[At,A ∗ t ] � ye(g) � =<br />

�<br />

Atxe(f),Atye(g) � + t � xe(f),ye(g) � ��f,χ[0,t] ��<br />

= χ[0,t],g � ��xe(f),ye(g) �<br />

+ t<br />

.<br />

Inf<strong>in</strong>itesimally that reads d(AtA ∗ t )=AtdA ∗ t + A ∗ t dAt + dt. Note that the Itô<br />

correction term f<strong>in</strong>ds its orig<strong>in</strong>s <strong>in</strong> the commutator between At and A ∗ t .<br />

Note that we need to check many conditions before we are allowed to<br />

apply the quantum Itô rule. In practice though we mostly work with ‘noisy


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 295<br />

Schröd<strong>in</strong>ger equations’, which have unitary and therefore bounded solutions.<br />

If the <strong>in</strong>tegrals and coefficients are bounded, then all the requirements of the<br />

theorem are satisfied. In the rema<strong>in</strong>der of these notes we will not worry much<br />

about these issues and just assume that we can apply the quantum Itô rule.<br />

In Section 1 we encountered the classical Wiener processes B α t = ie−iα At−<br />

ie iα A ∗ t for α ∈ [0,π). S<strong>in</strong>ce dB α t = ie −iα dAt−ie iα dA ∗ t , we recover the classical<br />

Itô rule for these Wiener processes, i.e. (dB α t ) 2 = dt, fromthequantumItô<br />

rule. For an f ∈ L 2 (R + ) we can write the Weyl operator W (f t]) asW (f t]) =<br />

exp �� t<br />

0 f(s)d(A∗ s − As) � . Therefore it follows from the quantum Itô rule that<br />

the Weyl operators W (f t]) satisfy equation (3.5), where − 1<br />

2 |f |2 W (f t])dt is<br />

the Itô correction term. The Poisson process of the previous section was given<br />

by Λ f<br />

t := W (f)∗ ΛtW (f) <strong>in</strong> the vacuum state, for which the quantum Itô rule<br />

gives<br />

dΛ f<br />

t = dΛt + f(t)dAt + f(t)dA ∗ t + |f(t)| 2 dt,<br />

which leads to the classical Itô rule(dΛ f<br />

t ) 2 = dΛ f<br />

t for the Poisson process.<br />

Furthermore, <strong>in</strong>tegrat<strong>in</strong>g the above equation, we see that Λ f<br />

t = Λt +B(if t])+<br />

� t<br />

0 |f(s)|2 ds.<br />

3.3 Open Quantum Systems<br />

In quantum optics the basic model consists of some physical system of <strong>in</strong>terest,<br />

e.g. a two-level atom, a cloud of atoms, or an atom <strong>in</strong> a cavity, <strong>in</strong><br />

<strong>in</strong>teraction with the electromagnetic field. The <strong>in</strong>teraction between the electromagnetic<br />

field and the system of <strong>in</strong>terest is described by quantum electrodynamics,<br />

see e.g. [10]. The dynamics is given by a Schröd<strong>in</strong>ger equation<br />

<strong>in</strong> which the atomic dipole operator couples to a stationary Gaussian wide<br />

band noise. The next step is to approximate the wide band noise by white<br />

noise. See [1], [14] and [11] for rigorous limits implement<strong>in</strong>g this Markovian<br />

approximation. In this procedure the Schröd<strong>in</strong>ger equation given by quantum<br />

electrodynamics transforms to a quantum stochastic differential equation of<br />

the form<br />

�<br />

dUt = (S − I)dΛt + LdA ∗ t − L∗SdAt − 1<br />

2 L∗ �<br />

Ldt − iHdt Ut, U0 = I,<br />

(3.6)<br />

where S, L and H are operators on the system of <strong>in</strong>terest, H be<strong>in</strong>g selfadjo<strong>in</strong>t<br />

and S be<strong>in</strong>g unitary. See [14] for a description of how gauge terms can arise<br />

from QED Hamiltonians <strong>in</strong> a Markovian approximation. For simplicity we<br />

have considered only one channel <strong>in</strong> the field. In the examples below we will<br />

briefly describe systems that <strong>in</strong>teract with two field channels.<br />

A Picard iteration scheme shows that there exists a unique solution to<br />

Equation (3.6). Us<strong>in</strong>g the quantum Itô rule we can calculate dU ∗ t Ut =<br />

dUtU ∗ t = 0, i.e. the solution to Equation (3.6) is unitary. We def<strong>in</strong>e the


296 L. Bouten<br />

time evolution of observables of the system of <strong>in</strong>terest by the flow jt(X) =<br />

U ∗ t (X ⊗ I)Ut. Us<strong>in</strong>g the quantum Itô rule, we f<strong>in</strong>d<br />

� � � � � � � � ∗ ∗ ∗ ∗<br />

djt(X) =jt L(X) dt+jt S XS−X dΛt+jt [L ,XS] dAt+jt [S X, L] dAt ,<br />

where the L<strong>in</strong>dblad generator [18] is given by<br />

L(X) =i[H, X]+L ∗ XL − 1<br />

2 (L∗ LX − XL ∗ L).<br />

Instead of go<strong>in</strong>g through a rigorous Markov limit, we will always take the<br />

quantum stochastic differential equation (3.6) as our start<strong>in</strong>g po<strong>in</strong>t.<br />

Example 3.6. Suppose we are study<strong>in</strong>g a two-level atom <strong>in</strong> <strong>in</strong>teraction with<br />

the electromagnetic field. The two-level atom is described by the quantum<br />

probability space (M2(C),ρ) and the field is described by (W,φ). The operators<br />

S, L and H are given by<br />

S = I, L = γσ− =<br />

� 00<br />

γ 0<br />

�<br />

, H = �ω0<br />

2 σz<br />

� �ω0<br />

= 2<br />

0<br />

0 − �ω0<br />

2<br />

where � is Planck’s constant, γ ≥ 0 is a decay parameter, and ω0 ∈ R is<br />

the so-called atomic frequency, determ<strong>in</strong>ed by the fact that �ω0 is the energy<br />

difference between the two levels. That is, the <strong>in</strong>teraction between the twolevel<br />

atom and the electromagnetic field is given by the follow<strong>in</strong>g quantum<br />

stochastic differential equation (σ+ = σ ∗ − )<br />

dUt =<br />

�<br />

γσ−dA ∗ t − γσ+dAt − γ2<br />

2 σ+σ−dt − i �ω0<br />

2 σzdt<br />

�<br />

Ut, U0 = I.<br />

A laser driv<strong>in</strong>g the two-level atom could be modelled by an additional Hamiltonian<br />

Ωσx = Ω(σ− +σ+), with Ω ∈ R. A more realistic model is obta<strong>in</strong>ed by<br />

add<strong>in</strong>g an extra field channel <strong>in</strong> a coherent state (the laser). If we dist<strong>in</strong>guish<br />

two field channels, then the two-level atom and the field together are described<br />

by the quantum probability space (M2(C)⊗W1⊗W2, ρ⊗φ⊗σ) where<br />

σ is the state given by the <strong>in</strong>ner product with the vector ψ(f) =W (f)e(0),<br />

for some f ∈ L2 (R + ). Here |f| is the amplitude of the laser and the phase of f<br />

represents the phase of the driv<strong>in</strong>g laser. The quantum stochastic differential<br />

equation is given by<br />

�<br />

dUt = γ1σ−dA ∗ 1t − γ1σ+dA1t + γ2σ−dA ∗ 2t − γ2σ+dA2t −<br />

γ2 1 + γ2 2<br />

σ+σ−dt − i<br />

2<br />

�ω0<br />

2 σzdt<br />

�<br />

Ut, U0 = I,<br />

with γ1,γ2 ≥ 0. If we are only <strong>in</strong>terested <strong>in</strong> the evolution of adapted<br />

observables, then we have W (f) ∗ U ∗ t XUtW (f) = W (f t]) ∗ U ∗ t XUtW (f t]).<br />

Therefore we can describe the system by the quantum probability space<br />

�<br />

,


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 297<br />

(M2(C) ⊗W1 ⊗W2, ρ⊗ φ ⊗ φ), and <strong>in</strong>teraction given by Ũt = UtW (ft]). Us<strong>in</strong>g the quantum Itô ruleitiseasytoseethat<br />

dŨt �<br />

= γ1σ−dA ∗ 1t − γ1σ+dA1t + LdA ∗ 2t − L∗dA2t −<br />

γ2 1σ+σ− + L∗L dt − i<br />

2<br />

�<br />

2 σz<br />

�<br />

− iHdt Ũt, Ũ0 = I,<br />

with L = γ2σ− + f(t) andH = i(γ2f(t)σ+ − γ2f(t)σ−)/2. The Hamiltonian<br />

H reduces to the simpler form Ωσx if Ω = iγ2f(t)/2 andf(t) =−i.<br />

Example 3.7. We consider a two-level atom <strong>in</strong>side a cavity. The cavity has<br />

one leaky mirror which couples it to the outside field. The cavity is described<br />

by the Hilbert space ℓ2 (N). Let n ∈ N. Denotebyδnthe element of ℓ2 (N)<br />

given by δn(n) =1andδn(m) = 0 for all m ∈ N\{n}. The annihilation<br />

operator is given by bδn = √ nδn−1, n ∈ N∗ and bδ0 = 0 (we will not worry<br />

about def<strong>in</strong><strong>in</strong>g a doma<strong>in</strong> for b). The creation operator b∗ is the adjo<strong>in</strong>t of b.<br />

The two-level atom <strong>in</strong> the cavity coupled to the field can be described by the<br />

quantum probability space (M2(C) ⊗B(ℓ2 (N)) ⊗W,ρ⊗ σ ⊗ φ), where ρ is<br />

some state of the two-level atom, σ is some state on B(ℓ2 (N)) (the cavity)<br />

and φ is the vacuum state of the field. The quantum stochastic differential<br />

equation describ<strong>in</strong>g the two-level atom, the cavity, and the field together is<br />

dUt =<br />

�<br />

γbdA ∗ t − γb ∗ dAt − γ2<br />

2 b∗bdt + g(σ+b − σ−b ∗ )dt −<br />

i(∆aσz + ∆cb ∗ �<br />

b)dt Ut, U0 = I.<br />

Here γ and g are positive parameters, determ<strong>in</strong>ed by the cavity. The parameters<br />

∆a and ∆c are real. More <strong>in</strong>formation about cavity QED can be found<br />

<strong>in</strong> [10]. See [13] for rigorous results on systems <strong>in</strong>clud<strong>in</strong>g a cavity.<br />

Example 3.8. We consider a gas of atoms <strong>in</strong> free space <strong>in</strong> <strong>in</strong>teraction with a<br />

l<strong>in</strong>early polarized laser beam [25]. Suppose the atoms are all two-level atoms,<br />

i.e. the Hilbert space of the gas is given by H := C2N .HereNrepresents the<br />

number of atoms <strong>in</strong> the gas. The gas is described by the quantum probability<br />

space (B(H),ρ), where ρ is a state on the algebra B(H) of all bounded<br />

operators on H. We model the laser beam by the symmetric Fock space over<br />

the one photon space C2 ⊗ L2 (R + ). Here C2 describes the polarization and<br />

L2 (R + ) describes the spatial degree of freedom of a photon <strong>in</strong> the beam.<br />

S<strong>in</strong>ce the speed of light is constant, we can identify the spatial degree of<br />

freedom with time. Let {ex,ey} be the orthonormal basis <strong>in</strong> C2 such that ex<br />

corresponds to an x-polarized photon and ey to a y-polarized photon. We can<br />

identify C2 ⊗ L2 (R + )withL2 (R + ) ⊕ L2 (R + ) by the l<strong>in</strong>ear map that maps<br />

ex ⊗ f to (f,0) and ey ⊗ g to (0,g) for all f,g ∈ L2 (R + ). That is, the laser<br />

beam consists out of two polarization channels<br />

F(C 2 ⊗ L 2 (R + )) = Fx ⊗Fy.


298 L. Bouten<br />

Let Wx and Wy be the algebras of all bounded operators on Fx and Fy,<br />

respectively. The gas and the x-polarized laser beam together are modelled<br />

by the quantum probability space � B(H) ⊗Wx ⊗Wy,ρ⊗ σ � where σ is given<br />

by the <strong>in</strong>ner product with the coherent vector ψ(ex ⊗ f) <strong>in</strong>F(C 2 ⊗ L 2 (R + ))<br />

given by<br />

�<br />

ψ(ex ⊗ f) =exp − 1<br />

2 �f�2<br />

�<br />

1 ⊕<br />

∞�<br />

n=1<br />

(ex ⊗ f) ⊗n<br />

√ .<br />

n!<br />

The modulus and phase of f(t) correspond to the amplitude and phase of<br />

the laser at time t.<br />

We can also choose the circularly polarized basis {e+,e−} <strong>in</strong> C 2 ,givenby<br />

e+ = −(ex + iey)/ √ 2ande− =(ex − iey)/ √ 2. In an analogous way as for<br />

the l<strong>in</strong>ear basis {ex,ey} this leads to an identification<br />

F(C 2 ⊗ L 2 (R + )) = F+ ⊗F−.<br />

In a suitable appproximation [25], the <strong>in</strong>teraction between the laser beam<br />

and the gas of atoms can be described by the follow<strong>in</strong>g quantum stochastic<br />

differential equation [5]<br />

dUt =<br />

� �e iκFz − I � dΛ ++<br />

t<br />

+ � e −iκFz − I � dΛ −−<br />

�<br />

t Ut, U0 = I. (3.7)<br />

Here κ is a real coupl<strong>in</strong>g parameter. The processes Λ ++<br />

t<br />

and Λ −−<br />

t<br />

gauge processes on F+ and F−, respectively. The operator Fz : C 2N<br />

is the collective z-component of the sp<strong>in</strong> of the atoms<br />

Fz =<br />

N�<br />

i=1<br />

I ⊗i−1<br />

2<br />

⊗ σz ⊗ I ⊗N−i<br />

2 .<br />

are the<br />

→ C2N Equation (3.7) shows that right polarized photons rotate the sp<strong>in</strong> over a<br />

positive angle κ along the z-axis and left polarized photons rotate the sp<strong>in</strong><br />

over a negative angle κ along the z-axis.<br />

4 Quantum Filter<strong>in</strong>g<br />

In this section we formulate the filter<strong>in</strong>g problem and solve it by a change<br />

of measure technique that is <strong>in</strong>spired by classical filter<strong>in</strong>g theory [20, 12, 29].<br />

Quantum filter<strong>in</strong>g has been <strong>in</strong>troduced by Belavk<strong>in</strong> us<strong>in</strong>g mart<strong>in</strong>gale methods<br />

[3, 4], see also [6]. The treatment below follows [8, 9] and is based on the<br />

quantum Bayes formula and a trick <strong>in</strong>troduced by Holevo [15].


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 299<br />

4.1 The Filter<strong>in</strong>g Problem<br />

Let (B,ρ) be the quantum probability space with which we model a certa<strong>in</strong><br />

system of <strong>in</strong>terest, e.g. a two-level atom, a dilute gas of atoms, an atom <strong>in</strong><br />

a cavity, or a harmonic oscillator. The system of <strong>in</strong>terest is coupled to the<br />

electromagnetic field (W,φ) which we will take to be <strong>in</strong> the vacuum state.<br />

The comb<strong>in</strong>ed system is then described by the quantum probability space<br />

(B ⊗W, P) whereP = ρ ⊗ φ. Suppose the <strong>in</strong>teraction between the system<br />

of <strong>in</strong>terest and the electromagnetic field is given by a QSDE of the follow<strong>in</strong>g<br />

type<br />

�<br />

dUt = LdA ∗ t − L ∗ dAt − 1<br />

2 L∗ �<br />

Ldt − iHdt Ut, U0 = I, (4.1)<br />

where L and H are elements <strong>in</strong> B, H be<strong>in</strong>g selfadjo<strong>in</strong>t. For reasons of convenience<br />

we have restricted ourselves to only a s<strong>in</strong>gle channel <strong>in</strong> the field and<br />

no gauge terms <strong>in</strong> the QSDE.<br />

We will work <strong>in</strong> the Heisenberg picture, i.e. the state P rema<strong>in</strong>s fixed and<br />

the observables evolve <strong>in</strong> time accord<strong>in</strong>g to jt(S) =U ∗ t SUt where S is an<br />

element <strong>in</strong> B⊗W.LetXbe a system operator, i.e. X ∈B, then it follows<br />

from the quantum Itô rule that<br />

� � � � � � ∗ ∗<br />

djt(X) =jt L(X) dt + jt [L ,X] dAt + jt [X, L] dAt . (4.2)<br />

We will call Equation (4.2) the system.<br />

In the field we are do<strong>in</strong>g a measurement cont<strong>in</strong>uously <strong>in</strong> time. We will allow<br />

for two possible measurement setups, direct photon detection, for which the<br />

observations are given by Y Λ<br />

t = U ∗ t ΛtUt, and homodyne detection, for which<br />

the observations are given by Y W<br />

t<br />

= U ∗ t (e −iφt At+e iφt A ∗ t )Ut. For convenience<br />

we will choose φt = 0. We will not go further <strong>in</strong>to the details of the homodyne<br />

detection setup here, but <strong>in</strong>stead refer to [2]. Us<strong>in</strong>g the quantum Itô rulewe<br />

obta<strong>in</strong><br />

dY Λ<br />

t = dΛt + jt(L)dA ∗ t + jt(L ∗ )dAt + jt(L ∗ L)dt,<br />

dY W<br />

t<br />

= jt(L + L ∗ )dt + dAt + dA ∗ t .<br />

(4.3)<br />

The quantum probability space (B ⊗W, P) together with Equations (4.2)<br />

and (4.3) def<strong>in</strong>e a system-observations model, <strong>in</strong> analogy to the systemobservation<br />

models <strong>in</strong> classical stochastic control theory.<br />

Only if the observations at different times commute with each other, can we<br />

observe them <strong>in</strong> a s<strong>in</strong>gle realization of an experiment. The requirement that<br />

the observations are commutative is called the self-nondemolition property.<br />

Note that if the observations are self-nondemolition, then they can be mapped<br />

onto a classical stochastic process via the spectral theorem. This classical<br />

stochastic process can be read off from the measurement apparatus while the<br />

measurement is tak<strong>in</strong>g place cont<strong>in</strong>uously <strong>in</strong> time. Let us now check that the


300 L. Bouten<br />

observation processes Y Λ<br />

t and Y W<br />

t are <strong>in</strong>deed self-nondemolition. Let Z be<br />

an operator of the form I ⊗ Z ⊗ I [s on H⊗Fs] ⊗F [s. It follows from the<br />

quantum Itô rulesthatforallt≥s U ∗ t ZUt = U ∗ � t<br />

s ZUs+ U<br />

s<br />

∗ � t<br />

r L(Z)Urdr+ U<br />

s<br />

∗ r [Z, L]UrdA ∗ � t<br />

r+ U<br />

s<br />

∗ r [L ∗ ,Z]UrdAr.<br />

S<strong>in</strong>ce Z is a field operator and L and H are system operators, we have that<br />

L(Z) =[L, Z] =[L∗ ,Z] = 0. Therefore we have U ∗ t ZUt = U ∗ s ZUs. Us<strong>in</strong>gthis<br />

result with Z = Λs, we f<strong>in</strong>d<br />

[Y Λ<br />

t ,YΛ<br />

s ]=[U ∗ t ΛtUt,U ∗ s ΛsUs] =[U ∗ t ΛtUt,U ∗ t ΛsUt] =U ∗ t [Λt,Λs]Ut =0,<br />

whereweusedthat[Λt,Λs] = 0. In a similar way we f<strong>in</strong>d [Y W<br />

t ,Y W<br />

s ]=0.<br />

That is, both processes Y Λ and Y W are self-nondemolition. Note however<br />

that [Y Λ<br />

t ,YW<br />

t ] �= 0, we can not observe both processes <strong>in</strong> a s<strong>in</strong>gle realization<br />

of the experiment.<br />

Def<strong>in</strong>e the von Neumann algebra generated by the observations up to<br />

time t by Yt =vN{Ys; 0 ≤ s ≤ t}, whereYs is either Y Λ<br />

s <strong>in</strong> case of direct<br />

photodetection, or Y W<br />

s<br />

<strong>in</strong> case of homodyne detection. We would like to<br />

estimate jt(X) for all operators X ∈Bbased on the observations up to time<br />

t. That is, we would like to calculate the conditional expectation P(jt(X)|Yt)<br />

which is the mean least squares estimate of jt(X) givenYt. However, <strong>in</strong><br />

order for P(jt(X)|Yt) to be well-def<strong>in</strong>ed, we have to show that jt(X) is<strong>in</strong>the<br />

commutant of Yt. This requirement is called the nondemolition property. To<br />

show that jt(X) is nondemolished by Yt we aga<strong>in</strong> use that U ∗ s ZUs = U ∗ t ZUt<br />

for all operators of the form I⊗Z⊗I [s on H⊗F s]⊗F [s.IfwetakeZ = As+A ∗ s<br />

when Ys = Y W<br />

s<br />

and Z = Λs when Ys = Y Λ<br />

s , then we obta<strong>in</strong><br />

[jt(X),Ys] =[U ∗ t XUt,U ∗ s ZUs] =[U ∗ t XUt,U ∗ t ZUt] =U ∗ t [X, Z]Ut =0,<br />

where we used that [X, Z] =0s<strong>in</strong>ceXis a system operator and Z is a field<br />

operator. This establishes the nondemolition property and therefore also that<br />

P(jt(X)|Yt) is well-def<strong>in</strong>ed.<br />

We will now focus on f<strong>in</strong>d<strong>in</strong>g the filter<strong>in</strong>g equation, i.e. a recursive equation<br />

for updat<strong>in</strong>g P(jt(X)|Yt) <strong>in</strong> real time. Note that s<strong>in</strong>ce P(jt(X)|Yt) depends<br />

l<strong>in</strong>early on X and is positive and normalized, we can def<strong>in</strong>e an <strong>in</strong>formation<br />

state πt on B by<br />

πt(X) =ι � P � jt(X) � ��<br />

�Yt ,<br />

where ι maps Yt to a classical process via the spectral theorem. Note that<br />

πt is a stochastic state, it depends on the observations up to time t. The<br />

filter<strong>in</strong>g equation propagates the <strong>in</strong>formation state πt <strong>in</strong> time and is driven<br />

by the observations.


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 301<br />

4.2 Change of State<br />

Homodyne Detection<br />

Suppose that we are do<strong>in</strong>g a homodyne detection experiment and that the<br />

<strong>in</strong>teraction between the system of <strong>in</strong>terest and the field is given by Equation<br />

(4.1). Our system-observations pair is then given by the quantum probability<br />

space (B⊗W, P) and the equations<br />

� � � � � � ∗ ∗<br />

djt(X) =jt L(X) dt + jt [L ,X] dAt + jt [X, L] dAt ,<br />

dYt = jt(L + L ∗ )dt + dAt + dA ∗ t .<br />

Solv<strong>in</strong>g the correspond<strong>in</strong>g filter<strong>in</strong>g problem means to f<strong>in</strong>d a recursive equation<br />

that propagates P(jt(X)|Yt) <strong>in</strong> real time. Our strategy will be to change<br />

to a different state Rt , the so-called reference state. We will choose the reference<br />

state Rt such that the observations have <strong>in</strong>dependent <strong>in</strong>crements. That<br />

will make it easier to manipulate conditional expectations. Us<strong>in</strong>g the Bayes<br />

formula we can relate conditional expectations with respect to Rt back to<br />

conditional expectations with respect to P.<br />

Note that the process As + A ∗ s<br />

, 0 ≤ s ≤ t is a Wiener process under P,<br />

i.e. it has <strong>in</strong>dependent <strong>in</strong>crements under P. Therefore, if we def<strong>in</strong>e a reference<br />

state<br />

R t (S) =P(UtSU ∗ t ), S ∈ U ∗ t B⊗WUt,<br />

then we see that under this state the observations Ys = js(As + A∗ s) =<br />

jt(As +A∗ s ) form a Wiener process. That is, under Rt , the observation process<br />

Ys = js(As + A∗ s), 0 ≤ s ≤ t has <strong>in</strong>dependent <strong>in</strong>crements. Moreover, under<br />

Rt the system jt(X) (X ∈ B) and the observations Ys, 0 ≤ s ≤ t are<br />

<strong>in</strong>dependent. From the def<strong>in</strong>ition it is immediate that<br />

P(S) =R t (U ∗ t SUt), S ∈B⊗W.<br />

We would now like to use the Bayes formula with Q = Ut to relate conditional<br />

expectations with respect to P to conditional expectations with respect to R t .<br />

However, Ut is not <strong>in</strong> the commutant of Yt. The follow<strong>in</strong>g trick, <strong>in</strong> the spirit<br />

of [15], resolves this problem.<br />

We search for an element Qt that satisfies the follow<strong>in</strong>g two requirements<br />

1. R t (U ∗ t SUt) =R t (Q ∗ t SQt), S ∈B⊗W,<br />

2. Qtis affiliated to the commutant of Yt.<br />

(4.4)<br />

We could then use the Bayes formula with Q = Qt, toobta<strong>in</strong><br />

P � jt(X) � �<br />

�<br />

R<br />

Yt = t�Q∗ � �<br />

t jt(X)Qt�Yt<br />

� � . (4.5)<br />

Rt�Q∗ t Qt�Yt


302 L. Bouten<br />

The two conditions of Equation (4.4) are satisfied if we can f<strong>in</strong>d an element<br />

Qt that satisfies<br />

1. QtU ∗ t v ⊗ Φ = v ⊗ Φ, ∀v ∈H,<br />

2. Qt is affiliated to the commutant of Yt.<br />

Or equivalently, if we can f<strong>in</strong>d an element Vt that satisfies<br />

1. Vtv⊗Φ = Utv ⊗ Φ, ∀v ∈H,<br />

2. Vtis affiliated to the commutant of Ct,<br />

(4.6)<br />

(4.7)<br />

where Ct = UtYtU ∗ t =vN{As + A ∗ s ; 0 ≤ s ≤ t}. Indeed, if Vt satisfies the<br />

conditions <strong>in</strong> Equation (4.7), then Qt = jt(Vt) satisfies the conditions <strong>in</strong><br />

Equation (4.6) and subsequently Equation (4.4). Let Vs for 0 ≤ s ≤ t be<br />

given by<br />

dVt =<br />

�<br />

L � dA ∗ � 1<br />

t + dAt −<br />

2 L∗ �<br />

Ldt − iHdt Vt, V0 = I, (4.8)<br />

then Vtv ⊗ Φ = Utv ⊗ Φ for all v ∈H. Moreover, the equation is driven by<br />

dAt + dA ∗ t and the coefficients L and H are <strong>in</strong> the commutant of Ct, i.e. Vs<br />

is for all 0 ≤ s ≤ t affiliated to the commutant of Ct. We conclude that Vt<br />

given by Equation (4.8) satisfies the conditions <strong>in</strong> Equation (4.7).<br />

On B we def<strong>in</strong>e an unnormalized <strong>in</strong>formation state by<br />

σt(X) =R t (Q ∗ t jt(X)Qt|Yt) =R t (jt(V ∗<br />

t XVt)|Yt).<br />

From the Bayes rule Equation (4.5) it is immediate that<br />

πt(X) = σt(X)<br />

. (4.9)<br />

σt(I)<br />

This is the quantum analogue of the classical Kallianpur-Striebel formula. We<br />

will now focus on f<strong>in</strong>d<strong>in</strong>g an equation that propagates σt(X)<strong>in</strong>time.Fromthe<br />

def<strong>in</strong>ition of conditional expectations it easily follows that Rt (jt(V ∗<br />

t XVt)|Yt)<br />

= U ∗ t P(V ∗<br />

t XVt|Ct)Ut. S<strong>in</strong>ceVtis driven by classical noise, we can resort to<br />

the classical Itô calculus from here onwards. The Itô rulegives<br />

dV ∗<br />

t XVt = V ∗<br />

t L(X)Vtdt + V ∗<br />

t (L∗ X + XL)Vt(dAt + dA ∗ t ).<br />

In <strong>in</strong>tegral form this reads<br />

V ∗<br />

t XVt = X +<br />

� t<br />

0<br />

V ∗<br />

s L(X)Vsds +<br />

� t<br />

0<br />

V ∗<br />

s (L∗ X + XL)Vs(dAs + dA ∗ s ).<br />

Tak<strong>in</strong>g the conditional expectation P( ·|Ct) of this equation gives


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 303<br />

P � V ∗<br />

t XVt<br />

�<br />

� Ct<br />

0<br />

� t<br />

� � � �<br />

= P X�Ct + P<br />

0<br />

� V ∗<br />

s L(X)Vs<br />

� �<br />

�Ct ds +<br />

� t<br />

P � V ∗<br />

s (L∗ � �<br />

X + XL)Vs�Ct<br />

(dAs + dA ∗ s ),<br />

where we have approximated the <strong>in</strong>tegrals by simple functions <strong>in</strong> the usual<br />

way and have taken the <strong>in</strong>tegrators out us<strong>in</strong>g the module property of the<br />

conditional expectation. For adapted processes Ls we have that P(Ls|Ct) =<br />

P(Ls|Cs) for all s ≤ t. This is where our clever choice for the reference state Rt pays off. We use here that C [s,t] is <strong>in</strong>dependent of Cs under P, orequivalently,<br />

that Y [s,t] is <strong>in</strong>dependent of Ys under Rt . Note that that is exactly how we<br />

had chosen the reference state Rt . This means we now have<br />

P � V ∗<br />

�<br />

� � t<br />

t XVt�Ct<br />

= P(X)+ P<br />

0<br />

� V ∗<br />

� �<br />

s L(X)Vs�Cs<br />

ds +<br />

� t<br />

P � V ∗<br />

s (L∗ � �<br />

X + XL)Vs�Cs<br />

(dAs + dA ∗ s ).<br />

0<br />

Sandwich<strong>in</strong>g with Ut leads to the follow<strong>in</strong>g l<strong>in</strong>ear quantum filter<strong>in</strong>g equation<br />

� �<br />

dσt(X) =σt L(X) dt + σt(L ∗ X + XL)dYt.<br />

This equation is the quantum analogue of the classical Duncan-Mortensen-<br />

Zakai equation [20, 12, 29]. Us<strong>in</strong>g the Kallianpur-Striebel formula Equation<br />

(4.9) and the Itô rule, we obta<strong>in</strong> the normalized quantum filter<strong>in</strong>g equation<br />

� � �<br />

dπt(X) =πt L(X) dt+ πt(L ∗ X+XL)−πt(L ∗ ��<br />

+L)πt(X) dYt−πt(L ∗ �<br />

+L)dt .<br />

(4.10)<br />

This equation is a quantum analogue of the classical Kushner-Stratonovich<br />

equation. The process dYt − πt(L∗ + L)dt driv<strong>in</strong>g the quantum filter is called<br />

the <strong>in</strong>novations or the <strong>in</strong>novat<strong>in</strong>g mart<strong>in</strong>gale. In subsection 4.3 we will prove<br />

that the <strong>in</strong>novations <strong>in</strong>deed form a mart<strong>in</strong>gale.<br />

Direct Photodetection<br />

Now suppose that <strong>in</strong>stead of a homodyne detection experiment we are<br />

directly count<strong>in</strong>g photons <strong>in</strong> the field. Our system-observations pair is then<br />

given by the quantum probability space (B ⊗W, P) and the equations<br />

� � � � � � ∗ ∗<br />

djt(X) =jt L(X) dt + jt [L ,X] dAt + jt [X, L] dAt ,<br />

dYt = dΛt + jt(L)dA ∗ t + jt(L ∗ )dAt + jt(L ∗ L)dt.<br />

To solve the correspond<strong>in</strong>g filter<strong>in</strong>g problem, we aga<strong>in</strong> change to a different<br />

state R t , the so-called reference state for the count<strong>in</strong>g experiment. We<br />

choose the reference state such that the observation process has <strong>in</strong>dependent<br />

<strong>in</strong>crements. Moreover, we want the measure <strong>in</strong>duced on the observations Yt<br />

by the reference measure R t to be absolutely cont<strong>in</strong>uous with respect to the<br />

measure <strong>in</strong>duced by P. This last requirement rules out the reference state we


304 L. Bouten<br />

used previously for the homodyne case, s<strong>in</strong>ce under this measure the count<strong>in</strong>g<br />

observations will be identical zero, whereas under P this is not the case. If<br />

we would ignore this problem and would proceed naively with the reference<br />

state from the previous section, we would f<strong>in</strong>d that we are unable to f<strong>in</strong>d a<br />

suitable Radon-Nikodym derivative Qt.<br />

We solve the above difficulty by us<strong>in</strong>g a reference state that turns the<br />

observations <strong>in</strong>to a Poisson process. Let the Weyl operator Wt be given by<br />

�<br />

dWt = dAt − dA ∗ 1<br />

t −<br />

2 dt<br />

�<br />

Wt, W0 = I.<br />

satisfies the follow<strong>in</strong>g QSDE<br />

�<br />

(L − I)dA ∗ t − (L∗ − I)dAt − 1<br />

2 (L∗ �<br />

L + I − 2L +2iH)dt<br />

Def<strong>in</strong>e U ′ t = WtUt, thenU ′ t<br />

dU ′ t =<br />

Let the reference state R t be given by<br />

R t (S) =P(U ′ tSU ′ ∗ ′ ∗ ′<br />

t ), S ∈ U t B⊗WU t .<br />

, U ′ 0<br />

= I.<br />

(4.11)<br />

Note that under R t we have that the process Ys, 0 ≤ s ≤ t is a Poisson<br />

process, i.e. the observations have <strong>in</strong>dependent <strong>in</strong>crements. Moreover, note<br />

that under R t jt(X) andYs = js(Λs) =jt(Λs) are <strong>in</strong>dependent for all X ∈B.<br />

From the def<strong>in</strong>ition it is immediate that<br />

P(S) =R t (U ′ ∗ ′<br />

t SU t ), S ∈B⊗W.<br />

As before we would now like to apply the Bayes formula. However, Equation<br />

(4.11) shows that U ′ t is not <strong>in</strong> the commutant of Yt. We will use the same<br />

trick as before to solve this problem [15].<br />

We search for an element Qt that satisfies the follow<strong>in</strong>g two requirements<br />

1. R t (U ′ ∗ ′<br />

t SU t )=R t (Q ∗ t SQt), S ∈B⊗W,<br />

2. Qtis affiliated to the commutant of Yt.<br />

These conditions are satisfied if we can f<strong>in</strong>d an element Qt that satisfies<br />

1. QtU ′ ∗<br />

t v ⊗ Φ = v ⊗ Φ, ∀v ∈H,<br />

2. Qtis affiliated to the commutant of Yt.<br />

Or equivalently, if we can f<strong>in</strong>d an element Vt that satisfies<br />

1. V ′<br />

t v ⊗ Φ = U ′ tv ⊗ Φ, ∀v ∈H,<br />

2. V ′<br />

t<br />

where Ct = U ′ tYtU ′ ∗<br />

t =vN{WsΛsW∗ s<br />

satisfy<strong>in</strong>g Equation (4.14), then Qt = U ′ t<br />

is affiliated to the commutant of Ct,<br />

(4.12)<br />

(4.13)<br />

(4.14)<br />

′<br />

;0≤ s ≤ t}. Indeed,ifwecanf<strong>in</strong>daV t<br />

satisfies Equations (4.13) and<br />

∗ V ′<br />

t U ′ t


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 305<br />

(4.12). Note that Zt = WtΛtW ∗ t is given by<br />

Let V ′<br />

s<br />

dZt = dΛt + dAt + dA ∗ t + dt, Z0 =0.<br />

for 0 ≤ s ≤ t be given by<br />

dV ′<br />

t =<br />

then V ′<br />

t v ⊗ Φ = U ′ t<br />

�<br />

(L − 1)dZt − 1<br />

2 (L∗ L + I +2iH)dt<br />

�<br />

V ′<br />

t , V0 = I, (4.15)<br />

v ⊗ Φ for all v ∈H. Moreover, the equation is driven<br />

by dZt and the coefficients L and H are <strong>in</strong> the commutant of Ct, i.e. V ′<br />

s is<br />

affiliated to the commutant of Ct for all 0 ≤ s ≤ t. This means that V ′<br />

t given<br />

by Equation (4.15) satisfies the conditions <strong>in</strong> Equation (4.14).<br />

We can now apply the Bayes formula with Qt = U ′ ∗ ′<br />

t V t U ′ t , i.e.<br />

P � jt(X) � � R<br />

�Yt = t�Q∗ � �<br />

t jt(X)Qt�Yt<br />

Rt (Q∗ t Qt|Yt)<br />

, X ∈B.<br />

On B we def<strong>in</strong>e an unnormalized <strong>in</strong>formation state by<br />

σt(X) =R t� Q ∗ � �<br />

t jt(X)Qt�<br />

t<br />

Yt = R � U ′ t<br />

∗ ′ ∗ ′<br />

V t XV t U ′ �<br />

t Yt<br />

where <strong>in</strong> the second step we used that X ∈Bcommutes with Wt. Fromthe<br />

Bayes formula it is immediate that<br />

πt(X) = σt(X)<br />

, X ∈B. (4.16)<br />

σt(I)<br />

We will now focus on f<strong>in</strong>d<strong>in</strong>g an equation that propagates σt(X) <strong>in</strong>time.<br />

From the def<strong>in</strong>ition of the conditional expectation it easily follows that<br />

Rt�U ′ ∗ ′ ∗ ′<br />

t V t XV t U ′ � �<br />

�Yt<br />

′ ∗ ′ ∗ ′<br />

t = U t P(V t XV t |Ct)U ′ t .TheItôrulegives<br />

dV ′ ∗ ′<br />

t XV t = V ′ ∗ ′<br />

t L(X)V t dt + V ′<br />

t (L ∗ XL − X)V ′<br />

�<br />

� ,<br />

t (dZt − dt).<br />

Us<strong>in</strong>g analogous arguments as <strong>in</strong> the homodyne case, we obta<strong>in</strong> the l<strong>in</strong>ear<br />

quantum filter<br />

� � �<br />

σt(X) =σt L(X) dt + σt(L ∗ XL) − σt(X) �� dYt − dt � .<br />

Us<strong>in</strong>g Equation (4.16) and the Itô rule, we obta<strong>in</strong> the normalized quantum<br />

filter for photon count<strong>in</strong>g<br />

�<br />

� � πt(L<br />

dπt(X) =πt L(X) dt +<br />

∗XL) πt(L∗ �<br />

�dYt<br />

− πt(X) − πt(L<br />

L) ∗ L)dt � .<br />

The process dYt − πt(L ∗ L)dt is called the <strong>in</strong>novations or the <strong>in</strong>novat<strong>in</strong>g mart<strong>in</strong>gale<br />

for the photon count<strong>in</strong>g experiment. In the next subsection we prove<br />

that this process <strong>in</strong>deed forms a mart<strong>in</strong>gale.


306 L. Bouten<br />

4.3 Innovations<br />

In mart<strong>in</strong>gale based approaches to quantum filter<strong>in</strong>g [4],[6], the follow<strong>in</strong>g theorem<br />

is the start<strong>in</strong>g po<strong>in</strong>t. The proof below is an adaptation to the Heisenberg<br />

picture of the proof <strong>in</strong> [6]. It can also be found <strong>in</strong> [8], [7].<br />

Theorem 4.1. Let Yt be given by Yt = U ∗ t ZtUt, wheredZt = atdΛt +btdAt +<br />

btdA ∗ t , Z0 =0and at ∈ R and bt ∈ C for all t ≥ 0. DenotebyYt the von<br />

Neumann algebra generated by Ys for 0 ≤ s ≤ t. Def<strong>in</strong>e the <strong>in</strong>novations ˜ Yt by<br />

˜Yt = Yt −<br />

� t<br />

0<br />

�<br />

P<br />

asU ∗ s L ∗ LUs + bsU ∗ s LUs + bsU ∗ s L ∗ Us�Ys<br />

�<br />

�<br />

ds.<br />

The <strong>in</strong>novations are a mart<strong>in</strong>gale, i.e. P( ˜ Yt|Ys) = ˜ Ys for all 0 ≤ s ≤ t.<br />

Proof. We have to show that P( ˜ Yt − ˜ Ys|Ys) = 0 for all 0 ≤ s ≤ t. This means<br />

we need to prove that P � ( ˜ Yt − ˜ Ys)K � =0forall0≤s≤t and K ∈Ys. This<br />

is equivalent to show<strong>in</strong>g that<br />

� t �<br />

P(YtK) − P(YsK) = P P(arU ∗ r L∗LUr + brU ∗ r LUr + brU ∗ r L∗ �<br />

Ur|Yr)K dr,<br />

s<br />

for all 0 ≤ s ≤ t and for all K ∈Ys. Fort = s the above equation is obviously<br />

true. It therefore rema<strong>in</strong>s to show that for all K ∈Ys and all s ≤ r ≤ t<br />

�<br />

dP(YrK) =P P(arU ∗ r L ∗ LUr + brU ∗ r LUr + brU ∗ r L ∗ �<br />

Ur|Yr)K dr<br />

�<br />

= P arU ∗ r L∗LUrK + brU ∗ r LUrK + brU ∗ r L∗ �<br />

UrK dr.<br />

This is just a simple exercise <strong>in</strong> apply<strong>in</strong>g the quantum Itô rule and the relation<br />

Yr = U ∗ r ZrUr.<br />

Acknowledgment. I would like to thank Mart<strong>in</strong> L<strong>in</strong>dsay and Uwe Franz<br />

for a critical read<strong>in</strong>g of a first version of this text.<br />

References<br />

1. L. Accardi, A. Frigerio, and Y. Lu. The weak coupl<strong>in</strong>g limit as a quantum functional<br />

central limit. Commun. Math. Phys., 131:537–570, 1990.<br />

2. A. Barchielli. Direct and heterodyne detection and other applications of quantum<br />

stochastic calculus to quantum optics. Quantum Opt., 2:423–441, 1990.<br />

3. V. P. Belavk<strong>in</strong>. Nondemolition stochastic calculus <strong>in</strong> Fock space and nonl<strong>in</strong>ear filter<strong>in</strong>g<br />

and control <strong>in</strong> quantum systems. In R. Guelerak and W. Karwowski, editors, Proceed<strong>in</strong>gs<br />

XXIV Karpacz w<strong>in</strong>ter school, Stochastic methods <strong>in</strong> mathematics and physics,<br />

pages 310–324. World Scientific, S<strong>in</strong>gapore, 1988.<br />

4. V. P. Belavk<strong>in</strong>. Quantum stochastic calculus and quantum nonl<strong>in</strong>ear filter<strong>in</strong>g. J. Multivar.<br />

Anal., 42:171–201, 1992.


Applications of Quantum Stochastic Processes <strong>in</strong> Quantum Optics 307<br />

5. L. Bouten, J. Stockton, G. Sarma, and H. Mabuchi. Scatter<strong>in</strong>g of polarized laser light<br />

by an atomic gas <strong>in</strong> free space: a quantum stochastic differential equation approach.<br />

Phys. Rev. A, 75:052111, 2007.<br />

6. L. M. Bouten, M. I. Gut¸ă, and H. Maassen. Stochastic Schröd<strong>in</strong>ger equations. J. Phys.<br />

A, 37:3189–3209, 2004.<br />

7. L. M. Bouten and R. Van Handel. On the separation pr<strong>in</strong>ciple of quantum control,<br />

2005. math-ph/0511021.<br />

8. L. M. Bouten and R. Van Handel. Quantum filter<strong>in</strong>g: a reference probability approach,<br />

2005. math-ph/0508006.<br />

9. L. M. Bouten, R. Van Handel, and M. James. An <strong>in</strong>troduction to quantum filter<strong>in</strong>g.<br />

SIAM J. Control Optim., 46:2199–2241, 2007.<br />

10. C. Cohen Tannoudji, J. Dupont Roc, and G. Grynberg. Photons and Atoms: Introduction<br />

to Quantum Electrodynamics. Wiley, 1989.<br />

11. J. Derez<strong>in</strong>ski and W. De Roeck. Extended weak coupl<strong>in</strong>g limit for Pauli-Fierz operators,<br />

2006. math-ph/0610054.<br />

12. T. Duncan. Evaluation of likelihood functions. Information and Control, (13):62–74,<br />

1968.<br />

13. F. Fagnola. On quantum stochastic differential equations with unbounded coefficients.<br />

Probab. Th. Rel. Fields, 86:501–516, 1990.<br />

14. J. Gough. Quantum flows as Markovian limit of emission, absorption and scatter<strong>in</strong>g<br />

<strong>in</strong>teractions. Commun. Math. Phys., 254:489–512, 2005.<br />

15. A. Holevo. Quantum stochastic calculus. J. Soviet Math., 56:2609–2624, 1991. Translation<br />

of Itogi Nauki i Tekhniki, ser. sovr. prob. mat. 36, 3–28, 1990.<br />

16. R. L. Hudson and K. R. Parthasarathy. Quantum Itô’s formula and stochastic evolutions.<br />

Commun. Math. Phys., 93:301–323, 1984.<br />

17. R. V. Kadison and J. R. R<strong>in</strong>grose. Fundamentals of the <strong>Theory</strong> of Operator Algebras,<br />

volume I. Academic Press, San Diego, 1983.<br />

18. G. L<strong>in</strong>dblad. On the generators of quantum dynamical semigroups. Commun. Math.<br />

Phys., 48:119–130, 1976.<br />

19. H. Maassen. Quantum probability applied to the damped harmonic oscillator. In<br />

S. Attal and J. L<strong>in</strong>dsay, editors, Quantum <strong>Probability</strong> Communications, volumeXII,<br />

pages 23–58. World Scientific, S<strong>in</strong>gapore, 2003.<br />

20. R. Mortensen. Optimal Control of Cont<strong>in</strong>uous-Time Stochastic Systems. PhDthesis,<br />

Univ. California, Berkeley, 1966.<br />

21. K. R. Parthasarathy. An Introduction to Quantum Stochastic Calculus. Birkhäuser,<br />

Basel, 1992.<br />

22. D. Petz. An Invitation to the Algebra of Canonical Commutation Relations. Leuven<br />

University Press, Leuven, 1990.<br />

23. M. Reed and B. Simon. Functional Analysis, volume 1 of Methods of Modern Mathematical<br />

Physics. Elsevier, 1980.<br />

24. I. Segal. Tensor algebras over Hilbert spaces. I. Trans. Am. Math. Soc., 81:106–134,<br />

1956.<br />

25. J. K. Stockton. Cont<strong>in</strong>uous quantum measurement of cold alkali-atom sp<strong>in</strong>s. PhD<br />

thesis, California Institute of Technology, 2006.<br />

26. M. Takesaki. Conditional expectations <strong>in</strong> von Neumann algebras. J. Funct. Anal.,<br />

9:306–321, 1971.<br />

27. J. von Neumann. Zur Algebra der Funktionaloperationen und Theorie der normalen<br />

Operatoren. Math. Ann., 102:370–427, 1929.<br />

28. D. Williams. <strong>Probability</strong> with Mart<strong>in</strong>gales. Cambridge University Press, Cambridge,<br />

1991.<br />

29. M. Zakai. On the optimal filter<strong>in</strong>g of diffusion processes. Z. Wahrsch. Verw. Geb.,<br />

11:230–243, 1969.


Quantum Walks<br />

Norio Konno<br />

Summary. Quantum walks can be considered as a generalized version of<br />

the classical random walk. There are two classes of quantum walks, that is,<br />

the discrete-time (or co<strong>in</strong>ed) and the cont<strong>in</strong>uous-time quantum walks. This<br />

manuscript treats the discrete case <strong>in</strong> Part I and cont<strong>in</strong>uous case <strong>in</strong> Part II,<br />

respectively. Most of the contents are based on our results. Furthermore, papers<br />

on quantum walks are listed <strong>in</strong> References. Studies of discrete-time walks<br />

appeared from the late 1980s from Gudder (1988), for example. Meyer (1996)<br />

<strong>in</strong>vestigated the model as a quantum lattice gas automaton. Nayak and Vishwanath<br />

(2000) and Amba<strong>in</strong>is et al. (2001) studied <strong>in</strong>tensively the behaviour<br />

of discrete-time walks, <strong>in</strong> particular, the Hadamard walk. In contrast with the<br />

central limit theorem for the classical random walks, Konno (2002a, 2005a)<br />

showed a new type of weak limit theorems for the one-dimensional lattice.<br />

Grimmett, Janson, and Scudo (2004) extended the limit theorem to a wider<br />

range of the walks. On the other hand, the cont<strong>in</strong>uous-time quantum walk<br />

was <strong>in</strong>troduced and studied by Childs, Farhi, and Gutmann (2002). Excellent<br />

reviews on quantum walks are found <strong>in</strong> Kempe (2003), Tregenna et al.<br />

(2003), Amba<strong>in</strong>is (2003), Kendon (2007).<br />

Acknowledgement<br />

A Japanese version of these notes with additional content, exercises and diagrams<br />

has recently appeared <strong>in</strong> Konno (2008). We thank Sangyo Tosho for<br />

grant<strong>in</strong>g permission to <strong>in</strong>clude here material from Konno (2008).<br />

N. Konno<br />

Department of Applied Mathematics, Yokohama National University,<br />

79-5 Tokiwadai, Yokohama 240-8501, Japan<br />

e-mail: konno@ynu.ac.jp<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 309<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


Part I: Discrete-Time Quantum Walks<br />

1 Limit Theorems<br />

1.1 Two-State Case<br />

Introduction<br />

The classical random walk on the l<strong>in</strong>e is the motion of a particle that lives<br />

on the set of <strong>in</strong>tegers. The particle moves at each time step either one unit<br />

to the right with probability p or one unit to the left with probability 1 − p.<br />

The directions of different steps are <strong>in</strong>dependent of each other. This classical<br />

random walk is often called the Bernoulli random walk. For general random<br />

walks on a countable space, there is a beautiful theory (see Spitzer (1976)).<br />

In this section, we consider quantum variations of the Bernoulli random walk<br />

and refer to such processes as quantum walks here.<br />

The quantum walk considered here is determ<strong>in</strong>ed by a 2 × 2 unitary matrix<br />

U that will be <strong>in</strong>troduced below. We will <strong>in</strong>troduce four other matrices<br />

P, Q, R and S which will all be determ<strong>in</strong>ed by U. The matrices P, Q, R and<br />

S will allow us to obta<strong>in</strong> a comb<strong>in</strong>atorial expression for the characteristic<br />

function of the walk and will clarify the dependence of the m-th moment and<br />

the symmetry of the distribution of the walk on the unitary matrix U and the<br />

<strong>in</strong>itial qubit state ϕ. Furthermore we give a new type of weak limit theorems<br />

for the quantum walk by us<strong>in</strong>g our results. Our limit theorem shows that the<br />

behavior of quantum walk is remarkably different from that of the classical<br />

random walk. As a corollary, it reveals whether some simulation results already<br />

known are accurate or not. The results on this section are based on<br />

Konno (2002a, 2005a).<br />

The time evolution of the one-dimensional quantum walk studied here is<br />

given by the follow<strong>in</strong>g unitary matrix:<br />

� �<br />

ab<br />

U = ∈ U(2)<br />

cd<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 311<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


312 N. Konno<br />

where a, b, c, d ∈ C. HereC is the set of complex numbers, and U(2) is the<br />

set of 2 × 2 unitary matrices. So we have<br />

|a| 2 + |c| 2 = |b| 2 + |d| 2 =1, ac + bd =0,<br />

c = −△b, d = △a,<br />

where z is a complex conjugate of z ∈ C and △ =detU = ad−bc. We should<br />

note that the unitarity of U gives |△| =1.<br />

The quantum walk is a quantum generalization of the classical Bernoulli<br />

random walk <strong>in</strong> one dimension with an additional degree of freedom called<br />

the chirality. The walk can also be considered as a quantum version of the socalled<br />

correlated random walk. There is a strong structural similarity between<br />

them (see Konno (2003)). The chirality takes values left and right, andmeans<br />

the direction of the motion of the particle. The evolution of the quantum walk<br />

is given by the follow<strong>in</strong>g rule. At each time step, if the particle has the left<br />

chirality, it moves one unit to the left, and if it has the right chirality, it<br />

moves one unit to the right.<br />

More precisely, the unitary matrix U acts on two chirality states |L〉<br />

and |R〉:<br />

|L〉 → a|L〉 + c|R〉, |R〉 → b|L〉 + d|R〉,<br />

where L and R refer to the right and left chirality state respectively. In fact,<br />

def<strong>in</strong>e<br />

� �<br />

� �<br />

1<br />

0<br />

|L〉 = , |R〉 = ,<br />

0<br />

1<br />

so we have<br />

U|L〉 = a|L〉 + c|R〉, U|R〉 = b|L〉 + d|R〉.<br />

The study of the dependence of some important quantities (e.g., characteristic<br />

function, the m-th moment, limit theorem) on <strong>in</strong>itial qubit state is<br />

one of the essential parts, so we def<strong>in</strong>e the set of <strong>in</strong>itial qubit states as follows:<br />

Φ =<br />

�<br />

ϕ =<br />

� α<br />

β<br />

�<br />

∈ C 2 : |α| 2 + |β| 2 �<br />

=1 .<br />

Let Xn be the quantum walk at time n start<strong>in</strong>g from <strong>in</strong>itial qubit state<br />

ϕ ∈ Φ. It should be noted that P (Xn =0)=1. In contrast with classical<br />

random walks, Xn cannotbewrittenasXn = Y1 + ···+ Yn, where Y1,Y2,...<br />

are <strong>in</strong>dependent and identically distributed random variables. The precise<br />

def<strong>in</strong>ition of Xn will appear <strong>in</strong> the next subsection. Here we expla<strong>in</strong> Xn<br />

briefly. To do so, we <strong>in</strong>troduce the follow<strong>in</strong>g matrices P and Q:<br />

P =<br />

� ab<br />

00<br />

�<br />

, Q =<br />

� 00<br />

cd<br />

�<br />

.


Quantum Walks 313<br />

Remark that<br />

U = P + Q.<br />

For fixed l and m with l + m = n and −l + m = x,<br />

Ξn(l, m) = �<br />

lj,mj<br />

P l1 Q m1 P l2 Q m2 ···P ln Q mn<br />

summed over all lj,mj ≥ 0 satisfy<strong>in</strong>g l1 +···+ln = l, m1 +···+mn = m and<br />

lj + mj = 1. Moreover, to def<strong>in</strong>e P (Xn = x), it is convenient to <strong>in</strong>troduce<br />

R =<br />

� cd<br />

00<br />

�<br />

, S =<br />

� 00<br />

ab<br />

An expression of Ξn(l, m) can be given by us<strong>in</strong>g P, Q, R and S (see Lemma<br />

1.1). The def<strong>in</strong>ition of Ξn(l, m) gives<br />

P (Xn = x) =(Ξn(l, m)ϕ) ∗ (Ξn(l, m)ϕ) =�Ξn(l, m)ϕ� 2 ,<br />

where n = l + m, x = −l + m and ∗ means the adjo<strong>in</strong>t operator. From this<br />

expression, we will obta<strong>in</strong> the characteristic function of Xn (Theorem 1.3)<br />

and the m-th moment of it (Corollary 1.4). One of the <strong>in</strong>terest<strong>in</strong>g results is<br />

that when m is even, the m-th moment of Xn is <strong>in</strong>dependent of the <strong>in</strong>itial<br />

qubit state ϕ ∈ Φ. On the other hand, when m is odd, the m-th moment<br />

depends on the <strong>in</strong>itial qubit state. So the standard deviation of Xn is not<br />

<strong>in</strong>dependent of the <strong>in</strong>itial qubit state ϕ ∈ Φ.<br />

By us<strong>in</strong>g the characteristic function of Xn, we give the follow<strong>in</strong>g new type<br />

of limit theorems <strong>in</strong> abcd �= 0 case (Theorem 1.6). If n →∞,then<br />

where Z has a density<br />

f(x) =f(x; ϕ = T [α, β])<br />

�<br />

1 −|a| 2<br />

=<br />

π(1 − x2 ) � |a| 2 − x2 Xn<br />

n<br />

⇒ Z,<br />

�<br />

.<br />

� �<br />

1 − |α| 2 −|β| 2 +<br />

aαbβ + aαbβ<br />

|a| 2<br />

� �<br />

x ,<br />

for x ∈ (−|a|, |a|), and f(x) =0for|x| ≥|a|, whereYn ⇒ Y means that Yn<br />

converges weakly to a limit Y and T <strong>in</strong>dicates the transposed operator. We<br />

remark that standard deviation of Z is not <strong>in</strong>dependent of <strong>in</strong>itial qubit state<br />

ϕ = T [α, β].<br />

Moreover if b =0,thenc =0and|a| = |d| =1.Sothiscaseistrivial.In<br />

fact, Theorem 1.3 (ii) implies that Xn/n ⇒ W (1) , where W (1) is determ<strong>in</strong>ed<br />

by P (W (1) = −1) = |α| 2 and P (W (1) =1)=|β| 2 .Ifa =0,thend =0<br />

and |b| = |c| = 1. So this case is also trivial. Theorem 1.3 (iii) implies that<br />

Xn/n ⇒ W (2) , where W (2) is determ<strong>in</strong>ed by δ0(x), where δa(x) denotes the<br />

po<strong>in</strong>tmass at a.


314 N. Konno<br />

The above weak limit theorem suggests the follow<strong>in</strong>g result on the symmetry<br />

of the distribution for quantum walks (Theorem 1.5). Def<strong>in</strong>e<br />

Φs = {ϕ ∈ Φ : P (Xn = x) =P (Xn = −x) for any n ∈ Z+ and x ∈ Z},<br />

Φ0 = {ϕ ∈ Φ : E(Xn) = 0 for any n ∈ Z+} ,<br />

Φ⊥ = � ϕ = T [α, β] ∈ Φ : |α| = |β|, aαbβ + aαbβ =0 � ,<br />

and Z (resp. Z+) is the set of (resp. non-negative) <strong>in</strong>tegers. For ϕ ∈ Φs, the<br />

probability distribution of Xn is symmetric for any n ∈ Z+. Us<strong>in</strong>g explicit<br />

forms of distribution of Xn (Lemma 1.2) and E(Xn) (Corollary 1.4 (i) for<br />

m =1case),wehave<br />

Φs = Φ0 = Φ⊥.<br />

Nayak and Vishwanath (2000) discussed the symmetry of distribution and<br />

showed that T [1/ √ 2, ±i/ √ 2] ∈ Φs for the Hadamard walk.<br />

In the rest of this subsection, we focus on the Hadamard walk which has<br />

been extensively <strong>in</strong>vestigated <strong>in</strong> the study of quantum walks. The unitary<br />

matrix U of the Hadamard walk is def<strong>in</strong>ed by the follow<strong>in</strong>g Hadamard gate<br />

(see Nielsen and Chuang (2000)):<br />

U = 1 � �<br />

1 1<br />

√ .<br />

2 1 −1<br />

The dynamics of this walk corresponds to that of the symmetric random<br />

walk <strong>in</strong> the classical case. However the symmetry of the Hadamard walk<br />

depends heavily on <strong>in</strong>itial qubit state, see Konno, Namiki, and Soshi (2004).<br />

For example, <strong>in</strong> the case of the Hadamard walk with <strong>in</strong>itial qubit state<br />

ϕ = T [1/ √ 2,i/ √ 2] (symmetric case), direct computation gives<br />

P (X4 = −4) = P (X4 =4)=1/16,<br />

P (X4 = −2) = P (X4 =2)=6/16, P(X4 =0)=2/16.<br />

In contrast with the above result, as for the classical symmetric random walk<br />

Y o n start<strong>in</strong>g from the orig<strong>in</strong>, we see that<br />

P (Y o<br />

4<br />

= −4) = P (Y o<br />

4 =4)=1/16,<br />

P (Y o<br />

4 = −2) = P (Y o<br />

4 =2)=4/16, P(Y o<br />

4 =0)=6/16.<br />

In fact, quantum walks behave quite differently from classical random walks.<br />

For the classical walk, the probability distribution is a b<strong>in</strong>omial distribution.<br />

On the other hand, the probability distribution <strong>in</strong> the quantum walk has a<br />

complicated, oscillatory form.<br />

Now we compare our analytical results (Theorem 1.6) with the numerical<br />

ones given by Mackay et al. (2002), Travaglione and Milburn (2002) for the<br />

the Hadamard walk. In this case, our weak limit theorem (Theorem 1.6)<br />

implies that if −1/ √ 2 ≤ a


Quantum Walks 315<br />

P (a ≤ Xn/n ≤ b) →<br />

� b<br />

1 − (|α|<br />

a<br />

2 −|β| 2 + αβ + αβ)x<br />

π(1 − x2 ) √ 1 − 2x2 for any <strong>in</strong>itial qubit state ϕ = T [α, β]. For the classical symmetric random<br />

walk Y o n start<strong>in</strong>g from the orig<strong>in</strong>, the well-known central limit theorem implies<br />

that if −∞


316 N. Konno<br />

In another asymmetric case ϕ = T [e iθ , 0] where θ ∈ [0, 2π), a similar<br />

argument implies that if −1/ √ 2 ≤ a


Quantum Walks 317<br />

where<br />

0=<br />

� �<br />

0<br />

, ϕ =<br />

0<br />

� �<br />

α<br />

.<br />

β<br />

The follow<strong>in</strong>g equation def<strong>in</strong>es the time evolution of the quantum walk:<br />

Ψn+1(x) =(U N Ψn)x = PΨn(x +1)+QΨn(x − 1),<br />

where (Ψn)x = Ψn(x), −N ≤ x ≤ N and 1 ≤ n


318 N. Konno<br />

Let |x〉 ∈Hp(x ∈ Z) be the position states of the quantum walk. A unitary<br />

shift operator is def<strong>in</strong>ed by<br />

ˆS = �<br />

�<br />

|x +1〉〈x|, Sˆ −1<br />

= Sˆ ∗<br />

= |x − 1〉〈x|.<br />

Then<br />

x∈Z<br />

Remark that<br />

⎡<br />

. ..<br />

⎢ . . . . . .<br />

⎢ ... 00000...<br />

⎢ ... 10000...<br />

ˆS = ⎢ ... 01000...<br />

⎢ ... 00100...<br />

⎢<br />

⎣ ... 00010...<br />

... . . . . . . ⎤<br />

⎥ ,<br />

⎥<br />

⎦<br />

. .<br />

x∈Z<br />

ˆS|x〉 = |x +1〉, ˆ S −1 |x〉 = |x − 1〉.<br />

⎡<br />

. ..<br />

⎢ . . . . . .<br />

⎢ ...01000...<br />

⎢ ...00100...<br />

Sˆ −1<br />

= ⎢ ...00010...<br />

⎢ ...00001...<br />

⎢<br />

⎣ ...00000...<br />

... . . . . . . ⎤<br />

⎥ .<br />

⎥<br />

⎦<br />

. .<br />

If ˆ P = |L〉〈L| and ˆ Q = |R〉〈R| are operations onto the two states of the co<strong>in</strong>,<br />

then one step of the quantum walk is given by the follow<strong>in</strong>g unitary operator:<br />

U =( ˆ S ⊗ ˆ Q + ˆ S −1 ⊗ ˆ P )(I ⊗ U).<br />

Not<strong>in</strong>g that ˆ PU = P and ˆ QU = Q, wesee<br />

⎡<br />

⎤<br />

. ..<br />

⎢ . . . . . . ⎥<br />

⎢ ... O P O O O ... ⎥<br />

⎢ ... Q O P O O ... ⎥<br />

U = ⎢ ... O Q O P O ... ⎥<br />

⎢ ... O O Q O P ... ⎥<br />

⎢<br />

⎣<br />

... O O O Q O ... ⎥<br />

⎦<br />

.<br />

... . . . . . ..<br />

with O =<br />

Then after n steps of the walk, the state is<br />

Ψn = U n Ψ0.<br />

� �<br />

00<br />

.<br />

00<br />

In the next subsection, we will obta<strong>in</strong> an explicit form for Ψn(x).<br />

Comb<strong>in</strong>atorial Approach<br />

This subsection gives a comb<strong>in</strong>atorial expression for the characteristic function<br />

of the quantum walk Xn. As a corollary, we obta<strong>in</strong> the m-th moment


Quantum Walks 319<br />

of Xn. As we expla<strong>in</strong>ed <strong>in</strong> the previous subsection, we need to know the follow<strong>in</strong>g<br />

comb<strong>in</strong>atorial expression to study P (Xn = x) forn + x =even.For<br />

fixed l and m with l + m = n and −l + m = x, weconsider<br />

Ξn(l, m) = �<br />

P l1 Q m1 P l2 Q m2 ···P ln Q mn<br />

lj,mj<br />

summed over all lj,mj ≥ 0 satisfy<strong>in</strong>g l1 + ···+ ln = l, m1 + ···+ mn = m<br />

and lj + mj = 1. We should note that<br />

Ψn(x) =Ξn(l, m)ϕ,<br />

s<strong>in</strong>ce Ψn(x) = T [Ψ L n (x),ΨR n (x)](∈ C2 ) is a two component vector of amplitudes<br />

of the particle be<strong>in</strong>g at site x at time n for <strong>in</strong>itial qubit state ϕ ∈ Φ<br />

and Ξn(l, m) is the sum of all possible paths <strong>in</strong> the trajectory consist<strong>in</strong>g of<br />

l steps left and m steps right with l =(n − x)/2 andm =(n + x)/2. For<br />

example, <strong>in</strong> the case of P (X4 = −2), we have the follow<strong>in</strong>g expression:<br />

Here we f<strong>in</strong>d a next nice relation:<br />

Ξ4(3, 1) = QP 3 + PQP 2 + P 2 QP + P 3 Q.<br />

P 2 = aP.<br />

By us<strong>in</strong>g this, the above example becomes<br />

Ξ4(3, 1) = a 2 QP + aP QP + aP QP + a 2 PQ.<br />

In general, we obta<strong>in</strong> the next table of products of the matrices P, Q, R and S:<br />

P Q R S<br />

P aP bR aR bP<br />

Q cS dQ cQ dS<br />

R cP dR cR dP<br />

S aS bQ aQ bS<br />

Table 1<br />

where PQ = bR, for example. S<strong>in</strong>ce P, Q, R and S form an orthonormal basis<br />

of the vector space of complex 2 × 2 matrices with respect to the trace <strong>in</strong>ner<br />

product 〈A|B〉 =tr(A ∗ B), Ξn(l, m) has the follow<strong>in</strong>g form:<br />

Ξn(l, m) =pn(l, m)P + qn(l, m)Q + rn(l, m)R + sn(l, m)S.<br />

Next problem is to obta<strong>in</strong> explicit forms of pn(l, m),qn(l, m),rn(l, m) and<br />

sn(l, m). In the above example of n = l + m =4case,wehave


320 N. Konno<br />

So, for example,<br />

Ξ4(4, 0) = a 3 P, Ξ4(3, 1) = 2abcP + a 2 bR + a 2 cS,<br />

Ξ4(2, 2) = bcdP + abcQ + b(ad + bc)R + c(ad + bc)S,<br />

Ξ4(1, 3) = 2bcdQ + bd 2 R + cd 2 S, Ξ4(2, 2) = d 3 Q.<br />

p4(3, 1) = 2abc, q4(3, 1) = 0, r4(3, 1) = a 2 b, s4(3, 1) = a 2 c.<br />

In a general case, the next key lemma is obta<strong>in</strong>ed.<br />

Lemma 1.1. We consider quantum walks <strong>in</strong> one dimension with abcd �= 0.<br />

Let<br />

� � � � � � � �<br />

ab<br />

00<br />

cd<br />

00<br />

P = , Q = , R = , S = .<br />

00<br />

cd<br />

00 ab<br />

and △ =detUwith U = P + Q. Suppose that l, m ≥ 0 with l + m = n, then<br />

we have<br />

(i) for l ∧ m(= m<strong>in</strong>{l, m}) ≥ 1,<br />

Ξn(l, m) =a l a m △ m<br />

l∧m<br />

(ii) for l(= n) ≥ 1,m=0,<br />

(iii) for l =0,m(= n) ≥ 1,<br />

�<br />

�<br />

−<br />

γ=1<br />

|b|2<br />

|a| 2<br />

�γ � �� �<br />

l − 1 m − 1<br />

γ − 1 γ − 1<br />

�<br />

l − γ m − γ 1 1<br />

× P + Q − R +<br />

aγ △aγ △b b S<br />

�<br />

,<br />

Ξn(l, 0) = a n−1 P,<br />

Ξn(0,m)=△ n−1 a n−1 Q.<br />

Proof. (a) pn(l, m) case:Firstweassumel ≥ 2andm ≥ 1. From Table 1, it<br />

is sufficient to consider only the follow<strong>in</strong>g case <strong>in</strong> order to compute pn(l, m):<br />

C(P, w) (2γ+1)<br />

n (l, m) =<br />

w1<br />

� �� �<br />

PP ···P<br />

w2<br />

w3<br />

w2γ<br />

w2γ+1<br />

� �� � � �� � � �� � � �� �<br />

QQ ···Q PP ···P ··· QQ ···Q PP ···P,<br />

where w = (w1,w2,...,w2γ+1) ∈ Z 2γ+1<br />

+ with w1,w2,...,w2γ+1 ≥ 1and<br />

γ ≥ 1. For example, PQP case is w1 = w2 = w3 =1andγ = 1. We should<br />

remark that l is the number of P ’s and m is the number of Q’s, so we have<br />

l = w1 + w3 + ···+ w2γ+1, m = w2 + w4 + ···+ w2γ.


Quantum Walks 321<br />

Moreover 2γ + 1 is the number of clusters of P ’s and Q’s. Next we consider<br />

the range of γ. The m<strong>in</strong>imum is γ = 1, that is, 3 clusters. This case is<br />

P ···PQ···QP ···P.<br />

The maximum is γ =(l − 1) ∧ m. Thiscaseis<br />

PQPQPQ···PQPQPP ···PP (l − 1 ≥ m),<br />

PQPQPQ···PQPQQ···QQP (l − 1 ≤ m),<br />

for example. Here we <strong>in</strong>troduce a set of sequences with 2γ + 1 components:<br />

for fixed γ ∈ [1, (l − 1) ∧ m]<br />

W (P, 2γ +1)={w =(w1,w2, ··· ,w2γ+1) ∈ Z 2γ+1 : w1 + w3 + ···+ w2γ+1 = l,<br />

w2 + w4 + ···+ w2γ = m, w1,w2,...,w2γ,w2γ+1 ≥ 1}.<br />

By us<strong>in</strong>g Table 1, we have<br />

C(P, w) (2γ+1)<br />

n (l, m) =a w1−1 Pd w2−1 Qa w3−1 P ···d w2γ −1 Qa w2γ+1−1 P<br />

= a l−(γ+1) d m−γ (PQ) γ P<br />

= a l−(γ+1) d m−γ b γ R γ P<br />

= a l−(γ+1) d m−γ b γ c γ−1 RP<br />

= a l−(γ+1) d m−γ b γ c γ P,<br />

where w ∈ W (P, 2γ +1).Forl ≥ 2,m≥ 1, that is, γ ≥ 1, we obta<strong>in</strong><br />

C(P, w) (2γ+1)<br />

n (l, m) =a l−(γ+1) b γ c γ d m−γ P.<br />

Note that the right-hand side of the above equation does not depend on<br />

w ∈ W (P, 2γ+1). So we write C(P ) (2γ+1)<br />

n (l, m) =C(P, w) (2γ+1)<br />

n (l, m)forw∈ W (P, 2γ+1). F<strong>in</strong>ally we compute the number of w =(w1,w2,...,w2γ,w2γ+1)<br />

satisfy<strong>in</strong>g w ∈ W (P, 2γ +1) by a standard comb<strong>in</strong>atorial argument as follows:<br />

� �� �<br />

l − 1 m − 1<br />

|W (P, 2γ +1)| =<br />

.<br />

γ γ − 1<br />

From the above observation, we obta<strong>in</strong><br />

pn(l, m)P =<br />

=<br />

=<br />

(l−1)∧m �<br />

γ=1<br />

(l−1)∧m �<br />

γ=1<br />

(l−1)∧m �<br />

γ=1<br />

�<br />

w∈W (P,2γ+1)<br />

C(P, w) (2γ+1)<br />

n (l, m)<br />

|W (P, 2γ +1)|C(P ) (2γ+1)<br />

n (l, m)<br />

� �� �<br />

l − 1 m − 1<br />

a<br />

γ γ − 1<br />

l−(γ+1) b γ c γ d m−γ P.


322 N. Konno<br />

So we conclude that<br />

pn(l, m) =<br />

(l−1)∧m �<br />

γ=1<br />

When l ≥ 1andm = 0, it is easy to see that<br />

� �� �<br />

l − 1 m − 1<br />

a<br />

γ γ − 1<br />

l−(γ+1) b γ c γ d m−γ .<br />

pn(l, 0)P = P l = a l−1 P.<br />

Furthermore, when l =1,m≥ 1andl =0,m≥ 0, it is clear that pn(l, m)=0.<br />

(b) qn(l, m) case : We beg<strong>in</strong> with l ≥ 1andm ≥ 2. In this case, it is enough<br />

to study only the follow<strong>in</strong>g case:<br />

w1<br />

w2<br />

w3<br />

w2γ<br />

w2γ+1<br />

� �� � � �� � � �� � � �� � � �� �<br />

QQ ···Q PP ···P QQ ···Q ··· PP ···P QQ ···Q,<br />

where w = (w1,w2,...,w2γ+1) ∈ Z 2γ+1<br />

+ with w1,w2,...,w2γ+1 ≥ 1and<br />

γ ≥ 1. Note that<br />

l = w2 + w4 + ···+ w2γ, m = w1 + w3 + ···+ w2γ+1.<br />

Then 2γ + 1 is the number of clusters of P ’s and Q’s. As <strong>in</strong> the part (a), we<br />

obta<strong>in</strong><br />

qn(l, m) =<br />

l∧(m−1) �<br />

γ=1<br />

� l − 1<br />

γ − 1<br />

When l =0andm ≥ 1, this case is<br />

�� �<br />

m − 1<br />

a<br />

γ<br />

l−γ b γ c γ d m−(γ+1) .<br />

qn(0,m)Q = Q m = d m−1 Q.<br />

Furthermore, when m =1,l≥ 1andm =0,l≥ 0, we see that qn(l, m) =0.<br />

(c) rn(l, m) case : Suppose l ≥ 1andm ≥ 1. In this case, it is sufficient to<br />

consider only the follow<strong>in</strong>g case:<br />

w1<br />

� �� �<br />

PP ···P<br />

w2<br />

w3<br />

w2γ<br />

� �� � � �� � � �� �<br />

QQ ···Q PP ···P ··· QQ ···Q,<br />

where w =(w1,w2,...,w2γ+1) ∈ Z 2γ<br />

+ with w1,w2,...,w2γ ≥ 1andγ ≥ 1.<br />

For example, PQPQ case is w1 = w2 = w3 = w4 =1andγ = 2. It should<br />

be noted that l is the number of P ’s and m is the number of Q’s, so we have<br />

l = w1 + w3 + ···+ w2γ−1, m = w2 + w4 + ···+ w2γ.<br />

Then 2γ is the number of clusters of P ’s and Q’s. The range of γ is determ<strong>in</strong>ed<br />

by the follow<strong>in</strong>g observation. The m<strong>in</strong>imum is γ = 1, that is, 2 clusters. This<br />

case is


Quantum Walks 323<br />

P ···PQ···Q.<br />

The maximum is γ = l ∧ m. Thiscaseis<br />

PQPQPQ···PQPP ···PPQ (l ≥ m),<br />

PQPQPQ···PQPQQ···QQ (m ≥ l),<br />

for examples. As <strong>in</strong> the part (a), we get<br />

l∧m �<br />

� �� �<br />

l − 1 m − 1<br />

rn(l, m) =<br />

a<br />

γ − 1 γ − 1<br />

l−γ b γ c γ−1 d m−γ .<br />

γ=1<br />

If l ∧ m =0,thenrn(l, m) =0.<br />

(d) sn(l, m) case : To beg<strong>in</strong>, we assume l ≥ 1andm ≥ 1. To compute sn(l, m),<br />

it is enough to consider only the follow<strong>in</strong>g case:<br />

w1<br />

w2<br />

w3<br />

w2γ<br />

� �� � � �� � � �� � � �� �<br />

QQ ···Q PP ···P QQ ···Q ··· PP ···P,<br />

where w =(w1,w2,...,w2γ+1) ∈ Z 2γ<br />

+ with w1,w2,...,w2γ ≥ 1andγ ≥ 1.<br />

Then<br />

l = w2 + w4 + ···+ w2γ, m = w1 + w3 + ···+ w2γ−1.<br />

As <strong>in</strong> the case of part (c), we obta<strong>in</strong><br />

l∧m �<br />

� �� �<br />

l − 1 m − 1<br />

sn(l, m) =<br />

a<br />

γ − 1 γ − 1<br />

l−γ b γ−1 c γ d m−γ .<br />

γ=1<br />

When l ∧ m =0,thiscaseissn(l, m) =0.<br />

We assume abcd �= 0.Whenl ∧ m ≥ 1, comb<strong>in</strong><strong>in</strong>g the above results (a) -<br />

(d) gives<br />

�<br />

Ξn(l, m) =a l d m<br />

l∧m<br />

γ=1<br />

� �γ � �� �<br />

bc l − 1 m − 1<br />

ad γ − 1 γ − 1<br />

�<br />

l − γ m − γ 1 1<br />

× P + Q + R +<br />

aγ dγ c b S<br />

�<br />

.<br />

From c = −△b, d = △a, the proof of Lemma 1.1 (i) is complete. Furthermore,<br />

parts (ii) and (iii) are easily shown, so we will omit the proofs of them.<br />

Then the distribution of Xn can be derived from Lemma 1.1 by direct<br />

computation.


324 N. Konno<br />

Lemma 1.2. For k =1, 2,...,[n/2], we have<br />

P (Xn = n − 2k)<br />

= |a| 2(n−1)<br />

k� k�<br />

�<br />

− |b|2<br />

|a| 2<br />

�γ+δ� �� �� �� �<br />

k − 1 k − 1 n − k − 1 n − k − 1<br />

γ − 1 δ − 1 γ − 1 δ − 1<br />

γ=1δ=1<br />

� �<br />

1<br />

×<br />

γδ<br />

�<br />

γ=1 δ=1<br />

{k 2 |a| 2 +(n − k) 2 |b| 2 − (γ + δ)(n − k)}|α| 2<br />

+{k 2 |b| 2 +(n − k) 2 |a| 2 − (γ + δ)k}|β| 2<br />

+ 1<br />

|b| 2<br />

�<br />

{(n − k)γ − kδ + n(2k − n)|b| 2 }aαbβ<br />

+{−kγ +(n− k)δ + n(2k − n)|b| 2 �<br />

}aαbβ + γδ<br />

�<br />

,<br />

P (Xn = −(n − 2k))<br />

= |a| 2(n−1)<br />

k� k�<br />

�<br />

− |b|2<br />

|a| 2<br />

�γ+δ � �� �� �� �<br />

k − 1 k − 1 n − k − 1 n − k − 1<br />

γ − 1 δ − 1 γ − 1 δ − 1<br />

� �<br />

1<br />

×<br />

γδ<br />

�<br />

{k 2 |b| 2 +(n − k) 2 |a| 2 − (γ + δ)k}|α| 2<br />

+{k 2 |a| 2 +(n − k) 2 |b| 2 − (γ + δ)(n − k)}|β| 2<br />

+ 1<br />

|b| 2<br />

�<br />

{kγ − (n − k)δ − n(2k − n)|b| 2 }aαbβ<br />

+{−(n − k)γ + kδ − n(2k − n)|b| 2 �<br />

}aαbβ + γδ<br />

�<br />

,<br />

P (Xn = n) =|a| 2(n−1) {|b| 2 |α| 2 + |a| 2 |β| 2 − (aαbβ + aαbβ)},<br />

P (Xn = −n) =|a| 2(n−1) {|a| 2 |α| 2 + |b| 2 |β| 2 +(aαbβ + aαbβ)},<br />

where [x] is the <strong>in</strong>teger part of x.<br />

By us<strong>in</strong>g Lemma 1.2, we obta<strong>in</strong> a comb<strong>in</strong>atorial expression for the characteristic<br />

function of Xn as follows. This result will be used <strong>in</strong> order to obta<strong>in</strong><br />

a limit theorem of Xn.


Quantum Walks 325<br />

Theorem 1.3. (i) When abcd �= 0, we have E(eiξXn )=<br />

|a| 2(n−1)<br />

�� cos(nξ) − �� |a| 2 −|b| 2�� |α| 2 −|β| 2� +2(aαbβ + aαbβ) � �<br />

i s<strong>in</strong>(nξ)<br />

+<br />

[ n−1<br />

2 ]<br />

�<br />

k=1<br />

k�<br />

k�<br />

γ=1 δ=1<br />

×<br />

�<br />

n<br />

+I<br />

2 −<br />

�<br />

− |b|2<br />

|a| 2<br />

�γ+δ � �� �� �� �<br />

k − 1 k − 1 n − k − 1 n − k − 1<br />

γ − 1 δ − 1 γ − 1 δ − 1<br />

���<br />

(n − k) 2 + k 2 − n(γ + δ)+ 2γδ<br />

|b| 2<br />

�<br />

cos((n − 2k)ξ)<br />

�<br />

+(n − 2k) −{n(|a| 2 −|b| 2 )+γ + δ}(|α| 2 −|β| 2 )<br />

� �<br />

�<br />

�<br />

γ + δ<br />

+ − 2n (aαbβ + aαbβ) i s<strong>in</strong>((n − 2k)ξ)<br />

|b| 2<br />

� 1<br />

γδ<br />

�<br />

n<br />

� �<br />

, 0<br />

2<br />

×<br />

n<br />

2�<br />

n<br />

2�<br />

γ=1 δ=1<br />

�<br />

− |b|2<br />

|a| 2<br />

�γ+δ � n<br />

2<br />

where I(x, y) =1(resp. =0)if x = y (resp. x �= y).<br />

(ii) When b =0, that is,<br />

�<br />

iθ e 0<br />

U =<br />

0 △e−iθ �<br />

,<br />

�2� n �2 − 1 2 − 1<br />

γ − 1 δ − 1<br />

� ��<br />

1<br />

× n<br />

4γδ<br />

2 − 2n(γ + δ)+ 4γδ<br />

|b| 2<br />

�� ,<br />

where θ ∈ R and △ =detU ∈ C with |△| =1, then we have<br />

(iii) When a =0, that is,<br />

E(e iξXn )=cos(nξ)+i(|β| 2 −|α| 2 )s<strong>in</strong>(nξ).<br />

�<br />

U =<br />

0 e iθ<br />

−△e −iθ 0<br />

where θ ∈ R and △ =detU∈ C with |△| =1, then we have<br />

E(e iξXn �<br />

2 2 cos ξ + i(|α| −|β| )s<strong>in</strong>ξ if n is odd,<br />

)=<br />

1 if n is even.<br />

From this theorem, we have the m-th moment of Xn <strong>in</strong> the standard fashion.<br />

The follow<strong>in</strong>g result can be used <strong>in</strong> order to study symmetry of distribution<br />

of Xn.<br />

�<br />

,


326 N. Konno<br />

Corollary 1.4. (i) We assume that abcd �= 0.Whenm is odd, we have<br />

E((Xn) m )=|a| 2(n−1)<br />

+<br />

× (n − 2k)m+1<br />

[ n−1<br />

2 ]<br />

�<br />

k=1<br />

k�<br />

� �<br />

−n m �� |a| 2 −|b| 2�� |α| 2 −|β| 2� +2(aαbβ + aαbβ) ��<br />

k�<br />

γ=1 δ=1<br />

γδ<br />

�<br />

− |b|2<br />

|a| 2<br />

� �<br />

γ+δ<br />

k − 1<br />

γ − 1<br />

��<br />

k − 1<br />

δ − 1<br />

��<br />

n − k − 1<br />

γ − 1<br />

�<br />

−{n(|a| 2 −|b| 2 )+γ + δ}(|α| 2 −|β| 2 )<br />

We assume that abcd �= 0.Whenm is even, we have<br />

E((Xn) m )=|a| 2(n−1)<br />

+<br />

× (n − 2k)m<br />

[ n−1<br />

2 ]<br />

�<br />

k=1<br />

�<br />

k�<br />

n m<br />

k�<br />

γ=1 δ=1<br />

γδ<br />

(ii) When b =0, that is,<br />

�<br />

− |b|2<br />

|a| 2<br />

� �<br />

γ+δ<br />

k − 1<br />

γ − 1<br />

��<br />

� �<br />

�<br />

γ + δ<br />

+ − 2n (aαbβ + aαbβ)<br />

|b| 2 �<br />

.<br />

��<br />

k − 1<br />

δ − 1<br />

��<br />

�<br />

(n − k) 2 + k 2 − n(γ + δ)+ 2γδ<br />

|b| 2<br />

U =<br />

�<br />

iθ e 0<br />

0 △e−iθ �<br />

,<br />

n − k − 1<br />

γ − 1<br />

�� .<br />

where θ ∈ [0, 2π) and △ =detU∈C with |△| =1, then we have<br />

E((Xn) m �<br />

m 2 2 n (|β| −|α| ) if m is odd,<br />

)=<br />

nm if m is even.<br />

(iii) When a =0, that is,<br />

�<br />

U =<br />

0 e iθ<br />

−△e −iθ 0<br />

where θ ∈ [0, 2π) and △ =detU∈C with |△| =1, then we have<br />

E((Xn) m ⎧<br />

⎨<br />

)=<br />

⎩<br />

�<br />

,<br />

|α| 2 −|β| 2 if n and m are odd,<br />

1 if n is odd and m is even,<br />

0 if n is even.<br />

��<br />

�<br />

n − k − 1<br />

δ − 1<br />

�<br />

n − k − 1<br />

δ − 1


Quantum Walks 327<br />

For any case, when m is even, E((Xn) m ) is <strong>in</strong>dependent of <strong>in</strong>itial qubit<br />

state ϕ. Therefore a parity law of the m-th moment can be derived from the<br />

above theorem.<br />

Symmetry of Distribution<br />

In this subsection, the follow<strong>in</strong>g necessary and sufficient condition for symmetry<br />

of distribution of Xn is given.<br />

Theorem 1.5. We assume abcd �= 0. Then we have<br />

where<br />

Φs = Φ0 = Φ⊥,<br />

Φs = {ϕ ∈ Φ : P (Xn = x) =P (Xn = −x) for any n ∈ Z+ and x ∈ Z},<br />

Φ0 = {ϕ ∈ Φ : E(Xn) =0 for any n ∈ Z+} ,<br />

Φ⊥ = � ϕ = T [α, β] ∈ Φ : |α| = |β|, aαbβ + aαbβ =0 � .<br />

This result is a generalization of Konno, Namiki, and Soshi (2004) for the<br />

Hadamard walk.<br />

Proof. (i) Φs ⊂ Φ0. This is obvious by the def<strong>in</strong>itions of Φs and Φ0.<br />

(ii) Φ0 ⊂ Φ⊥. By Corollary 1.4 (i) (m = 1 case), we see that<br />

E(X1) =E(X2) =0<br />

if and only if<br />

� |a| 2 −|b| 2 �� |α| 2 −|β| 2� +2(aαbβ + aαbβ) =0. (1.1.1)<br />

Then (1.1.1) implies that for n ≥ 3, Corollary 1.4 (i) (m =1case)canbe<br />

rewritten as<br />

E(Xn) =−(|α| 2 −|β| 2 ) |a|2(n−1)<br />

2|b| 2<br />

[ n−1<br />

2 ]<br />

�<br />

k� k�<br />

�<br />

×<br />

−<br />

k=1 γ=1 δ=1<br />

|b|2<br />

|a| 2<br />

�γ+δ � �� �� �<br />

k − 1 k − 1 n − k − 1<br />

γ − 1 δ − 1 γ − 1<br />

� � 2 n − k − 1 (n − 2k) (γ + δ)<br />

×<br />

.<br />

δ − 1 γδ


328 N. Konno<br />

Therefore E(Xn) =0(n ≥ 3) gives |α| = |β|. Comb<strong>in</strong><strong>in</strong>g |α| = |β| with<br />

(1.1.1), we have the desired result.<br />

(iii) Φ⊥ ⊂ Φs. We assume that<br />

|α| = |β|, aαbβ + aαbβ =0. (1.1.2)<br />

By us<strong>in</strong>g Lemma 1.2 and (1.1.2), we see that for k =1, 2,...,[n/2],<br />

P (Xn = n − 2k) =P (Xn = −(n − 2k))<br />

= |a| 2(n−1)<br />

k� k�<br />

�<br />

−<br />

γ=1 δ=1<br />

|b|2<br />

|a| 2<br />

�γ+δ � �� �� �� �<br />

k − 1 k − 1 n − k − 1 n − k − 1<br />

γ − 1 δ − 1 γ − 1 δ − 1<br />

�<br />

�<br />

×<br />

,<br />

and<br />

|α| 2<br />

γδ {(n − k)2 + k 2 − n(γ + δ)} + 1<br />

|b| 2<br />

P (Xn = n) =P (Xn = −n) =|a| 2(n−1) |α| 2 .<br />

So the desired conclusion is obta<strong>in</strong>ed.<br />

Weak Limit Theorem<br />

In the subsection, we give a new type of limit theorems for Xn with abcd �= 0<br />

as follows.<br />

Theorem 1.6. If n →∞,then<br />

where Z has a density<br />

f(x) =f(x; ϕ = T [α, β])<br />

�<br />

1 −|a| 2<br />

=<br />

π(1 − x2 ) � |a| 2 − x2 Xn<br />

n<br />

⇒ Z,<br />

� �<br />

1 − |α| 2 −|β| 2 +<br />

for x ∈ (−|a|, |a|) and f(x) =0for |x| ≥|a| with<br />

�<br />

E(Z) =− |α| 2 −|β| 2 +<br />

E(Z 2 )=1− � 1 −|a| 2 ,<br />

aαbβ + aαbβ<br />

|a| 2<br />

� �<br />

x ,<br />

aαbβ + aαbβ<br />

|a| 2<br />

�<br />

× (1 − � 1 −|a| 2 ),<br />

and Yn ⇒ Y means that Yn converges weakly to a limit Y .


Quantum Walks 329<br />

Proof. We beg<strong>in</strong> with <strong>in</strong>troduc<strong>in</strong>g the Jacobi polynomial P ν,µ<br />

n (x), where<br />

P ν,µ<br />

n (x) is orthogonal on [−1, 1] with respect to (1 − x)ν (1 + x) µ with<br />

ν, µ > −1. Then the follow<strong>in</strong>g relation holds:<br />

P ν,µ<br />

n (x) =<br />

Γ (n + ν +1)<br />

Γ (n +1)Γ (ν +1) 2F1(−n, n + ν + µ +1;ν +1;(1− x)/2),<br />

(1.1.3)<br />

where 2F1(a, b; c; z) is the hypergeometric series and Γ (z) is the gamma function.<br />

In general, as for orthogonal polynomials, see Andrews, Askey, and Roy<br />

(1999). Then we see that<br />

k�<br />

γ=1<br />

�<br />

− |b|2<br />

|a| 2<br />

�γ−1 1<br />

γ<br />

� k − 1<br />

γ − 1<br />

�� �<br />

n − k − 1<br />

γ − 1<br />

= 2F1(−(k − 1), −{(n − k) − 1};2;−|b| 2 /|a| 2 )<br />

= |a| −2(k−1) 2F1(−(k − 1),n− k +1;2;1−|a| 2 )<br />

= 1<br />

(2|a| 2 − 1).<br />

k |a|−2(k−1) P 1,n−2k<br />

k−1<br />

The first equality is given by the def<strong>in</strong>ition of the hypergeometric series. The<br />

second equality comes from the follow<strong>in</strong>g relation:<br />

2F1(a, b; c; z) =(1− z) −a 2F1(a, c − b; c; z/(z − 1)).<br />

The last equality follows from (1.1.3). In a similar way, we have<br />

k�<br />

γ=1<br />

�<br />

− |b|2<br />

|a| 2<br />

�γ−1 � �� �<br />

k − 1 n − k − 1<br />

γ − 1 γ − 1<br />

= |a| −2(k−1) P 0,n−2k<br />

k−1 (2|a| 2 − 1).<br />

By us<strong>in</strong>g the above relations and Theorem 1.3, we obta<strong>in</strong> the next asymptotics<br />

of characteristic function E(e iξXn/n ):<br />

Lemma 1.7. if n →∞with k/n = x ∈ (−(1 −|a|)/2, (1 + |a|)/2), then<br />

E(e iξXn/n ) ∼<br />

�� 2 2x − 2x +1<br />

×<br />

x2 [(n−1)/2] �<br />

k=1<br />

(P 1,n−2k<br />

k−1<br />

|a| 2n−4k−2 |b| 4<br />

) 2 − 2<br />

x<br />

� ��<br />

1 − 2x<br />

+<br />

−<br />

x<br />

1<br />

�<br />

−2 |α| 2 −|β| 2 aαbβ + aαbβ<br />

−<br />

|b| 2<br />

�<br />

1,n−2k<br />

Pk−1 P 0,n−2k<br />

k−1<br />

+ 2<br />

|b|<br />

0,n−2k<br />

(P 2 k−1<br />

) 2<br />

�<br />

× cos((1 − 2x)ξ)<br />

x {(|a|2 −|b| 2 )(|α| 2 −|β| 2 )+2(aαbβ + aαbβ)}(P 1,n−2k<br />

k−1<br />

P 0,n−2k<br />

k−1 P 1,n−2k<br />

k−1<br />

�<br />

i s<strong>in</strong>((1 − 2x)ξ)<br />

) 2<br />

�<br />

,


330 N. Konno<br />

where f(n) ∼ g(n) means f(n)/g(n) → 1(n → ∞), and P i,n−2k<br />

k−1<br />

P i,n−2k<br />

k−1 (2|a| 2 − 1) (i =0, 1).<br />

Next we use an asymptotic result on the Jacobi polynomial P α+an,β+bn<br />

n (x)<br />

derived by Chen and Ismail (1991). By us<strong>in</strong>g (2.16) <strong>in</strong> their paper with α → 0<br />

or 1,a → 0,β = b → (1 − 2x)/x, x → 2|a| 2 − 1and△→4(1 −|a| 2 )(4x2 −<br />

4x +1−|a| 2 )/x2 , we have the follow<strong>in</strong>g lemma. It should be noted that<br />

there are some m<strong>in</strong>or errors <strong>in</strong> (2.16) <strong>in</strong> that paper, for example, � (−△) →<br />

� −1<br />

(−△) .<br />

Lemma 1.8. If n →∞with k/n = x ∈ (−(1 −|a|)/2, (1 + |a|)/2), then<br />

P 0,n−2k<br />

k−1 ∼ 2|a|2k−n<br />

� πn √ −Λ cos(An + B),<br />

P 1,n−2k<br />

k−1 ∼ 2|a|2k−n<br />

�<br />

x<br />

� √<br />

πn −Λ (1 − x)(1 −|a| 2 cos(An + B + θ),<br />

)<br />

where Λ = (1 −|a| 2 )(4x 2 − 4x +1−|a| 2 ), A and B are some constants<br />

(which are <strong>in</strong>dependent of n), and θ ∈ [0,π/2] is determ<strong>in</strong>ed by cos θ =<br />

� (1 −|a| 2 )/4x(1 − x).<br />

From the Riemann-Lebesgue lemma (see Durrett (2004), for example) and<br />

Lemmas 1.7 and 1.8, we see that, if n →∞,then<br />

E(e iξXn/n ) →<br />

1 −|a| 2 � 1<br />

2<br />

1<br />

dx<br />

π 1−|a| x(1 − x) 2<br />

� (|a| 2 − 1)(4x2 − 4x +1−|a| 2 )<br />

�<br />

�<br />

× cos((1 − 2x)ξ) − (1 − 2x) |α| 2 −|β| 2 aαbβ + aαbβ<br />

+<br />

|a| 2<br />

�<br />

�<br />

i s<strong>in</strong>((1 − 2x)ξ)<br />

� �<br />

1 −|a| 2 |a|<br />

1<br />

=<br />

dx<br />

π −|a| (1 − x2 ) � |a| 2 − x2 �<br />

�<br />

× cos(xξ) − x |α| 2 −|β| 2 aαbβ + aαbβ<br />

+<br />

|a| 2<br />

� �<br />

i s<strong>in</strong>(xξ)<br />

� �<br />

|a|<br />

1 −|a| 2<br />

=<br />

−|a| π(1 − x2 ) � |a| 2 − x2 � �<br />

1 − |α| 2 −|β| 2 aαbβ + aαbβ<br />

+<br />

|a| 2<br />

� �<br />

x e iξx dx<br />

= φ(ξ).<br />

Then φ(ξ) is cont<strong>in</strong>uous at ξ = 0, so the cont<strong>in</strong>uity theorem implies that<br />

Xn/n converges weakly to a limit Z with characteristic function φ. Moreover<br />

the density function of Z is given by<br />

=


Quantum Walks 331<br />

f(x; T �<br />

1 −|a| 2<br />

[α, β]) =<br />

π(1 − x2 ) � |a| 2 − x2 � �<br />

1 − |α| 2 −|β| 2 aαbβ + aαbβ<br />

+<br />

|a| 2<br />

� �<br />

x ,<br />

for x ∈ (−|a|, |a|). So the proof of Theorem 1.6 is complete.<br />

It can be confirmed that f(x; T [α, β]) satisfies the property of a density<br />

function as follows: It is easily checked that f(x; T [α, β]) ≥ 0, s<strong>in</strong>ce<br />

�<br />

1 ≥± |α| 2 −|β| 2 aαbβ + aαbβ<br />

+<br />

|a| 2<br />

�<br />

|a|. (1.1.4)<br />

On the other hand,<br />

� |a|<br />

f(x;<br />

−|a|<br />

T � �<br />

1 −|a| 2 1<br />

[α, β]) dx =<br />

π 0<br />

�<br />

1 −|a| 2<br />

= Γ (1/2)<br />

π<br />

2 2F1(1/2, 1; 1; |a| 2 )<br />

=1.<br />

t −1/2 (1 − t) −1/2 (1 −|a| 2 t) −1 dt<br />

The last equality comes from Γ (1/2) = √ π and 2F1(1/2, 1; 1; |a| 2 ) =<br />

1/ � 1 −|a| 2 .<br />

Moreover, us<strong>in</strong>g (1.1.4), we see that for any m ≥ 1,<br />

|E(Z m )|≤2|a| m .<br />

In the symmetric case of the Hadamard walk, we have the follow<strong>in</strong>g result.<br />

Proposition 1.9. For n ≥ 1,<br />

E(Z 2n )=1− 1 √ 2<br />

Proof. We beg<strong>in</strong> with<br />

where<br />

E(Z 2n )=<br />

In =<br />

� 1/ √ 2<br />

� 1<br />

−1/ √ 2<br />

It should be noted that<br />

0<br />

I2n+2 =<br />

n−1 �<br />

k=0<br />

1<br />

2 3k<br />

x 2n<br />

π(1 − x 2 ) √ 1 − 2x<br />

yn (2 − y2 ) � dy =<br />

1 − y2 � �<br />

2k<br />

, E(Z<br />

k<br />

2n−1 )=0.<br />

π<br />

2 dx = 2(3−2n)/2<br />

� π/2<br />

0<br />

s<strong>in</strong> n θ<br />

1+cos 2 θ dθ.<br />

� π/2<br />

{2 − (1 + cos<br />

0<br />

2 θ)} s<strong>in</strong> 2n θ<br />

1+cos2θ =2I2n −<br />

(2n − 1)(2n − 3) ···1<br />

2n(2n − 2) ···2<br />

dθ<br />

× π<br />

2 .<br />

× I2n,


332 N. Konno<br />

Let Jn = I2n/2 n .Thenweseethatforn ≥ 0<br />

where<br />

So we have<br />

Therefore for n ≥ 1,<br />

Jn+1 − Jn = − π<br />

2 3n+2<br />

� 2n<br />

n<br />

�<br />

,<br />

J0 = π<br />

2 √ 2 , J1 = 2 − √ 2<br />

4 √ 2 π.<br />

I2n =2 n Jn =2 n<br />

So the proof is complete.<br />

Fourier Analysis<br />

�<br />

1<br />

2 √ 2 −<br />

n−1 �<br />

k=0<br />

E(Z 2n )=1− 1<br />

n−1 �<br />

√<br />

2 k=0<br />

1<br />

2 3k+2<br />

1<br />

2 3k<br />

� �<br />

2k<br />

k<br />

�<br />

π.<br />

� �<br />

2k<br />

.<br />

k<br />

The two-component wave function of the quantum walk was given by<br />

� �<br />

L Ψn (x)<br />

Ψn(x) = .<br />

Ψ R n (x)<br />

Follow<strong>in</strong>g the def<strong>in</strong>ition of the model, we have<br />

�<br />

L aΨn (x +1)+bΨ<br />

Ψn+1(x) =<br />

R �<br />

n (x +1)<br />

, (1.1.5)<br />

(x − 1)<br />

where U =<br />

� �<br />

ab<br />

∈ U(2). If we set<br />

cd<br />

and assume the relations<br />

cΨ L n (x − 1) + dΨ R n<br />

ˆΨn(k) =<br />

Ψ j n (x) =<br />

� π<br />

−π<br />

�<br />

ˆΨ L<br />

n (k)<br />

ˆΨ R �<br />

,<br />

n (k)<br />

dk<br />

2π eikx ˆ Ψ j n (k),<br />

ˆΨ j n(k) = �<br />

e −ikx Ψ j n(x),<br />

x∈Z


Quantum Walks 333<br />

for j = L, R, then (1.1.5) is rewritten <strong>in</strong> the wave-number space (k-space) of<br />

the Fourier transformation, k ∈ [−π, π) as<br />

where U(k) isgivenby<br />

ˆΨn+1(k) =U(k) ˆ Ψn(k), n =0, 1, 2, ··· ,<br />

U(k) =<br />

�<br />

ik e 0<br />

0 e−ik �<br />

U.<br />

The state at time step n is then obta<strong>in</strong>ed from the <strong>in</strong>itial state ˆ Ψ0(k) =<br />

� α<br />

β<br />

�<br />

,α,β ∈ C, |α| 2 + |β| 2 =1, by<br />

ˆΨn(k) =U(k) n ˆ Ψ0(k), (1.1.6)<br />

whose Fourier transformation gives the wave function at time step n as<br />

Ψn(x) =<br />

=<br />

� π<br />

−π<br />

� π<br />

−π<br />

dk<br />

2π eikx ˆ Ψn(k)<br />

dk<br />

2π eikx U(k) n ˆ Ψ0(k).<br />

The probability distribution function <strong>in</strong> the real space at time step n is<br />

given by<br />

Pn(x) =||Ψn(x)|| 2<br />

=<br />

� π<br />

−π<br />

dk ′<br />

2π<br />

� π<br />

−π<br />

dk<br />

2π ei(k−k′ )x � Ψˆ ∗<br />

0 (k ′ )U ∗ (k ′ ) n��<br />

U(k) n �<br />

Ψ0(k) ˆ .<br />

(1.1.7)<br />

Let Xn denote the position of the one-dimensional quantum walk at time<br />

step n =0, 1, 2, ··· and consider a function f of x ∈ Z. The expectation of<br />

f(Xn) is def<strong>in</strong>ed by<br />

E(f(Xn)) = �<br />

f(x)Pn(x)<br />

x∈Z<br />

= �<br />

x∈Z<br />

� π<br />

dk<br />

f(x)<br />

−π<br />

′<br />

2π e−ik′ x<br />

Ψˆ ∗<br />

n (k ′ � π<br />

dk<br />

)<br />

−π 2π eikxΨn(k). ˆ<br />

If f(x) =xr ,r =0, 1, 2, ···, it is written as<br />

E(X r n )=�<br />

� π<br />

dk<br />

−π<br />

′<br />

2π e−ik′ x<br />

Ψˆ ∗<br />

n (k ′ � π ��<br />

dk<br />

)<br />

−i<br />

−π 2π<br />

d<br />

�r e<br />

dk<br />

ikx<br />

�<br />

ˆΨn(k).<br />

x∈Z


334 N. Konno<br />

We note that ˆ Ψn(k) should be a periodic function of k ∈ [−π, π), and then<br />

by partial <strong>in</strong>tegrations, we will have<br />

� π ��<br />

dk<br />

−i<br />

2π<br />

d<br />

�r e<br />

dk<br />

ikx<br />

� � π<br />

ˆΨn(k)<br />

dk<br />

=<br />

2π eikx<br />

�<br />

i d<br />

�r ˆΨn(k),<br />

dk<br />

−π<br />

and thus<br />

E(X r � π<br />

dk<br />

n)=<br />

−π 2π ˆ Ψ ∗ �<br />

n(k) i d<br />

�r ˆΨn(k), (1.1.8)<br />

dk<br />

wherewehaveusedthesummationformula �<br />

x∈Z eikx =2πδ(k). Then, if<br />

f(x) is analytic around x = 0, that is, if it has a converg<strong>in</strong>g Taylor expansion<br />

<strong>in</strong> the form f(x) = �∞ j=0 ajxj , we will have the formula<br />

E(f(Xn)) =<br />

� π<br />

−π<br />

−π<br />

dk<br />

2π ˆ Ψ ∗ n(k)f<br />

Method of Grimmett, Janson, and Scudo (2004)<br />

�<br />

i d<br />

�<br />

ˆΨn(k). (1.1.9)<br />

dk<br />

The unitary matrix U(k) has two eigenvalues λ1(k) andλ2(k) and has correspond<strong>in</strong>g<br />

eigenvectors v1(k) andv2(k) that def<strong>in</strong>e an orthonormal basis for<br />

H = L2 (K) ⊗ Hc, whereK =[−π, π) andHc = C2 .PutD = id/dk. Then<br />

D r 2�<br />

Ψn(k) = (n)rλj(k)<br />

j=1<br />

n−r (Dλj(k)) r 〈vj(k),Ψ0(k)〉vj(k)+O(n r−1 ),<br />

(1.1.10)<br />

where (n)r = n(n − 1) ···(n − r + 1). On the other hand, (1.1.8) becomes<br />

E(X r n )=<br />

� π<br />

−π<br />

Comb<strong>in</strong><strong>in</strong>g (1.1.10) with (1.1.11) yields<br />

ˆΨ ∗ n (k)DrΨn(k) ˆ dk<br />

. (1.1.11)<br />

2π<br />

E(X r n )=<br />

� π 2�<br />

(n)rλj(k)<br />

−π j=1<br />

n−r (Dλj(k)) r 〈vj(k),Ψ0(k)〉〈Ψ0(k),vj(k)〉 dk<br />

2π<br />

=<br />

� π<br />

−π j=1<br />

2�<br />

(n)r<br />

+O(n −1 )<br />

� �r Dλj(k)<br />

2 dk<br />

|〈vj(k),Ψ0(k)〉|<br />

λj(k)<br />

2π + O(n−1 ). (1.1.12)<br />

Therefore let Ω = K ×{1, 2} and µ denote the probability measure on Ω<br />

given by dk/2π ×|〈vj(k),Ψ0(k)〉| 2 on K ×{j}. Put<br />

hj(k) = Dλj(k)<br />

λj(k) .


Quantum Walks 335<br />

Def<strong>in</strong>e h : Ω → R by h(k, j) =hj(k). By (1.1.12),<br />

�<br />

h r dµ.<br />

lim<br />

n→∞ E((Xn/n) r )=<br />

Therefore Grimmett, Janson, and Scudo (2004) obta<strong>in</strong>ed<br />

Theorem 1.10. If n →∞,then<br />

Xn<br />

n<br />

Ω<br />

⇒ Y = h(Z),<br />

where Z is a random variable of Ω with distribution µ.<br />

Here we consider the Hadamard walk case. In this case,<br />

U(k) = 1 �<br />

ik e e<br />

√<br />

2<br />

ik<br />

e−ik −e−ik �<br />

.<br />

The eigenvalues are<br />

λ1(k) = 1 √ 2 ( √ I + i s<strong>in</strong> k), λ2(k) = 1<br />

√ 2 (− √ I + i s<strong>in</strong> k), (1.1.13)<br />

where I =1+cos2k.Then cos k<br />

cos k<br />

h1(k) =− √ , h2(k) = √ ,<br />

I I<br />

and<br />

�√I +cosk<br />

v1(k) =<br />

2 √ �<br />

e<br />

I<br />

ik �<br />

�√I √<br />

− cos k<br />

, v2(k) =<br />

I − cos k<br />

2 √ �<br />

e<br />

I<br />

ik<br />

− √ �<br />

,<br />

I − cos k<br />

with 〈vj(k),vj(k)〉 =1(j =1, 2). Let pj(k) =|〈vj(k),Ψ0(k)〉| 2 for j =1, 2.<br />

Note that p1(k)+p2(k) =1(k∈ [−π, π)). From Theorem 1.10, we have<br />

�<br />

P (Y ≤ y) =<br />

p1(k) dk<br />

2π +<br />

�<br />

p2(k) dk<br />

2π .<br />

{k∈[−π,π):h1(k)≤y}<br />

{k∈[−π,π):h2(k)≤y}<br />

Let k = k(y) ∈ [0,π) denote the unique solution of h1(k) =− cos k/ √ I = y<br />

for −1/ √ 2


336 N. Konno<br />

f(y) = d<br />

P (Y ≤ y)<br />

dy<br />

= 1<br />

2π {p1(k(y)) + p1(−k(y)) + p2(−π + k(y)) + p2(π − k(y))} dk(y)<br />

dy .<br />

Not<strong>in</strong>g that<br />

we have<br />

(1.1.14)<br />

y<br />

cos(k(y)) = − � , (1.1.15)<br />

1 − y2 dk(y)<br />

dy =<br />

1<br />

(1 − y2 ) � . (1.1.16)<br />

1 − 2y2 On the other hand, from (1.1.15) and<br />

�<br />

1 − 2y<br />

s<strong>in</strong>(k(y)) =<br />

2<br />

1 − y2 , I =1+cos2 (k(y)) = 1<br />

,<br />

1 − y2 we get<br />

� � �<br />

1 −y + i 1 − 2y2 v1(k(y)) = � ,<br />

2(1 + y) 1+y<br />

� � �<br />

1 −y − i 1 − 2y2 v1(−k(y)) = � ,<br />

2(1 + y) 1+y<br />

� � �<br />

1 y − i 1 − 2y2 v2(−π + k(y)) = � ,<br />

2(1 + y) −1 − y<br />

� � �<br />

1 y + i 1 − 2y2 v2(π − k(y)) = � .<br />

2(1 + y) −1 − y<br />

Comb<strong>in</strong><strong>in</strong>g these with (1.1.14) and (1.1.16) yields<br />

f(y) =<br />

1<br />

π(1 − y 2 ) � 1 − 2y 2 {1 − (|α|2 −|β| 2 + αβ + αβ)y}I (−1/ √ 2,1/ √ 2) (y).<br />

(1.1.17)<br />

In Katori, Fuj<strong>in</strong>o, and Konno (2005), the time-evolution equation of a<br />

one-dimensional quantum walker is exactly mapped to the three-dimensional<br />

Weyl equation for a zero-mass particle with sp<strong>in</strong> 1/2, <strong>in</strong> which each wave<br />

number k of walker’s wave function is mapped to a po<strong>in</strong>t q(k) <strong>in</strong> the threedimensional<br />

momentum space and q(k) makes a planar orbit as k changes<br />

its value <strong>in</strong> [−π, π). The <strong>in</strong>tegration over k provid<strong>in</strong>g the real-space wave


Quantum Walks 337<br />

function for a quantum walker corresponds to consider<strong>in</strong>g an orbital state of<br />

a Weyl particle, which is def<strong>in</strong>ed as a superposition (curvil<strong>in</strong>ear <strong>in</strong>tegration)<br />

of the energy-momentum eigenstates of a free Weyl equation along the orbit.<br />

The above density function (1.1.17) of quantum-walker’s pseudo-velocities <strong>in</strong><br />

the long-time limit is fully controlled by the shape of the orbit and how the<br />

orbit is embedded <strong>in</strong> the three-dimensional momentum space. The family of<br />

orbital states can b e regarded as a geometrical representation of the unitary<br />

group U(2).<br />

1.2 Three-State Case<br />

Introduction<br />

In this section we consider the three-state Grover walk. The Hadamard walk<br />

plays a key role <strong>in</strong> studies of the quantum walk and it has been analyzed<br />

<strong>in</strong> detail. Thus the generalization of the Hadamard walk is a fasc<strong>in</strong>at<strong>in</strong>g<br />

challenge. The simplest classical random walker on a one-dimensional lattice<br />

moves to the left or to the right with probability 1/2. On the other hand,<br />

the quantum walker accord<strong>in</strong>g to the Hadamard walk on a l<strong>in</strong>e moves both<br />

to the left and to the right. It is well known that the spatial distribution of<br />

the probability of f<strong>in</strong>d<strong>in</strong>g a particle governed by the Hadamard walk after<br />

long time is quite different from that of the classical random walk. However<br />

there are some common properties. In both walks the probability of f<strong>in</strong>d<strong>in</strong>g<br />

the particle at a fixed lattice site converges to zero as time goes to <strong>in</strong>f<strong>in</strong>ity.<br />

Let us consider a classical random walker that can stay at the same position<br />

with non-zero probability <strong>in</strong> a s<strong>in</strong>gle time-evolution. In the limit of long time,<br />

does the probability of f<strong>in</strong>d<strong>in</strong>g the walker at a fixed position converge to<br />

a positive value ? If the probability of stay<strong>in</strong>g at the same position <strong>in</strong> a<br />

s<strong>in</strong>gle step is very close to 1, the walker diffuses very slowly. However the<br />

probability of existence at a fixed position surely converges to zero if the<br />

jump<strong>in</strong>g probability is not zero. The classical random walker who can rema<strong>in</strong><br />

at the same position is essentially regarded as the same process as a walk<br />

that jumps right or left with probability 1/2, by scal<strong>in</strong>g the time. Is the same<br />

conclusion valid for a quantum walk ? The answer is “No”. We show that<br />

the profile of the Hadamard walk changes drastically by append<strong>in</strong>g only one<br />

degree of freedom to the <strong>in</strong>ner states.<br />

If the quantum particle <strong>in</strong> a three-state Grover walk exists at only one<br />

site <strong>in</strong>itially, the particle is trapped with high probability near the <strong>in</strong>itial<br />

position. Similar localization has already been seen <strong>in</strong> other quantum walks.<br />

The first simulation show<strong>in</strong>g the localization was presented by Mackay et al.<br />

(2002) <strong>in</strong> study<strong>in</strong>g the two-dimensional Grover walk. After that, more ref<strong>in</strong>ed<br />

simulations were performed by Tregenna et al. (2003) and an exact proof on<br />

the localization was given by Inui, Konishi, and Konno (2004). The second


338 N. Konno<br />

is found <strong>in</strong> the four-state Grover walk (Inui and Konno (2005)). In this walk,<br />

a particle moves not only to the nearest sites but also the second nearest<br />

sites accord<strong>in</strong>g to the four <strong>in</strong>ner states. The four-state Grover walk is also a<br />

generalized Hadamard walk, and it is similar to the three-state Grover walk.<br />

The significant difference between them is that the wave function of the<br />

four-state Grover walk does not converge, but that of the three-state Grover<br />

walk converges as time tends to <strong>in</strong>f<strong>in</strong>ity. We focus on localized stationary<br />

distribution of the three-state quantum particle and calcul ate it rigorously.<br />

On the other hand, Konno (2005c) proved that a weak limit distribution of a<br />

cont<strong>in</strong>uous-time rescaled two-state quantum walk has a similar form to that<br />

of the discrete-time one. In fact, both limit density functions for two-state<br />

quantum walks have two peaks at the two end po<strong>in</strong>ts of the supports. As a<br />

corollary, it is easily shown that a weak limit distribution of the correspond<strong>in</strong>g<br />

cont<strong>in</strong>uous-time three-state quantum walk has the same shape as that of the<br />

two-state walk. So the localization does not occur <strong>in</strong> the cont<strong>in</strong>uous-time case<br />

<strong>in</strong> contrast with the discrete-time case. In this situation the ma<strong>in</strong> aim of this<br />

section is to show rigorously that the localization occurs for a discrete-time<br />

three-state Grover walk. The results <strong>in</strong> this section are based on Inui, Konno,<br />

and Segawa (2005). Furthermore, we give a weak limit theorem for the threestate<br />

Grover walk and clarify the relation between the localization and the<br />

limit theorem.<br />

Def<strong>in</strong>ition of the Three-state Grover Walk<br />

The three-state quantum walk considered here is a k<strong>in</strong>d of generalized<br />

Hadamard walk on a l<strong>in</strong>e. The particle ruled by three-state quantum walk is<br />

characterized <strong>in</strong> the Hilbert space which is def<strong>in</strong>ed by a direct product of a<br />

chirality state space {|L〉, |0〉, |R〉} and a position space {|x〉 : x ∈ Z}. The<br />

chirality states are transformed at each time step by the follow<strong>in</strong>g unitary<br />

transformation:<br />

Let Ψn(x) = T [Ψ L n (x),Ψ0 n (x),ΨR n<br />

|L〉 = 1<br />

(−|L〉 +2|0〉 +2|R〉),<br />

3<br />

|0〉 = 1<br />

(2|L〉−|0〉 +2|R〉),<br />

3<br />

|R〉 = 1<br />

(2|L〉 +2|0〉−|R〉).<br />

3<br />

(x)] be the amplitude of the wave function of<br />

the particle correspond<strong>in</strong>g to the chiralities “L”, “0” and “R” attheposition<br />

x ∈ Z and the time n ∈{0, 1, 2,...}. We assume that a particle exists <strong>in</strong>itially<br />

at the orig<strong>in</strong>. Then the <strong>in</strong>itial quantum states are determ<strong>in</strong>ed by Ψ0(0) =<br />

T 2 2 2 [α, β, γ], where α, β, γ ∈ C with |α| + |β| + |γ| =1.<br />

Before we def<strong>in</strong>e the time evolution of the wave function, we <strong>in</strong>troduce the<br />

follow<strong>in</strong>g three operators:


Quantum Walks 339<br />

UL = 1<br />

⎡<br />

⎣<br />

3<br />

−122<br />

⎤<br />

0 0 0⎦<br />

,<br />

0 0 0<br />

U0 = 1<br />

⎡ ⎤<br />

0 0 0<br />

⎣ 2 −1 2⎦<br />

,<br />

3<br />

0 0 0<br />

UR = 1<br />

⎡ ⎤<br />

00 0<br />

⎣ 00 0 ⎦ .<br />

3<br />

22−1<br />

The three-state quantum walk considered here given by<br />

U (G,3) = UL + U0 + UR = 1<br />

⎡ ⎤<br />

−1 2 2<br />

⎣ 2 −1 2 ⎦<br />

3<br />

2 2 −1<br />

is often called three-state Grover walk. In general, n-state Grover walk is<br />

def<strong>in</strong>ed by a unitary matrix U (G,n) =[u (G,n) (i, j) :1≤ i, j ≤ n], where<br />

u (G,n) (i, j) isthe(i, j) component of the matrix U (G,n) and<br />

u (G,n) (i, i) =2/n − 1, u (G,n) (i, j) =2/n if i �= j.<br />

If the matrix UL is applied to the function Ψn(x), the “L”-component is<br />

(x), and<br />

selected after carry<strong>in</strong>g out the superimposition between Ψ L n (x),Ψ0 n<br />

Ψ R n (x). Similarly the “0”-component and “R”-component are selected <strong>in</strong><br />

U0Ψn(x) andURΨn(x).<br />

We now def<strong>in</strong>e the time evolution of the wave function by<br />

Ψn+1(x) =ULΨn(x +1)+U0Ψn(x)+URΨn(x − 1).<br />

One f<strong>in</strong>ds clearly that the chiralities “L” and“R” correspond to the left<br />

and the right, and the chirality “0” corresponds to the neutral state for the<br />

motion.<br />

Us<strong>in</strong>g the Fourier analysis, which is often used <strong>in</strong> the calculations of quantum<br />

walks, we obta<strong>in</strong> the wave function. The spatial Fourier transformation<br />

of Ψn(x) is def<strong>in</strong>ed by<br />

ˆΨn(k) = �<br />

e −ikx Ψn(x).<br />

x∈Z<br />

The dynamics of the wave function <strong>in</strong> the Fourier doma<strong>in</strong> is given by<br />

ˆΨn+1(k) = 1<br />

⎡<br />

⎣<br />

3<br />

eik 0 0<br />

0 1 0<br />

0 0 e−ik ⎤ ⎡ ⎤<br />

−1 2 2<br />

⎦ ⎣ 2 −1 2 ⎦ Ψn(k) ˆ<br />

2 2 −1<br />

≡ U(k) ˆ Ψn(k). (1.2.18)<br />

Thus the solution of (1.2.18) is ˆ Ψn(k) =U(k) n ˆ Ψ0(k). Let e iθj,k and |vj(k)〉<br />

be the eigenvalues of U(k) and the orthonormal eigenvector correspond<strong>in</strong>g to<br />

e iθj,k (j =1, 2, 3). S<strong>in</strong>ce the matrix U(k) is a unitary matrix, it is diagonalizable.<br />

Therefore the wave unction ˆ Ψn(k) is expressed by


340 N. Konno<br />

⎛<br />

3�<br />

ˆΨn(k) = ⎝ e<br />

j=1<br />

iθj,kn<br />

⎞<br />

|vj(k)〉〈vj(k)| ⎠ Ψ0(k), ˆ<br />

where ˆ Ψ0(k) = T [α, β, γ] ∈ C3 with |α| 2 + |β| 2 + |γ| 2 =1. The eigenvalues are<br />

given by<br />

⎧<br />

⎨ 0, j =1,<br />

θj,k = θk, j =2,<br />

⎩<br />

−θk, j=3,<br />

cos θk = − 1<br />

(2 + cos k),<br />

3<br />

s<strong>in</strong> θk = 1 �<br />

(5 + cos k)(1 − cos k), (1.2.19)<br />

3<br />

for k ∈ [−π, π). The eigenvectors of U(k) which form an orthonormal basis<br />

are obta<strong>in</strong>ed after some algebra:<br />

⎡<br />

�<br />

⎢<br />

|vj(k)〉 = ck(θj,k) ⎣<br />

where<br />

�<br />

ck(θ) =2<br />

1<br />

1+cos(θ − k) +<br />

1<br />

1+e i(θj,k −k)<br />

1<br />

1+e iθj,k 1<br />

1+e i(θj,k +k)<br />

⎤<br />

⎥<br />

⎦ , (1.2.20)<br />

1<br />

1+cosθ +<br />

�−1 1<br />

.<br />

1+cos(θ + k)<br />

In the above calculation, it is most important to emphasize that there is<br />

the eigenvalue “1” <strong>in</strong>dependently on the value of k. The eigenvalues of the<br />

Hadamard walk are given by e iθk and e i(π−θk) with the arguments satisfy<strong>in</strong>g<br />

s<strong>in</strong> θk =s<strong>in</strong>k/ √ 2 (see (1.1.13)). Therefore the eigenvalues do not take the<br />

value “1” except the special case k = 0. We have shown the existence of<br />

strongly degenerated eigenvalue such as “1” <strong>in</strong> (1.2.19) is necessary condition<br />

<strong>in</strong> the quantum walks show<strong>in</strong>g the localization (see also Inui, Konishi, and<br />

Konno (2004)).<br />

Let us number each chirality “L”,“0”, and “R” us<strong>in</strong>gl =1,2,and3,<br />

respectively. Then the wave function <strong>in</strong> real space is obta<strong>in</strong>ed by the <strong>in</strong>verse<br />

Fourier transform: for α, β, γ ∈ C with |α| 2 + |β| 2 + |γ| 2 =1,<br />

Ψn(x) =Ψn(x; α, β, γ)<br />

= 1<br />

� π<br />

ˆΨn(k)e<br />

2π −π<br />

ikx dk<br />

= 1<br />

⎛<br />

� π 3�<br />

⎝ e<br />

2π −π<br />

iθj,kn<br />

|vj(k)〉〈vj(k)| ˆ ⎞<br />

Ψ0(k) ⎠ e ikx dk<br />

j=1


Quantum Walks 341<br />

where<br />

with<br />

= T [Ψn(x;1;α, β, γ),Ψn(x;2;α, β, γ),Ψn(x;3;α, β, γ)]<br />

3�<br />

T<br />

=<br />

� Ψ j n (x;1;α, β, γ),Ψj n (x;2;α, β, γ),Ψj n (x;3;α, β, γ)� ,<br />

j=1<br />

Ψ j n(x; l; α, β, γ) = 1<br />

2π<br />

� π<br />

−π<br />

ck(θj,k)ϕk(θj,k,l)e i(θj,kn+kx) dk (l =1, 2, 3),<br />

ϕk(θ, l) =ζl,k(θ)[αζ1,k(θ)+βζ2,k(θ)+γζ3,k(θ)] (l =1, 2, 3),<br />

(1.2.21)<br />

(1.2.22)<br />

ζ1,k(θ) =(1+e i(θ−k) ) −1 , ζ2,k(θ) =(1+e iθ ) −1 , ζ3,k(θ) =(1+e i(θ+k) ) −1 ,<br />

(1.2.23)<br />

where z is conjugate of z ∈ C. Sometimes we omit the <strong>in</strong>itial qubit state<br />

[α, β, γ] suchasΨ j n(x; l) = Ψ j n(x; l; α, β, γ). The probability of f<strong>in</strong>d<strong>in</strong>g the<br />

particle at the position x and time n with the chirality l is given by Pn(x; l) =<br />

|Ψn(x; l)| 2 . Thus the probability of f<strong>in</strong>d<strong>in</strong>g the particle at the position x and<br />

time n is Pn(x) = � 3<br />

l=1 Pn(x; l).<br />

Time-averaged <strong>Probability</strong><br />

We focus our attention on the spatial distribution of the probability of f<strong>in</strong>d<strong>in</strong>g<br />

the particle after long time. Then (1.2.21) is rather complicated, but the<br />

value of limn→∞ Pn(x) can be calculated <strong>in</strong> follow<strong>in</strong>g subsections. We start<br />

by show<strong>in</strong>g a numerical result for the probability of f<strong>in</strong>d<strong>in</strong>g the particle at the<br />

orig<strong>in</strong> before carry<strong>in</strong>g out an analytical calculation. The value at the orig<strong>in</strong><br />

of the time-averaged probability is def<strong>in</strong>ed by<br />

¯P∞(0; α, β, γ) = lim<br />

N→∞<br />

�<br />

lim<br />

T →∞<br />

1<br />

T<br />

3�<br />

�<br />

T� −1<br />

Pn,N (0; l; α, β, γ) ,<br />

l=1 n=0<br />

where Pn,N (x; l; α, β, γ) is the probability of f<strong>in</strong>d<strong>in</strong>g a particle with the chirality<br />

l at the position x and time n on a cyclic lattice conta<strong>in</strong><strong>in</strong>g N sites.<br />

The time-averaged probability <strong>in</strong>troduced here was already used to study the<br />

quantum walk. In the case of the Hadamard walk on a cycle conta<strong>in</strong><strong>in</strong>g odd<br />

sites, the time-averaged probability takes a value 1/N <strong>in</strong>dependently of the<br />

<strong>in</strong>itial states (see Aharonov et al. (2001), Inui, Konishi, Konno, and Soshi<br />

(2005)). Therefore the time-averaged probability converges to zero <strong>in</strong> the<br />

limit N →∞. On the other hand, the time-averaged probability of quantum


342 N. Konno<br />

walks which exhibit the localization converges to a non-zero value. For this<br />

reason, we first calculate the time-averaged probability of three-state Grover<br />

walk, and we will show <strong>in</strong> the next subsection that the probability Pn(x)<br />

itself converges as well.<br />

We calculate the time-averaged probability at the orig<strong>in</strong> <strong>in</strong> three-state<br />

Grover walk on a cycle with N sites. We assume that the number N is odd.<br />

The argument of eigenvalues of three-state Grover walk with a f<strong>in</strong>ite N is<br />

θ j,2mπ/N for m ∈ [−(N −1)/2, (N −1)/2]. S<strong>in</strong>ce the eigenvalue correspond<strong>in</strong>g<br />

to m is the same as the eigenvalues correspond<strong>in</strong>g to −m, the wave function<br />

is formally expressed by<br />

Ψn,N (0; l; α, β, γ) =<br />

3�<br />

j=1<br />

(N−1)/2 �<br />

m=0<br />

cj,m,l(N)e iθN,j,mn ,<br />

where θN,j,m is θ j,2mπ/N. We note here that the coefficients cj,m,l(N) depend<br />

on the <strong>in</strong>itial state [α, β, γ], but we omit to describe the dependence. Then<br />

the probability Pn,N (0; α, β, γ) at the orig<strong>in</strong> is given by<br />

Pn,N (0; α, β, γ)<br />

3�<br />

=<br />

(N−1)/2 �<br />

c<br />

l1,l2,j1,j2=1 m1,m2=0<br />

∗ j1,m1,l1 (N)cj2,m2,l2(N)e i(θN,j 2 ,m2 −θN,j 1 ,m1 )n .<br />

The coefficients cj,m,l(N) are determ<strong>in</strong>ed from the product of eigenvectors.<br />

Although lengthy calculations are required to express the coefficients<br />

cj,m,l(N) explicitly, it is shown below that coefficients cj,m,l(N) except j =1<br />

do not contribute Pn,N (x). Not<strong>in</strong>g that the follow<strong>in</strong>g equation<br />

we have<br />

¯PN (0; α, β, γ) =<br />

T −1<br />

1<br />

lim<br />

T →∞ T<br />

n=0<br />

⎛�<br />

3�<br />

�<br />

⎜�<br />

⎝�<br />

�<br />

�<br />

l=1<br />

�<br />

�<br />

� 3�<br />

+ �<br />

�<br />

�<br />

j=2<br />

�<br />

e iθn =<br />

(N−1)/2<br />

�<br />

m=0<br />

cj,0,l(N)<br />

c1,m,l(N)<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

2<br />

+<br />

� 1,θ=0,<br />

0, θ�= 0,<br />

3�<br />

j=2<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

2<br />

(N−1)/2 �<br />

m=1<br />

|cj,m,l(N)| 2<br />

⎞<br />

⎟<br />

⎠ . (1.2.24)<br />

The difference between the first term and the other terms is caused by the<br />

difference of the degree of degenerate eigenvalues.<br />

Let φj,k(N) be the eigenvectors correspond<strong>in</strong>g to the eigenvalue e iθj,kn of<br />

the time-evolution matrix of the three-state Grover walk with a matrix size<br />

3N × 3N. They are easily obta<strong>in</strong>ed from |vj(k)〉 by


Quantum Walks 343<br />

|φj,k(N)〉 = 1<br />

√ 3N [|vj(k)〉,ω|vj(k)〉,ω 2 |vj(k)〉,...,ω N−1 |vj(k)〉],<br />

where ω = e 2πi/N . S<strong>in</strong>ce the coefficients cj,m,l(N) are proportional to the<br />

product of eigenvectors, the orders with respect to N of the first term and<br />

the second term <strong>in</strong> (1.2.24) are O(1) and O(N −1 ), respectively. Thus we can<br />

neglect the second term <strong>in</strong> the limit of N →∞. Us<strong>in</strong>g the eigenvectors <strong>in</strong><br />

(1.2.20), we have<br />

¯P∞(0; α, β, γ) =(5− 2 √ 6)(1 + |α + β| 2 + |β + γ| 2 − 2|β| 2 ).<br />

The time-averaged probability takes the maximum value 2(5−2 √ 6) at β =0.<br />

The component of ¯ P∞(0; α, β, γ) correspond<strong>in</strong>g to l =1, 2, 3 are respectively<br />

given by<br />

¯P∞(0; 1; α, β, γ) = |√6α − 2( √ 6 − 3)β +(12−5 √ 6)γ| 2<br />

,<br />

36<br />

¯P∞(0; 2; α, β, γ) = (√ 6 − 3) 2 |α + β + γ| 2<br />

,<br />

9<br />

¯P∞(0; 3; α, β, γ) = |√6γ − 2( √ 6 − 3)β +(12−5 √ 6)α| 2<br />

. (1.2.25)<br />

36<br />

We stress here that the time-averaged probability is not always positive. If<br />

α =1/ √ 6, β = −2/ √ 6andγ =1/ √ 6, then the time-averaged probability<br />

becomes zero, that is, ¯ P∞(0; 1/ √ 6, −2/ √ 6, 1/ √ 6) = 0.<br />

Stationary Distribution of the Particle<br />

We showed that the time-averaged probability of three-state Grover walk<br />

converges to non-zero values except special <strong>in</strong>itial states. In this subsection<br />

we calculate the limit P∗(x) ≡ limn→∞ Pn(x)(= limn→∞ Pn(x; α, β, γ)) rigorously<br />

and consider the dependence of P∗(x) onthepositionx.<br />

The wave function given by (1.2.21) is <strong>in</strong>f<strong>in</strong>ite superposition of the wave<br />

function ei(θj,kn+kx) . If the argument θj,k given <strong>in</strong> (1.2.19) for j =2, 3isnot<br />

absolute zero, then the eiθj,kn oscillates with high frequency <strong>in</strong> k-space for<br />

large n. On the other hand, ck(θj,k)ϕk(θj,k,l) <strong>in</strong> (1.2.22) changes smoothly<br />

with respect to k. Therefore we expect that the <strong>in</strong>tegration for j =2, 3<strong>in</strong><br />

(1.2.22) becomes small as the time <strong>in</strong>creases due to cancellation and converges<br />

to zero <strong>in</strong> the limit of n →∞.Thatis,<br />

Lemma 1.11. For fixed x ∈ Z,<br />

lim<br />

3�<br />

n→∞<br />

j=2<br />

Ψ j n (x; l; α, β, γ) =0, (1.2.26)<br />

for l =1, 2, 3 and α, β, γ ∈ C with |α| 2 + |β| 2 + |γ| 2 =1.


344 N. Konno<br />

Proof. Here we give an outl<strong>in</strong>e of the proof. A direct computation gives<br />

⎡<br />

3�<br />

⎣<br />

j=2<br />

Ψ j n (x;1;α, β, γ)<br />

Ψ j n(x;2;α, β, γ)<br />

Ψ j ⎤ ⎡ ⎤<br />

α<br />

⎦ ⎢ ⎥<br />

= M ⎣ β ⎦ ,<br />

n(x;3;α, β, γ) γ<br />

where M =(mij)1≤i,j≤3 with<br />

and<br />

m11 =3Jx,n + 1<br />

2 {Jx−1,n + Jx+1,n +(Kx−1,n − Kx+1,n)} ,<br />

m33 =3Jx,n + 1<br />

2 {Jx−1,n + Jx+1,n − (Kx−1,n − Kx+1,n)} ,<br />

m12 = −{Jx,n + Jx+1,n +(Kx,n − Kx+1,n)} ,<br />

m32 = −{Jx,n + Jx−1,n +(Kx,n − Kx−1,n)} ,<br />

m13 = −2Jx+1,n, m31 = −2Jx−1,n,<br />

m21 = −{Jx,n + Jx−1,n +(Kx−1,n − Kx,n)} ,<br />

m23 = −{Jx,n + Jx+1,n +(Kx+1,n − Kx,n)} ,<br />

m22 =4Jx,n,<br />

Jx,n = 1<br />

� π<br />

2π<br />

Kx,n = 1<br />

2π<br />

−π<br />

� π<br />

−π<br />

cos(kx)<br />

5+cosk cos(θkn)dk,<br />

cos(kx)<br />

� (5 + cos k)(1 − cos k) s<strong>in</strong>(θkn)dk.<br />

From the Riemann-Lebesgue lemma, we can show that<br />

lim<br />

n→∞ Jx,n =0, lim<br />

n→∞ (Kx,n − Kx+1,n) =0,<br />

for any x ∈ Z. Therefore we have the desired conclusion.<br />

From this lemma, the probability P∗(x) = P∗(x; α, β, γ) is determ<strong>in</strong>ed<br />

from the eigenvectors correspond<strong>in</strong>g to the eigenvalue “1”, that is, θ1,k =0<br />

(see (1.2.19)), and l-th component of P∗(x; l) =P∗(x; l; α, β, γ) isgivenby<br />

P∗(x; l; α, β, γ) = � � Ψ 1 n(x; l; α, β, γ) � � 2 . (1.2.27)<br />

Note that Ψ 1 n (x; l; α, β, γ) does not depend on time n,s<strong>in</strong>ceθ1,k =0.Bytransform<strong>in</strong>g<br />

<strong>in</strong>tegration <strong>in</strong> the left-hand side <strong>in</strong> (1.2.27) <strong>in</strong>to complex <strong>in</strong>tegral, we<br />

have


Quantum Walks 345<br />

P∗(x;1;α, β, γ) =|2αI(x)+βJ+(x)+2γK+(x)| 2 ,<br />

�<br />

�<br />

P∗(x;2;α, β, γ) = �<br />

β<br />

�αJ−(x)+ 2 L(x)+γJ+(x)<br />

�2<br />

�<br />

�<br />

� ,<br />

P∗(x;3;α, β, γ) =|2αK−(x)+βJ−(x)+2γI(x)| 2 , (1.2.28)<br />

where c = −5+2 √ 6(∈ (−1, 0)) and<br />

I(x) = 2c|x|+1<br />

c2 ,<br />

− 1<br />

L(x) =I(x − 1) + 2I(x)+I(x +1),<br />

J+(x) =I(x)+I(x +1), J−(x) =I(x − 1) + I(x),<br />

K+(x) =I(x +1), K−(x) =I(x − 1), (1.2.29)<br />

for any x ∈ Z. Therefore we have<br />

Theorem 1.12.<br />

P∗(x; α, β, γ) =<br />

3�<br />

P∗(x; l; α, β, γ) (x∈Z). l=1<br />

We should remark that it is confirmed ¯ P∞(0; l; α, β, γ) =P∗(0; l; α, β, γ)<br />

for l =1, 2, 3 by us<strong>in</strong>g (1.2.25), (1.2.28) and (1.2.29).<br />

Here we give an example. From (1.2.28) and (1.2.29), we obta<strong>in</strong><br />

P∗(0; i/ √ 2, 0, 1/ √ 2) = 4c2 (5c 2 +2c +5)<br />

(1 − c 2 ) 2<br />

=2(5− 2 √ 6) = 0.202 ...,<br />

(1.2.30)<br />

P∗(x; i/ √ 2, 0, 1/ √ 2) = 2(5c4 +2c 3 +10c 2 +2c +5)<br />

(1 − c 2 ) 2 c 2|x| , (1.2.31)<br />

for any x with |x| ≥1. Furthermore,<br />

0 < �<br />

P∗(x; i/ √ 2, 0, 1/ √ 2) = 1/ √ 6=0.408 ...


346 N. Konno<br />

have �<br />

x∈Z P∗(x) =0. The same conclusion can be obta<strong>in</strong>ed for the discretetime<br />

and cont<strong>in</strong>uous-time two-state quantum walks.<br />

Weak Limit Theorem<br />

In this subsection we give a weak limit theorem for the three-state Grover<br />

walk for any <strong>in</strong>itial state ϕ = T [α, β, γ] by us<strong>in</strong>g the method of Grimmett,<br />

Janson, and Scudo (2004). The unitary matrix of the walk is<br />

U = 1<br />

⎡ ⎤<br />

−1 2 2<br />

⎣ 2 −1 2 ⎦ .<br />

3<br />

2 2 −1<br />

So we have<br />

U(k) = 1<br />

⎡<br />

3<br />

⎣ −eik 2e ik 2e ik<br />

2 −1 2<br />

2e −ik 2e −ik −e −ik<br />

⎤<br />

⎦ .<br />

The eigenvalues and correspond<strong>in</strong>g eigenvectors of U(k) are<br />

where<br />

and<br />

where<br />

λ1(k) =1, λ2(k) =e iθk , λ3(k) =e −iθk ,<br />

cos θk = − 1<br />

3 (cos k +2), s<strong>in</strong> θk = 1�<br />

(5 + cos k)(1 − cos k),<br />

3<br />

v2(k) =<br />

v3(k) =<br />

v1(k) =<br />

⎡<br />

2<br />

⎣<br />

5+cosk<br />

1<br />

(1 + e −ik )/2<br />

e −ik<br />

⎤<br />

⎦ ,<br />

1<br />

�<br />

|1/w1(k)| 2 + |1/w2(k)| 2 + |1/w3(k)| 2<br />

⎡ ⎤<br />

1/w1(k)<br />

⎢ ⎥<br />

⎣ 1/w2(k) ⎦ ,<br />

1/w3(k)<br />

1<br />

�<br />

|1/w1(k)| 2 + |1/w2(k)| 2 + |1/w3(k)| 2<br />

⎡ ⎤<br />

1/w3(k)<br />

⎢ ⎥<br />

⎣ 1/w2(k) ⎦ ,<br />

1/w1(k)<br />

w1(k) =1+e i(θk−k) , w2(k) =1+e iθk , w3(k) =1+e i(θk+k) .<br />

Then we have


Quantum Walks 347<br />

h1(k) =0, h2(k) =<br />

s<strong>in</strong> k<br />

� (5 + cos k)(1 − cos k) ,<br />

s<strong>in</strong> k<br />

h3(k) =− � .<br />

(5 + cos k)(1 − cos k)<br />

From Theorem 1.10, we obta<strong>in</strong><br />

P (Y ≤ y) =<br />

3�<br />

�<br />

j=1<br />

{k∈[−π,π):hj(k)≤y}<br />

pj(k) dk<br />

2π .<br />

Let k(y) ∈ (0,π] be the unique solution of h2(k) fory≥0. Then we get<br />

� π<br />

P (Y ≤ y) = p1(k)<br />

−π<br />

dk<br />

2π +<br />

�� 0 � �<br />

π<br />

+ p2(k)<br />

−π k(y)<br />

dk<br />

2π +<br />

�� −k(y) � �<br />

π<br />

+ p3(k)<br />

−π 0<br />

dk<br />

2π .<br />

Let ˜ k(y)(= −k(y)) ∈ [−π, 0) be the unique solution of h2(k) fory ≤ 0.<br />

Similarly we have<br />

P (Y ≤ y) =<br />

� 0<br />

˜k(y)<br />

p2(k) dk<br />

2π +<br />

Therefore the limit measure is given by<br />

where<br />

Moreover<br />

� − ˜ k(y)<br />

0<br />

p3(k) dk<br />

2π .<br />

f(y) = d<br />

dy P (Y ≤ y) =∆δ0(y) − 1<br />

2π {p2(k(y)) + p3(−k(y))} dk(y)<br />

dy ,<br />

Remark that<br />

cos k(y) = 5y2 − 1<br />

cos θ k(y) = − 1+3y2<br />

1 − y2 , s<strong>in</strong> k(y) =2y� 2(1 − 3y2 )<br />

1 − y2 3(1 − y2 ) , s<strong>in</strong> θk(y) = 2�2(1 − 3y2 )<br />

3(1 − y2 )<br />

v2(k(y)) = 1<br />

2 √ ⎡ � ⎤<br />

2(1 − 3y2 �<br />

)+i(1 − 3y)<br />

⎣ 2(1 − 3y2 ) − 2i ⎦<br />

3 �<br />

,<br />

2(1 − 3y2 )+i(1 + 3y)<br />

v3(−k(y)) = 1<br />

2 √ ⎡ � ⎤<br />

2(1 − 3y2 �<br />

) − i(1 − 3y)<br />

⎣ 2(1 − 3y2 )+2i ⎦<br />

3 �<br />

.<br />

2(1 − 3y2 ) − i(1 + 3y)<br />

dk(y)<br />

dy<br />

2<br />

= −<br />

√ 2<br />

(1 − y2 ) � .<br />

1 − 3y2 ,<br />

.


348 N. Konno<br />

Therefore we have<br />

Theorem 1.13. If n →∞,then<br />

Xn<br />

n<br />

where Z has the follow<strong>in</strong>g measure:<br />

where<br />

and<br />

f(y) =∆(α, β, γ) δ0(y)+<br />

� π<br />

⇒ Z,<br />

√ 2(c0 + c1y + c2y 2 )<br />

2π(1 − y 2 ) � 1 − 3y 2 I (−1/ √ 3,1/ √ 3) (y),<br />

∆(α, β, γ) = p1(k)<br />

−π<br />

dk<br />

2π<br />

����� = α + β<br />

�<br />

�2<br />

� �<br />

�<br />

�<br />

2 � + �<br />

β �2<br />

�γ + �<br />

2 �<br />

� √<br />

6<br />

6<br />

+ℜ � �<br />

�<br />

(2α + β)(2γ + β) 1 − 5<br />

�<br />

√<br />

6 ,<br />

12<br />

c0 = |α + γ| 2 +2|β| 2 ,<br />

Here ℜ(z) is the real part of z ∈ C.<br />

c1 =2 � −|α − β| 2 + |γ − β| 2� ,<br />

c2 = |α − γ| 2 − 2ℜ((2α + β)(2γ + β)).<br />

Therefore by Theorem 1.13 we check<br />

∆(i/ √ 2, 0, 1/ √ 2) = �<br />

P∗(x; i/<br />

x∈Z<br />

√ 2, 0, 1/ √ 2) = 1/ √ 6,<br />

∆(1/ √ 3, 1/ √ 3, 1/ √ 3) = �<br />

P∗(x;1/<br />

x∈Z<br />

√ 3, 1/ √ 3, 1/ √ 3) = 3 − √ 6,<br />

∆(1/ √ 3, −1/ √ 3, 1/ √ 3) = �<br />

P∗(x;1/ √ 3, −1/ √ 3, 1/ √ 3) = (3 − √ 6)/9.<br />

In general, we see that<br />

x∈Z<br />

∆(α, β, γ) = �<br />

P∗(x; α, β, γ).<br />

x∈Z


Quantum Walks 349<br />

Conclusions and Discussions<br />

Some properties which are not observed <strong>in</strong> two-state quantum walks were<br />

found <strong>in</strong> three-state Grover walk. The particle which exists at the orig<strong>in</strong> splits<br />

<strong>in</strong> a superposition of three pieces. Two of the three parts leave for po<strong>in</strong>ts at<br />

<strong>in</strong>f<strong>in</strong>ity and the rema<strong>in</strong>der stays at the orig<strong>in</strong>. In contrast with the ord<strong>in</strong>ary<br />

Hadamard walk, the probability of f<strong>in</strong>d<strong>in</strong>g the particle at the orig<strong>in</strong> does not<br />

vanish for large time, and its maximum probability is 2(5−2 √ 6) = 0.202 ....<br />

A simple reason why the three-state Grover walk is different from two-state<br />

quantum walks is the difference <strong>in</strong> the degree of degenerate eigenvalues. The<br />

necessary condition of the localization is the existence of degenerate eigenvalues.<br />

And furthermore, each degree of degeneration must be proportional to<br />

the dimension of the Hilbert space. In addition, if the degenerate eigenvalue<br />

is 1 only, then the probability of f<strong>in</strong>d<strong>in</strong>g the particle can converges <strong>in</strong> the<br />

limit of n →∞. The four-state Grover walk exhibits the localization, but<br />

the probability of f<strong>in</strong>d<strong>in</strong>g the particle oscillates, because there are degenerate<br />

eigenvalues with values “1” and “−1”. On the other hand, the degenerated<br />

eigenvalue <strong>in</strong> three-state Grover walk unrelated to the wave number is only<br />

“1”, therefore the probability converges. Although no experiment exists about<br />

the quantum walks yet, the stationary properties of three-state Grover walk<br />

may be useful for comparison wit h theoretical results.<br />

We discuss a relation between the limit distribution for the orig<strong>in</strong>al threestate<br />

Grover walk Xn as time n →∞and that of the rescaled Xn/n <strong>in</strong> the<br />

same limit. When we consider the three-state Grover walk start<strong>in</strong>g from a<br />

mixture of three pure states T [1, 0, 0], T [0, 1, 0], and T [0, 0, 1] with probability<br />

1/3 respectively, we can obta<strong>in</strong> a weak limit probability distribution f(x) for<br />

the rescaled three-state quantum walk Xn/n as n →∞:<br />

f(x) = 1<br />

3 δ0(x)+<br />

√<br />

8<br />

3π(1 − x2 ) √ 1 − 3x2 I (−1/ √ 3,1/ √ 3) (x), (1.2.33)<br />

for x ∈ R, whereδ0(x) denotes the po<strong>in</strong>tmass at the orig<strong>in</strong> and I (a,b)(x) =1,<br />

if x ∈ (a, b), =0, otherwise. The above derivation is due to the method by<br />

Grimmett, Janson, and Scudo (2004). We should note that the first term <strong>in</strong><br />

the right-hand side of (1.2.33) corresponds to the localization for the orig<strong>in</strong>al<br />

three-state quantum walk. In fact, we have<br />

1<br />

3<br />

�<br />

x∈Z<br />

[P∗(x;1, 0, 0) + P∗(x;0, 1, 0) + P∗(x;0, 0, 1)] = 1<br />

3 .<br />

The last value 1/3 is noth<strong>in</strong>g but the coefficient 1/3 ofδ0(x). Moreover, the<br />

second term <strong>in</strong> the right-hand side of (1.2.33) has a similar weak limit density<br />

function for the same rescaled Hadamard walk with two <strong>in</strong>ner states start<strong>in</strong>g<br />

from a uniform random mixture of two pure states T [1, 0] and T [0, 1]:


350 N. Konno<br />

fH(x) =<br />

1<br />

π(1 − x 2 ) √ 1 − 2x 2 I (−1/ √ 2,1/ √ 2) (x),<br />

for x ∈ R. This case does not have a delta measure term correspond<strong>in</strong>g to a<br />

localization.<br />

1.3 Multi-State Case<br />

Rigorous results for multi-state quantum walks are very limited. Localization<br />

of one-dimensional four-state Grover walk (Inui and Konno (2005)) and twodimensional<br />

four-state quantum walks (Inui, Konishi, and Konno (2004)) was<br />

<strong>in</strong>vestigated. Moreover, Miyazaki, Katori, and Konno (2007) obta<strong>in</strong>ed weak<br />

limit theorems for a class of multi-state quantum walks <strong>in</strong> one dimension.<br />

2 Disordered Case<br />

2.1 Introduction<br />

A unitary matrix correspond<strong>in</strong>g to the evolution of the quantum walk is usually<br />

determ<strong>in</strong>istic and <strong>in</strong>dependent of the time step. In this chapter we study<br />

a random unitary matrix case motivated by numerical results of Mackay<br />

et al. (2002) and Ribeiro, Milman, and Mosseri (2004). We obta<strong>in</strong> a classical<br />

random walk from the quantum walk by <strong>in</strong>troduc<strong>in</strong>g an <strong>in</strong>dependent<br />

random fluctuation at each time step and perform<strong>in</strong>g an ensemble average<br />

<strong>in</strong> a rigorous way based on a comb<strong>in</strong>atorial approach. The time evolution<br />

of the quantum walk is given by the follow<strong>in</strong>g random unitary matrices<br />

{Un : n =1, 2,...}:<br />

� �<br />

an bn<br />

Un = ,<br />

cn dn<br />

where an,bn,cn,dn ∈ C. The subscript n <strong>in</strong>dicates the time step. The unitarity<br />

of Un gives<br />

|an| 2 + |cn| 2 = |bn| 2 + |dn| 2 =1,<br />

ancn + bndn =0,cn = −△nbn, dn = △nan, (2.1.1)<br />

where z is the complex conjugate of z ∈ C and △n =detUn = andn −<br />

bncn with |△n| =1. Put wn =(an,bn,cn,dn). Let {wn : n =1, 2,...} be<br />

<strong>in</strong>dependent and identically distributed (or i.i.d. for short) on some space,<br />

(for example, [0, 2π)) with<br />

E(|a1| 2 )=E(|b1| 2 )=1/2, (2.1.2)<br />

E(a1c1) =0. (2.1.3)


Quantum Walks 351<br />

Remark that (2.1.2) implies E(|c1| 2 ) = E(|d1| 2 )=1/2, and (2.1.3) gives<br />

E(b1d1) = 0 by us<strong>in</strong>g (2.1.1). The set of <strong>in</strong>itial qubit states for the quantum<br />

walk is given by<br />

Φ = � ϕ = T [α, β] ∈ C 2 : |α| 2 + |β| 2 =1 � .<br />

Moreover we assume that {wn : n =1, 2,...} and {α, β} are <strong>in</strong>dependent.<br />

We call the above process a disordered quantum walk <strong>in</strong> this chapter. Let R<br />

be the set of the real numbers. Then we consider the follow<strong>in</strong>g two cases:<br />

Case I: an,bn,cn,dn ∈ R (n =1, 2,...)andE(αβ + αβ) =0.<br />

Case II: E(|α| 2 )=1/2 andE(αβ) =0.<br />

We should note that Case I corresponds to an example given by Ribeiro,<br />

Milman, and Mosseri (2004) and Case II corresponds to an example given by<br />

Mackay et al. (2002), respectively. The numerical simulations by two groups<br />

suggest that the probability distribution of a disordered quantum walk converges<br />

to a b<strong>in</strong>omial distribution by averag<strong>in</strong>g over many trials. So the ma<strong>in</strong><br />

purpose of this chapter is to prove the above numerical results by us<strong>in</strong>g a<br />

comb<strong>in</strong>atorial (path <strong>in</strong>tegral) approach. Results <strong>in</strong> this chapter are based on<br />

Konno (2005b). Ma<strong>in</strong> result here (Theorem 2.1) shows that the expectation<br />

of the probability distribution for the disordered quantum walk becomes the<br />

probability distribution of a classical symmetric random walk. However our<br />

result does not treat a crossover from quantum to classical walks.<br />

2.2 Def<strong>in</strong>ition of Disordered Quantum Walk<br />

First we divide Un <strong>in</strong>to two matrices:<br />

Pn =<br />

� an bn<br />

0 0<br />

�<br />

, Qn =<br />

� 0 0<br />

cn dn<br />

with Un = Pn + Qn. The important po<strong>in</strong>t is that Pn (resp. Qn) represents<br />

that the particle moves to the left (resp. right) at each time step n. Byus<strong>in</strong>g<br />

Pn and Qn, we def<strong>in</strong>e the dynamics of the disordered quantum walk on the<br />

l<strong>in</strong>e. To do so, we def<strong>in</strong>e a (4N +2)× (4N +2)matrixUnby ⎡<br />

⎤<br />

0 Pn 0 ... ... 0 Qn<br />

⎢ Qn ⎢ 0 Pn 0 ... ... 0 ⎥<br />

⎢ 0 Qn 0 Pn 0 ... 0 ⎥<br />

⎢<br />

U n = ⎢<br />

.<br />

⎢ . .. . .. . .. . .. . ..<br />

⎥<br />

. ⎥ ,<br />

⎥<br />

⎢ 0 ... 0 Qn 0 Pn 0 ⎥<br />

⎣ 0 ... ... 0 Qn 0 Pn ⎦<br />

Pn 0 ... ... 0 Qn 0<br />

�<br />

,


352 N. Konno<br />

where 0 = O2 is the 2 × 2 zero matrix. Note that Pn and Qn satisfy<br />

PnP ∗ n + QnQ ∗ n = P ∗ nPn + Q ∗ nQn = I2,<br />

PnQ ∗ n = QnP ∗ n = Q∗n Pn = P ∗ nQn = O2,<br />

where ∗ means the adjo<strong>in</strong>t operator and I2 is the 2×2 unit matrix. The above<br />

relations imply that U n is also a unitary matrix. Us<strong>in</strong>g U n, we can def<strong>in</strong>e a<br />

disordered quantum walk Xn at time n start<strong>in</strong>g from ϕ ∈ Φ. Let Ψn(x) be<br />

the two component vector of amplitudes of the particle be<strong>in</strong>g at site x and<br />

at time n. Then the probability of {Xn = x} is def<strong>in</strong>ed by<br />

P (Xn = x) =�Ψn(x)� 2 .<br />

Remark that {Xn = x} is an event generated by {wi : i =1, 2,...,n} and<br />

{α, β}. The unitarity of U m(m =1, 2,...,n) ensures<br />

n�<br />

x=−n<br />

P (Xn = x) =�U nU n−1 ···U 1ϕ� 2 = �ϕ� 2 = |α| 2 + |β| 2 =1,<br />

for any 1 ≤ n ≤ N. That is, the amplitude always def<strong>in</strong>es a probability<br />

distribution for the location.<br />

2.3 Result<br />

We now give a comb<strong>in</strong>atorial expression for the probability distribution of the<br />

disordered quantum walk. For fixed l and m with l + m = n and −l + m = x,<br />

we <strong>in</strong>troduce<br />

Ξn(l, m) = �<br />

P ln Q mn P ln−1 Q mn−1 ···P l1 Q m1<br />

lj,mj<br />

summed over all lj,mj ≥ 0 satisfy<strong>in</strong>g l1 + ···+ ln = l, m1 + ···+ mn = m<br />

and lj + mj = 1. We should note that<br />

Ψn(x) =Ξn(l, m)ϕ.<br />

For example, <strong>in</strong> the case of P (X4 =0),wehave<br />

Ξ4(2, 2) = P4Q3Q2P1 + Q4P3P2Q1 + P4P3Q2Q1<br />

+P4Q3P2Q1 + Q4P3Q2P1 + Q4Q3P2P1.<br />

To compute Ξn(l, m), we <strong>in</strong>troduce the follow<strong>in</strong>g useful random matrices:<br />

� �<br />

cn dn<br />

Rn = ,<br />

0 0<br />

�<br />

0<br />

Sn =<br />

�<br />

0<br />

.<br />

an bn


Quantum Walks 353<br />

In general, we obta<strong>in</strong> the follow<strong>in</strong>g table of products of the matrices Pj,Qj,Rj<br />

and Sj (j =1, 2,...): for any m, n ≥ 1,<br />

Pn Qn Rn Sn<br />

Pm amPn bmRn amRn bmPn<br />

Qm cmSn dmQn cmQn dmSn<br />

Rm cmPn dmRn cmRn dmPn<br />

Sm amSn bmQn amQn bmSn<br />

where PmQn = bmRn, for example. From this table, we obta<strong>in</strong><br />

Ξ4(2, 2) = b4d3c2P1 + c4a3b2Q1 +(a4b3d2 + b4c3b2)R1 +(c4b3c2 + d4c3a2)S1.<br />

We should note that P1,Q1,R1 and S1 form an orthonormal basis of the<br />

vector space of complex 2×2 matrices with respect to the trace <strong>in</strong>ner product<br />

〈A|B〉 =tr(A ∗ B). So Ξn(l, m) has the follow<strong>in</strong>g form:<br />

Ξn(l, m) =pn(l, m)P1 + qn(l, m)Q1 + rn(l, m)R1 + sn(l, m)S1. (2.3.4)<br />

In general, explicit forms for pn(l, m),qn(l, m),rn(l, m) andsn(l, m) arecomplicated.<br />

From (2.3.4), we have<br />

� �<br />

pn(l, m)a1 + rn(l, m)c1 pn(l, m)b1 + rn(l, m)d1<br />

Ξn(l, m) =<br />

.<br />

qn(l, m)c1 + sn(l, m)a1 qn(l, m)d1 + sn(l, m)b1<br />

Remark that pn(l, m),qn(l, m),rn(l, m),sn(l, m) depend only on {wi : i =<br />

2, 3,...n}, so they are <strong>in</strong>dependent of w1, wherewi =(ai,bi,ci,di). Moreover<br />

it should be noted that {wi : i = 1, 2,...} and {α, β} are <strong>in</strong>dependent.<br />

Therefore we obta<strong>in</strong><br />

P (Xn = x) =||Ξn(l, m)ϕ|| 2<br />

Then we see that<br />

= � |pn(l, m)a1 + rn(l, m)c1| 2 + |qn(l, m)c1 + sn(l, m)a1| 2� |α| 2<br />

+ � |pn(l, m)b1 + rn(l, m)d1| 2 + |qn(l, m)d1 + sn(l, m)b1| 2� |β| 2<br />

�<br />

+ (pn(l, m)a1 + rn(l, m)c1)(pn(l, m)b1 + rn(l, m)d1)<br />

�<br />

+(qn(l, m)c1 + sn(l, m)a1)(qn(l, m)d1 + sn(l, m)b1) αβ<br />

�<br />

+ (pn(l, m)a1 + rn(l, m)c1)(pn(l, m)b1 + rn(l, m)d1)<br />

�<br />

+(qn(l, m)c1 + sn(l, m)a1)(qn(l, m)d1 + sn(l, m)b1) αβ<br />

= C1|α| 2 + C2|β| 2 + C3αβ + C4αβ.<br />

E(C1|α| 2 )= 1<br />

2 E� |pn(l, m)| 2 + |sn(l, m)| 2 + |qn(l, m)| 2 + |rn(l, m)| 2� E(|α| 2 ),


354 N. Konno<br />

s<strong>in</strong>ce {pn(l, m),qn(l, m),rn(l, m),sn(l, m)} and w1 are <strong>in</strong>dependent, {wi : i =<br />

1, 2,...} and {α, β} are <strong>in</strong>dependent, (2.1.2) and (2.1.3). Similarly, we have<br />

E(C2|β| 2 )= 1<br />

2 E� |pn(l, m)| 2 + |sn(l, m)| 2 + |qn(l, m)| 2 + |rn(l, m)| 2� E(|β| 2 ).<br />

CaseI:FirstwenotethatC4 = C3. Soan,bn,cn,dn ∈ R gives C3 = C4. Then<br />

the condition αβ + αβ = 0 implies C3αβ + C4αβ = 0. Therefore we obta<strong>in</strong><br />

E(P (Xn = x))<br />

= E(||Ξn(l, m)ϕ|| 2 )<br />

= 1<br />

2 E� |pn(l, m)| 2 + |sn(l, m)| 2 + |qn(l, m)| 2 + |rn(l, m)| 2� E(|α| 2 )<br />

+ 1<br />

2 E� |pn(l, m)| 2 + |sn(l, m)| 2 + |qn(l, m)| 2 + |rn(l, m)| 2� E(|β| 2 )<br />

= 1 �<br />

E(|pn(l, m)|<br />

2<br />

2 )+E(|qn(l, m)| 2 )+E(|rn(l, m)| 2 )+E(|sn(l, m)| 2 ) � ,<br />

s<strong>in</strong>ce |α| 2 + |β| 2 =1.<br />

Case II: E(αβ) =E(αβ) = 0 gives the same conclusion as above <strong>in</strong> a similar<br />

fashion.<br />

To get a feel for the follow<strong>in</strong>g ma<strong>in</strong> result, we will look at the concrete<br />

case where n =4andl = m =2.<br />

E(P (X ϕ<br />

4 =0))<br />

= E(||Ξ4(2, 2)ϕ|| 2 )<br />

= 1 �<br />

E(|p4(2, 2)|<br />

2<br />

2 )+E(|q4(2, 2)| 2 )+E(|r4(2, 2)| 2 )+E(|s4(2, 2)| 2 ) �<br />

= 1 �<br />

E(|b4|<br />

2<br />

2 )E(|d3| 2 )E(|c2| 2 )+E(|c4| 2 )E(|a3| 2 )E(|b2| 2 )<br />

+E(|a4| 2 )E(|b3| 2 )E(|d2| 2 )+E(|b4| 2 )E(|c3| 2 )E(|b2| 2 )<br />

+E(a4b4)E(b3c3)E(d2b2)+E(a4b4)E(b3c3)E(d2b2)<br />

+E(|c4| 2 )E(|b3| 2 )E(|c2| 2 )+E(|d4| 2 )E(|c3| 2 )E(|a2| 2 )<br />

+E(c4d4)E(b3c3)E(c2a2)+E(c4d4)E(b3c3)E(c2a2) �<br />

= 1 1 2 2 1<br />

+ + + =<br />

16 16 16 16 24 � �<br />

4<br />

,<br />

2<br />

s<strong>in</strong>ce E(|a1| 2 ) = E(|b1| 2 ) = E(|c1| 2 ) = E(|d1| 2 ) = 1/2, and E(a1c1) =<br />

E(b1d1) =0. This result corresponds to P (Y o<br />

4 =0)=� 4<br />

2<br />

� /2 4 , where Y o n<br />

denotes the classical symmetric random walk start<strong>in</strong>g from the orig<strong>in</strong>. We<br />

generalize the above example as follows:<br />

Theorem 2.1. We assume that a disordered quantum walk start<strong>in</strong>g from ϕ<br />

satisfies Case I or Case II. For n =0, 1, 2,..., and x = −n, −(n − 1),...,<br />

n − 1,n, with n + x is even, we have


Quantum Walks 355<br />

E(P (Xn = x)) = P (Y o 1<br />

n = x) =<br />

2n � �<br />

n<br />

.<br />

(n + x)/2<br />

Proof. For n =0, 1, 2,..., and x = −n, −(n − 1),...,n− 1,n, with n + x is<br />

even, let l =(n− x)/2, m=(n + x)/2 andM = � � n<br />

(n+x)/2 . In general, as <strong>in</strong><br />

the case of n =4withl = m =2,wehave<br />

P (Xn = x) =E(||Ξn(l, m)ϕ|| 2 )<br />

= 1<br />

M�<br />

2<br />

j=1<br />

E(|u (j)<br />

n |2 )E(|u (j)<br />

n−1 |2 ) ···E(|u (j)<br />

2 |2 )+R(w1,w2,...,wn),<br />

where u (j)<br />

∈{ai,bi,ci,di} for any i =2, 3,...,n, j =1, 2,...,M and wk =<br />

i<br />

(ak,bk,ck,dk) for any k =1, 2,...,n. Then E(|a1| 2 )=E(|b1| 2 )=E(|c1| 2 )=<br />

E(|d1| 2 )=1/2 gives<br />

E(P (Xn = x)) = M<br />

+ R(w1,w2,...,wn).<br />

2n To get the desired conclusion, it suffices to show that R(w1,w2,...,wn) =0.<br />

When there are more than two terms <strong>in</strong> pn(l, m) orsn(l, m), we have to<br />

consider two cases: there is a k such that<br />

···Pk+1PkPk−1 ···P1, ···Qk+1PkPk−1 ···P1.<br />

For example, s4(2, 2) = d4c3a2 + c4b3c2 case is k =1,s<strong>in</strong>ce<br />

Q4Q3P2P1, Q4P3Q2P1.<br />

Then the correspond<strong>in</strong>g pn(l, m) orsn(l, m) aregivenby<br />

Therefore we have<br />

···ak+1akak−1 ···a2, ···ck+1akak−1 ···a2.<br />

E � (···ak+1akak−1 ···a2)(···ck+1akak−1 ···a2) �<br />

= E(···) E(ak+1ck+1) E(|ak| 2 ) E(|ak−1| 2 ) ···E(|a2| 2 )=0,<br />

s<strong>in</strong>ce E(a1c1) = 0. When there are more than two terms <strong>in</strong> qn(l, m) or<br />

rn(l, m), we have to consider two cases: there is a k such that<br />

···Qk+1QkQk−1 ···Q1, ···Pk+1QkQk−1 ···Q1.<br />

Therefore, a similar argument holds. So the proof is complete.<br />

The above theorem implies that the expectation of the probability distribution<br />

for the disordered quantum walk is noth<strong>in</strong>g but the probability<br />

distribution of a classical symmetric random walk.


356 N. Konno<br />

2.4 Examples<br />

The first example was <strong>in</strong>troduced and studied by Ribeiro, Milman, and<br />

Mosseri (2004). This corresponds to Case I:<br />

� �<br />

cos(θn) s<strong>in</strong>(θn)<br />

Un =<br />

,<br />

s<strong>in</strong>(θn) − cos(θn)<br />

where {θn : n =1, 2,...} are i.i.d. on [0, 2π) with<br />

E(cos 2 (θ1)) = E(s<strong>in</strong> 2 (θ1)) = 1/2, E(cos(θ1)s<strong>in</strong>(θ1)) = 0.<br />

Here we give two concrete cases:<br />

(i) θ1 is the uniform distribution on [0, 2π),<br />

(ii) P (θ1 = ξ) =P (θ1 = π/2+ξ) =1/2 forsomeξ ∈ [0,π).<br />

For an <strong>in</strong>itial qubit state, we choose a non-random ϕ = T [α, β] ∈ Ψ with<br />

|α| = |β| =1/ √ 2, and αβ + αβ =0.<br />

The second example was given by Mackay et al. (2002). This corresponds<br />

to Case II:<br />

Un = 1<br />

√ 2<br />

� 1 e iθn<br />

e −iθn −1<br />

where {θn : n =1, 2,...} are i.i.d. on [0, 2π) with<br />

�<br />

,<br />

E(cos(θ1)) = E(s<strong>in</strong>(θ1)) = 0.<br />

Moreover an <strong>in</strong>itial qubit state is chosen as ϕ = T [α, β] ∈ Ψ with E(|α| 2 )=<br />

1/2 andE(αβ) = 0. For example, both θ1 and θ∗ are uniform distributions<br />

on [0, 2π), and they are <strong>in</strong>dependent of each other. Let α = cos(θ∗) and<br />

β =cos(θ∗). So the above conditions hold.<br />

F<strong>in</strong>ally we discuss an example studied by Shapira et al. (2003). Let Wn =<br />

{Xn,Yn,Zn}, where {Wn : n =1, 2,...} is i.i.d., and {Xn,Yn,Zn} is also<br />

i.i.d., moreover, Xn is normally distributed with mean zero and variance σ 2 ,<br />

for some σ>0. Their case is<br />

where<br />

Vn = 1 √ 2<br />

Un = 1<br />

√ 2<br />

�<br />

cos(Rn)+iZn s<strong>in</strong>(Rn)<br />

Rn<br />

(−Yn + iXn) s<strong>in</strong>(Rn)<br />

Rn<br />

� �<br />

1 1<br />

× Vn,<br />

1 −1<br />

(Yn + iXn) s<strong>in</strong>(Rn)<br />

Rn<br />

cos(Rn) − iZn s<strong>in</strong>(Rn)<br />

Rn<br />

and Rn = � X2 n + Y 2 n + Z2 n . Then a direct computation gives<br />

�<br />

,


Quantum Walks 357<br />

E(|a1| 2 )=E(|b1| 2 )=1/2,<br />

E(a1c1) = 1 1<br />

+<br />

6 3 (1 − 4σ2 )e −2σ2<br />

≡ µ(σ). (2.4.5)<br />

We see that (2.4.5) implies 0 < µ( √ 3/2) = 0.0179 ... ≤ µ(σ) ≤ 1/2(=<br />

limσ↓0 µ(σ)) for any σ>0. Note that the limit σ ↓ 0 corresponds to the<br />

Hadamard walk case. So their example does not satisfy our condition (2.1.3).<br />

3 Reversible Cellular Automata<br />

3.1 Introduction<br />

In this chapter we consider some properties of a one-dimensional reversible<br />

cellular automaton (RCA) derived from a quantum walk on a l<strong>in</strong>e. Results<br />

<strong>in</strong> this chapter are based on Konno, Mitsuda, Soshi, and Yoo (2004). We<br />

present necessary and sufficient conditions on the <strong>in</strong>itial state for the existence<br />

of some conserved quantities of the RCA. One is the expectation of the<br />

distribution (the 1st moment), and the other is the squared norm of the distribution<br />

(the 0th moment). The former corresponds to the symmetry of the<br />

distribution, one of the results below (see Theorem 3.2) implies that the distribution<br />

is symmetric for any time step if and only if its expectation is zero<br />

for any time step. We should note that our RCA is different from the quantum<br />

cellular automaton studied by Gröss<strong>in</strong>g and Zeil<strong>in</strong>ger (1988a, 1988b). In<br />

Gröss<strong>in</strong>g and Zeil<strong>in</strong>ger (1988b), they found a conservation law which connects<br />

the strength of the mix<strong>in</strong>g of locally <strong>in</strong>teract<strong>in</strong>g states and the periodicity of<br />

the global structures. Their conservation law also differs from ours. For a recent<br />

review on quantum cellular automata, see Aoun and Tarifi (2004), and<br />

Schumacher and Wernerfor (2004), for examples. It is well known that cellular<br />

automata are models based on simple rules which upon determ<strong>in</strong>istic time<br />

evolution exhibit various complex behavior (Wolfram (2002)). Our <strong>in</strong>vestigation<br />

of the RCA might shed some additional light on the analysis of this<br />

complex behavior. Moreover, by apply<strong>in</strong>g our results to the quantum walk<br />

case, we may obta<strong>in</strong> more detailed <strong>in</strong>formation on quantum walks. The purpose<br />

of this chapter is to <strong>in</strong>vestigate some correspondence relation between<br />

quantum walks and RCAs. Then, we analyze some properties and conserved<br />

quantities of a quantum walk by exam<strong>in</strong><strong>in</strong>g a RCA. Thus, properties that are<br />

hidden or less obvious <strong>in</strong> a quantum walk may be more obvious when study<strong>in</strong>g<br />

the RCA. This method then provides a useful tool for study<strong>in</strong>g quantum<br />

walks.<br />

In this chapter, we consider an extension of the Hadamard walk as follows:<br />

� �<br />

cos θ s<strong>in</strong> θ<br />

H(θ) =<br />

,<br />

s<strong>in</strong> θ − cos θ


358 N. Konno<br />

where θ ∈ [0, 2π). Here we study ma<strong>in</strong>ly the case of θ ∈ [0,π/2) for the sake<br />

of simplicity. Remark that when θ = π/4 it becomes the Hadamard walk,<br />

that is, H(π/4) = H.<br />

The def<strong>in</strong>ition of the quantum walk gives<br />

for<br />

Ψ L n+1 (x) =aΨ L n (x +1)+bΨ R n<br />

Ψ R n+1 (x) =cΨ L n (x − 1) + dΨ R n<br />

U =<br />

� �<br />

ab<br />

∈ U(2).<br />

cd<br />

(x +1),<br />

(x − 1),<br />

Furthermore, we can rewrite the above equations to uncouple both chirality<br />

components:<br />

Ψ L n+2 (x) =aΨ L n+1 (x +1)+dΨ L n+1 (x − 1) − (ad − bc)Ψ L n (x),<br />

Ψ R n+2 (x) =aΨ R n+1 (x +1)+dΨ R n+1 (x − 1) − (ad − bc)Ψ R n (x).<br />

That is, the left and right chiralities both satisfy the follow<strong>in</strong>g partial difference<br />

equation:<br />

ηn+2(x) =aηn+1(x +1)+dηn+1(x − 1) − (ad − bc)ηn(x), (3.1.1)<br />

where ηn(x) ∈ C stands for the probability amplitudes of left or right chirality<br />

at time n ∈ Z+ at location x ∈ Z, whereZ+ is the set of non-negative<br />

<strong>in</strong>tegers. The above argument appeared <strong>in</strong> Knight, Roldán, and Sipe (2003a,<br />

2003b) (see also page 279 <strong>in</strong> Gudder (1988)). In other words, (3.1.1) implies<br />

a dynamical <strong>in</strong>dependence of the evolution of the two chiralities L and R.<br />

Thus there are two essentially <strong>in</strong>dependent walks, coupled only by the first<br />

two steps. After that the two walks can behave <strong>in</strong>dependently of each other.<br />

In this chapter, we call (3.1.1) a reversible cellular automata (RCA), s<strong>in</strong>ce<br />

(3.1.1) implies that ηn(x) can also be determ<strong>in</strong>ed by ηn+1(·) andηn+2(·).<br />

It should be noted that if we put (η0(0),η1(−1),η1(1)) = (Ψ L 0 (0),aΨL 0 (0)+<br />

bΨ R 0 (0), 0) on the <strong>in</strong>itial state of (3.1.1), then ηn(x) =Ψ L n (x), that is, ηn(x)<br />

represents the left chirality of the quantum walk. Similarly, an <strong>in</strong>itial qubit<br />

state of (3.1.1), that is, (η0(0),η1(−1),η1(1)) = (Ψ R 0 (0), 0,cΨL 0 (0) + dΨ R 0 (0))<br />

gives the right chirality of the quantum walk, namely, ηn(x) =Ψ R n (x).<br />

Here we consider a more general sett<strong>in</strong>g concern<strong>in</strong>g <strong>in</strong>itial states. The<br />

study on the dependence of some important quantities on the <strong>in</strong>itial state is<br />

one of the essential parts, so we def<strong>in</strong>e the set of <strong>in</strong>itial states for the RCA<br />

as follows:<br />

�Φ = { �ϕ =(η0(0),η1(−1),η1(1)) ≡ (α, β, γ) :α, β, γ ∈ C} .


Quantum Walks 359<br />

To dist<strong>in</strong>guish our <strong>in</strong>itial state for the RCA considered here from the <strong>in</strong>itial<br />

qubit state for quantum walks <strong>in</strong> our previous chapters, we put “∼” <strong>in</strong>Φ etc.<br />

From now on we consider only RCA’s def<strong>in</strong>ed by H(θ):<br />

ηn+2(x) =cosθ[ηn+1(x +1)− ηn+1(x − 1)] + ηn(x). (3.1.2)<br />

McGuigan (2003) studied some classes of quantum cellular automata. In his<br />

sett<strong>in</strong>g, (3.1.2) belongs to a class of fermionic quantum cellular automata<br />

whose update equation is given by<br />

ηn+2(x) =f(ηn+1(x +1),ηn(x),ηn+1(x − 1)) + ηn(x). (3.1.3)<br />

Our case is f(x, y, z) =(x − z)cosθ. By us<strong>in</strong>g (3.1.2), Romanelli et al. (2004)<br />

analyzed <strong>in</strong> detail discrete-time one-dimensional quantum walks by separat<strong>in</strong>g<br />

the quantum evolution <strong>in</strong>to Markovian and <strong>in</strong>terference terms.<br />

Now we def<strong>in</strong>e a distribution (<strong>in</strong> general, non-probability distribution) of<br />

RCA ηn at time n ∈ Z+ by<br />

{|ηn(x)| 2 : x ∈ Z}.<br />

In this chapter, we call {|ηn(x)| 2 : x ∈ Z} the distribution of ηn.<br />

One of the ma<strong>in</strong> results of this chapter is to give the follow<strong>in</strong>g necessary<br />

and sufficient conditions of the symmetry of distribution of ηn for any time<br />

step n.<br />

3.2 Symmetry of Distribution<br />

First we present the follow<strong>in</strong>g useful lemma to prove Theorem 3.2.<br />

Lemma 3.1. (i) We suppose that the <strong>in</strong>itial state is<br />

where α, β ∈ C. Then we have<br />

�ϕ =(η0(0),η1(−1),η1(1)) ≡ (α, β, −β),<br />

for any x ∈ Z and n ∈ Z+.<br />

(ii) We suppose that the <strong>in</strong>itial state is<br />

where β ∈ R, ξ ∈ R. Then we have<br />

for any x ∈ Z+ and n ∈ Z+.<br />

ηn(x) =(−1) n ηn(−x), (3.2.4)<br />

�ϕ =(η0(0),η1(−1),η1(1)) ≡ (0,β,e iξ β),<br />

ηn(x) =(−1) n+1 e iξ ηn(−x), (3.2.5)


360 N. Konno<br />

Proof. We prove the Lemma by <strong>in</strong>duction on the time step n. The proof of<br />

part (ii) is essentially the same as that of part (i), so we omit it.<br />

First we consider n =0, 1 case. We can easily check that (3.2.4) holds<br />

for n =0, 1. Next we assume that (3.2.4) holds for time n = m and m +1<br />

(m ≥ 0). Then we have<br />

ηm+2(x) =cosθ[ηm+1(x +1)− ηm+1(x − 1)] + ηm(x)<br />

=cosθ[(−1) m+1 ηm+1(−x − 1) − (−1) m+1 ηm+1(−x +1)]<br />

+(−1) m ηm(−x)<br />

=(−1) m+1 cos θ[ηm+1(−x − 1) − ηm+1(−x +1)]+(−1) m ηm(−x)<br />

=(−1) m+2 {cos θ[ηm+1(−x +1)−ηm+1(−x − 1)] + ηm(−x)}<br />

=(−1) m+2 ηm+2(−x).<br />

Remark that the first and last equalities come from (3.1.2). The second equality<br />

uses the <strong>in</strong>duction hypothesis. So we have shown that (3.2.4) is also correct<br />

for time n = m + 2. The proof is complete.<br />

We <strong>in</strong>troduce the follow<strong>in</strong>g three classes:<br />

�Φ⊥ = { �ϕ ∈ � Φ : β + γ = 0 (3.2.6)<br />

or |β| = |γ|(> 0), and α = 0 (3.2.7)<br />

or |β| = |γ|(> 0),α�= 0, and θβ + θγ − 2θα = π (mod 2π)},<br />

�Φs = { �ϕ ∈<br />

(3.2.8)<br />

� Φ : |ηn(x)| = |ηn(−x)| for any n ∈ Z+ and x ∈ Z},<br />

�<br />

�Φ0 = �ϕ ∈ � ∞�<br />

�<br />

Φ :<br />

.<br />

x=−∞<br />

x|ηn(x)| 2 =0 foranyn ∈ Z+<br />

Here θz is the argument of z ∈ C with z �= 0. It is noted that if �ϕ ∈ � Φs,<br />

then the distribution (<strong>in</strong> general, non-probability distribution) of ηn becomes<br />

symmetric for any time n ∈ Z+.<br />

Theorem 3.2. For the RCA def<strong>in</strong>ed by (3.1.2), we have<br />

�Φ⊥ = � Φs = � Φ0.<br />

Proof. First we see that the def<strong>in</strong>itions of � Φs and � Φ0 give immediately<br />

�Φs ⊂ � Φ0. (3.2.9)<br />

Next, we prove � Φ⊥ ⊂ � Φs. The proof is divided <strong>in</strong>to the follow<strong>in</strong>g three<br />

cases.<br />

Case 1. If the <strong>in</strong>itial states are satisfy<strong>in</strong>g (3.2.6), then we can obta<strong>in</strong> the<br />

follow<strong>in</strong>g relation from Lemma 3.1 (i):<br />

|ηn(x)| = |ηn(−x)|. (3.2.10)


Quantum Walks 361<br />

Case 2. We assume that the <strong>in</strong>itial states satisfy (3.2.7). By us<strong>in</strong>g Lemma 3.1<br />

(ii), we can obta<strong>in</strong> (3.2.10).<br />

Case 3. We assume that the <strong>in</strong>itial states satisfy (3.2.8). Let<br />

�ϕ1 =(|α|e iθα , |β|e iθβ , |β|e i(π−θβ+2θα) ).<br />

Then<br />

�ϕ2 = e −iθα �ϕ1 =(|α|, |β|e i(θβ−θα) i(π−θβ+θα)<br />

, |β|e ).<br />

We put θ = θα − θβ, then�ϕ2 =(|α|, |β|eiθ , |β|ei(π−θ) ). Therefore, real and<br />

imag<strong>in</strong>ary parts of �ϕ2 become<br />

ℜ( �ϕ2) =(|α|, |β| cos θ, −|β| cos θ),<br />

ℑ( �ϕ2) =(0, |β| s<strong>in</strong> θ, |β| s<strong>in</strong> θ),<br />

respectively. Here if we take ℜ( �ϕ2) as an <strong>in</strong>itial state of the RCA, then we<br />

can see that the distribution of ηn is symmetric by us<strong>in</strong>g Lemma 3.1 (i).<br />

Similarly, us<strong>in</strong>g Lemma 3.1 (ii), the <strong>in</strong>itial state ℑ( �ϕ2) also gives a symmetric<br />

distribution. From �ϕ2 = ℜ( �ϕ2)+ℑ( �ϕ2) ∈ � Φs, we can get �ϕ1 ∈ � Φs. Therefore<br />

for any �ϕ ∈ � Φ⊥, wehave|ηn(x)| = |ηn(−x)| for any x ∈ Z and n ∈ Z+. Then<br />

we conclude<br />

F<strong>in</strong>ally, a direct computation gives<br />

where m(n) ≡<br />

�Φ⊥ ⊂ � Φs. (3.2.11)<br />

m(1) = |γ| 2 −|β| 2 ,<br />

m(2) = 2 cos 2 θ(|γ| 2 −|β| 2 ),<br />

m(3) = 1<br />

2 (3 cos2 2θ +2cos2θ +1)(|γ| 2 −|β| 2 )<br />

∞�<br />

x=−∞<br />

then �ϕ ∈ � Φ⊥. Sowehave<br />

Comb<strong>in</strong><strong>in</strong>g (3.2.10) - (3.2.12) gives<br />

so the proof is complete.<br />

− 1<br />

s<strong>in</strong> θ s<strong>in</strong> 2θ {α(β + γ)+α(β + γ)},<br />

2<br />

x|ηn(x)| 2 . The above equations imply that if �ϕ ∈ � Φ0<br />

�Φ0 ⊂ � Φ⊥. (3.2.12)<br />

�Φ⊥ ⊂ � Φs ⊂ � Φ0 ⊂ � Φ⊥,


362 N. Konno<br />

3.3 Conserved Quantity for RCA<br />

From now on we assume that 0 0,<br />

�Φ∗(c) = � Φ(c).<br />

The above theorem implies that � Φ∗(c) gives the necessary and sufficient conditions<br />

we want to know.<br />

Proof. Case 1: � Φ(c) ⊂ � Φ∗(c). A little algebra reveals<br />

||η0|| 2 = |α| 2 , (3.3.17)<br />

||η1|| 2 = |β| 2 + |γ| 2 ,<br />

||η2||<br />

(3.3.18)<br />

2 = |α| 2 +2cos 2 θ(|β| 2 + |γ| 2 ) − cos 2 θ(βγ + βγ)<br />

− cos θ(α(β − γ)+α(β − γ)), (3.3.19)<br />

||η3|| 2 =2cos 2 θ|α| 2 + � 2cos 4 θ +(1−2cos 2 θ) 2� (|β| 2 + |γ| 2 )<br />

+2 cos 2 θ(1 − 2cos 2 θ)(βγ + βγ)<br />

+cosθ(1 − 3cos 2 θ){α(β − γ)+α(β − γ)}. (3.3.20)


Quantum Walks 363<br />

From (3.3.17) - (3.3.19), we have<br />

α(β − γ)+α(β − γ) =cosθ{2c − (βγ + βγ)}. (3.3.21)<br />

Comb<strong>in</strong><strong>in</strong>g (3.3.21) with (3.3.17), (3.3.18), (3.3.20) implies<br />

s<strong>in</strong> θ cos θ(βγ + βγ)=0,<br />

so we have βγ + βγ =0,s<strong>in</strong>ce0


364 N. Konno<br />

We def<strong>in</strong>e ϕ = ϕ(ξ) ∈ R by λ+(ξ) =eiξϕ ,thatis,<br />

�<br />

cos ϕ = 1 − cos2 θ s<strong>in</strong> 2 ξ, s<strong>in</strong> ϕ = − cos θ s<strong>in</strong> ξ.<br />

So λ−(ξ) =−e −iξϕ . In this sett<strong>in</strong>g, we have the follow<strong>in</strong>g result: for any<br />

n ∈ Z+,<br />

||ηn|| 2 = ||A|| 2 ∗ + ||B|| 2 ∗ +(−1) n {〈e <strong>in</strong>ϕ A|e −<strong>in</strong>ϕ B〉 + 〈e −<strong>in</strong>ϕ B|e <strong>in</strong>ϕ A〉}.<br />

A direct computation gives<br />

||A|| 2 ∗ = ||B||2 ∗<br />

= 1<br />

4s<strong>in</strong>θ<br />

Moreover we have<br />

�<br />

(|α| 2 + |β| 2 + |γ| 2 ) −{α(β − γ)+α(β − γ)}<br />

4π{〈e <strong>in</strong>ϕ A|e −<strong>in</strong>ϕ B〉 + 〈e −<strong>in</strong>ϕ B|e <strong>in</strong>ϕ A〉}<br />

� 2π<br />

0<br />

� �<br />

1 − s<strong>in</strong> θ<br />

cos θ<br />

� � �<br />

2<br />

1 − s<strong>in</strong> θ<br />

−(βγ + βγ)<br />

.<br />

cos θ<br />

= |α| 2 cos(2(n − 1)ϕ)<br />

0 cos2 dξ<br />

ϕ<br />

�� 2π<br />

cos ξ s<strong>in</strong>((2n − 1)ϕ)<br />

+(αγ − αβ)i<br />

0 cos2 � 2π<br />

s<strong>in</strong> ξ s<strong>in</strong>((2n − 1)ϕ)<br />

dξ + i<br />

ϕ<br />

0 cos2 ϕ<br />

�� 2π<br />

cos ξ s<strong>in</strong>((2n − 1)ϕ)<br />

+(αβ − αγ)i<br />

0 cos2 � 2π<br />

s<strong>in</strong> ξ s<strong>in</strong>((2n − 1)ϕ)<br />

dξ − i<br />

ϕ<br />

0 cos2 ϕ<br />

−(|β| 2 + |γ| 2 � 2π<br />

cos(2nϕ)<br />

)<br />

0 cos2 ϕ dξ<br />

�� 2π<br />

cos(2ξ)cos(2nϕ)<br />

−βγ<br />

0 cos2 � 2π<br />

s<strong>in</strong>(2ξ)cos(2nϕ)<br />

dξ − i<br />

ϕ<br />

0 cos2 �<br />

dξ<br />

ϕ<br />

�� 2π<br />

cos(2ξ)cos(2nϕ)<br />

−βγ<br />

cos2 � 2π<br />

s<strong>in</strong>(2ξ)cos(2nϕ)<br />

dξ + i<br />

ϕ<br />

cos2 �<br />

dξ ,<br />

ϕ<br />

for any n ∈ Z+. Furthermore we have the follow<strong>in</strong>g equation:<br />

||ηn|| 2 = 1<br />

�<br />

(|α|<br />

2s<strong>in</strong>θ<br />

2 + |β| 2 + |γ| 2 � �<br />

1 − s<strong>in</strong> θ<br />

) −{α(β − γ)+α(β − γ)}<br />

cos θ<br />

� � �<br />

2<br />

1 − s<strong>in</strong> θ<br />

−(βγ + βγ)<br />

cos θ<br />

+ (−1)n<br />

�<br />

|α|<br />

π<br />

2<br />

� 0<br />

cos(2(n − 1)x)<br />

θ−π/2 cos x � cos2 x − s<strong>in</strong> 2 θ dx<br />

0<br />

�<br />

dξ<br />

�<br />


Quantum Walks 365<br />

−{α(β − γ)+α(β − γ)} 1<br />

� 0<br />

s<strong>in</strong> x s<strong>in</strong>((2n − 1)x)<br />

cos θ θ−π/2 cos x � cos2 x − s<strong>in</strong> 2 θ dx<br />

−(|β| 2 + |γ| 2 � 0<br />

cos(2nx)<br />

)<br />

θ−π/2 cos x � cos2 x − s<strong>in</strong> 2 θ dx<br />

−(βγ + βγ)<br />

�<br />

2<br />

cos θ<br />

� 0<br />

−<br />

θ−π/2<br />

� 0<br />

θ−π/2<br />

cos(2nx) � cos 2 x − s<strong>in</strong> 2 θ<br />

cos x<br />

cos(2nx)<br />

cos x � cos 2 x − s<strong>in</strong> 2 θ dx<br />

dx<br />

��<br />

, (3.3.22)<br />

for any n ∈ Z+. The condition “|α| 2 = |β| 2 + |γ| 2 = c, α(β − γ)+α(β − γ) =<br />

2c cos θ, βγ + βγ =0”gives<br />

||ηn|| 2 = c + (−1)nc π<br />

� 0<br />

cos(2(n − 1)x) − 2s<strong>in</strong>xs<strong>in</strong>((2n − 1)x) − cos(2nx)<br />

×<br />

θ−π/2<br />

cos x � cos2 x − s<strong>in</strong> 2 dx,<br />

θ<br />

for any n ∈ Z+. On the other hand, a little algebra reveals that<br />

cos(2(n − 1)x) − 2s<strong>in</strong>x s<strong>in</strong>((2n − 1)x) − cos(2nx) =0,<br />

for any n ∈ Z+. Therefore we conclude that<br />

||ηn|| 2 = c (n ∈ Z+),<br />

that is, � Φ∗(c) ⊂ � Φ(c). The proof of Theorem 3.3 is complete.<br />

By (3.3.22) and the Riemann-Lebesgue lemma, we have<br />

lim<br />

n→∞ ||ηn|| 2 = 1<br />

2s<strong>in</strong>θ<br />

�<br />

(|α| 2 + |β| 2 + |γ| 2 ) −{α(β − γ)+α(β − γ)}<br />

� �<br />

� �2 �<br />

1 − s<strong>in</strong> θ<br />

1 − s<strong>in</strong> θ<br />

×<br />

− (βγ + βγ)<br />

. (3.3.23)<br />

cos θ<br />

cos θ<br />

The above equation will be used <strong>in</strong> the next section. As a corollary to Theorem<br />

3.3, we can show that the quantity is not conserved <strong>in</strong> any RCA with a<br />

symmetric distribution.<br />

Corollary 3.4. For any c>0,<br />

�Φs ∩ � Φ(c) =φ.


366 N. Konno<br />

Proof. We consider the follow<strong>in</strong>g three cases.<br />

Case 1. When β + γ =0,from|β| 2 + |γ| 2 = c, wehave<br />

β =<br />

�<br />

c<br />

2 eiθβ �<br />

c<br />

, γ = −<br />

2 eiθβ . (3.3.24)<br />

However (3.3.24) and βγ + βγ = 0 contradict each other, s<strong>in</strong>ce βγ + βγ =<br />

−c (< 0).<br />

Case 2. “|β| = |γ|(> 0), and α =0”and“|α| 2 = c (> 0)” contradict each<br />

other.<br />

Case 3. We assume that “|β| = |γ|(> 0),α �= 0, and θβ + θγ − 2θα =<br />

π (mod 2π)”. From now on we omit “mod 2π”. Then<br />

can be rewritten as<br />

α(β − γ)+α(β − γ) =2c cos θ<br />

|α|{|β| cos(θα − θβ) −|γ| cos(θα − θγ)} = c cos θ. (3.3.25)<br />

On the other hand, |α| = √ c, |β| = |γ| = � c/2. So (3.3.25) becomes<br />

cos(θα − θβ) − cos(θα − θγ) = √ 2cosθ. (3.3.26)<br />

Comb<strong>in</strong><strong>in</strong>g (3.3.26) with θβ + θγ − 2θα = π implies<br />

� �<br />

θβ − θγ cos θ<br />

s<strong>in</strong><br />

= − √ . (3.3.27)<br />

2<br />

2<br />

We see that βγ+βγ =0givesθβ−θγ = ±π/2, so (3.3.27) becomes cos θ = ±1.<br />

Then we have a contradiction, s<strong>in</strong>ce we assumed that 0


Quantum Walks 367<br />

respectively. By us<strong>in</strong>g s<strong>in</strong> θ �= 0(s<strong>in</strong>ce0


368 N. Konno<br />

lim<br />

n→∞ ||Ψ R n ||2 = 1<br />

�<br />

(1+ cos<br />

2s<strong>in</strong>θ<br />

2 θ)|αr| 2 +s<strong>in</strong> 2 θ|αl| 2 − s<strong>in</strong> θ cos θ(αlαr + αlαr)<br />

−{2cosθ|αr| 2 � �<br />

1 − s<strong>in</strong> θ<br />

− s<strong>in</strong> θ(αlαr + αlαr)}<br />

cos θ<br />

�<br />

.<br />

For example, if we choose θ ∈ (0,π/2) and ϕ = T [1, 0] (asymmetric case),<br />

then<br />

lim<br />

n→∞ ||Ψ L n ||2 s<strong>in</strong> θ<br />

=1−<br />

2 ∈<br />

� �<br />

1<br />

, 1 , lim<br />

2 n→∞ ||Ψ R n ||2 s<strong>in</strong> θ<br />

=<br />

2 ∈<br />

�<br />

0, 1<br />

�<br />

.<br />

2<br />

4 Quantum Cellular Automata<br />

4.1 Introduction<br />

Patel, Raghunathan, and Rungta (2005a) constructed a quantum walk on a<br />

l<strong>in</strong>e without us<strong>in</strong>g a co<strong>in</strong> toss <strong>in</strong>struction, and analyzed the asymptotic behavior<br />

of the walk on the l<strong>in</strong>e and its escape probability with an absorb<strong>in</strong>g<br />

wall. In fact the quantum walk <strong>in</strong>vestigated by them can be considered as<br />

a class of quantum cellular automata on the l<strong>in</strong>e (see Meyer (1996, 1997,<br />

1998), for examples), so we call their quantum walk a quantum cellular automaton<br />

(QCA) <strong>in</strong> this chapter. Results here are based on Hamada, Konno,<br />

and Segawa (2005).<br />

At a first glance, the QCA looks different from the quantum walk. However,<br />

we show that there exists a one-to-one correspondence between them <strong>in</strong> a<br />

more general sett<strong>in</strong>g. The purpose of this chapter is to clarify this connection.<br />

Once the connection is well understood, the result by Patel, Raghunathan,<br />

and Rungta (2005a) is straightforwardly obta<strong>in</strong>ed.<br />

4.2 Def<strong>in</strong>ition of QCA<br />

We def<strong>in</strong>e the dynamics of a one-dimensional QCA <strong>in</strong>clud<strong>in</strong>g the model <strong>in</strong>vestigated<br />

by Patel, Raghunathan, and Rungta (2005a). Let η (m)<br />

n (x)(∈ C) be<br />

the amplitude of the QCA at time n ∈ Z+ and at location x ∈ Z start<strong>in</strong>g<br />

from m ∈ Z, thatis,η (m)<br />

0 (m) =1andη (m)<br />

0 (x) =0ifx �= m. Moreover,let<br />

and<br />

ζ (m:±)<br />

n (x) =|αη (m)<br />

n (x)+βη (m±1)<br />

n<br />

ζ (m:±)<br />

n<br />

=(ζ (m:±)<br />

n (x) :x ∈ Z),<br />

(x)| 2 ,


Quantum Walks 369<br />

where α, β ∈ C with |α| 2 + |β| 2 =1. As we will show later, the ζ (m:±)<br />

n (x) is<br />

equivalent to a probability distribution of a quantum walk at time n, where<br />

the pair (α, β) corresponds to the <strong>in</strong>itial qubit state of the quantum walk.<br />

The evolution of the QCA on the l<strong>in</strong>e is given by<br />

where U is the unitary matrix<br />

U =<br />

η (m)<br />

n+1 = Uη(m) n ,<br />

... −3 −2 −1 0 +1 +2 +3 +4 ...<br />

⎛ . ⎞<br />

. .. · · · · · · · · ...<br />

⎜<br />

⎟<br />

−3 ⎜ ... b a 0 0 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−2 ⎜ ... a b c d 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−1 ⎜ ... d c b a 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

0 ⎜ ... 0 0 a b c d 0 0 ... ⎟<br />

⎜<br />

⎟,<br />

+1 ⎜ ... 0 0 d c b a 0 0 ... ⎟<br />

⎜<br />

⎟<br />

+2 ⎜ ... 0 0 0 0 a b c d ... ⎟<br />

⎜<br />

⎟<br />

+3 ⎜ ... 0 0 0 0 d c b a ... ⎟<br />

⎜<br />

⎟<br />

+4 ⎝ ... 0 0 0 0 0 0 a b ... ⎠<br />

.<br />

.<br />

. ... · · · · · · · ·<br />

..<br />

with a, b, c, d ∈ C and η (m)<br />

n<br />

η (m)<br />

n<br />

is the configuration<br />

= T [...,η (m)<br />

n (−1),η(m)<br />

n (0),η(m)<br />

n (+1),...],<br />

for any n ∈ Z+. Let ||u|| 2 = �∞ x=−∞ |u(x)|2 . The unitarity of U ensures that<br />

if ||η (m)<br />

0 || =1, then ||η (m)<br />

n || =1foranyn∈Z+. Furthermore, if ||ζ (m:±)<br />

0 || =1,<br />

then ||ζ (m:±)<br />

n || =1foranyn∈Z+. A little algebra reveals that U is unitary<br />

if and only if<br />

|a| 2 + |b| 2 + |c| 2 + |d| 2 =1, (4.2.1)<br />

ad + ad + bc + bc =0, (4.2.2)<br />

ac + bd =0, (4.2.3)<br />

ab + ab =0, (4.2.4)<br />

cd + cd =0, (4.2.5)<br />

where z is a complex conjugate of z ∈ C. Here we consider a, b, c, d satisfy<strong>in</strong>g<br />

(4.2.1) - (4.2.5). Trivial cases are “|a| =1,b = c = d = 0”, “|b| =1,a = c =<br />

d = 0”, “|c| =1,a = b = d = 0”, and “|d| =1,a = b = c = 0”. For other<br />

cases, we have the follow<strong>in</strong>g five types:<br />

Type I: |b| 2 + |c| 2 =1,bc + bc =0,bc�= 0,a= d =0.<br />

Type II: |a| 2 + |b| 2 =1,ab + ab =0,ab�= 0,c= d =0.<br />

Type III: |c| 2 + |d| 2 =1,cd + cd =0,cd�= 0,a= b =0.


370 N. Konno<br />

Type IV: |a| 2 + |d| 2 =1,ad + ad =0,ad�= 0,b= c =0.<br />

Type V: a, b, c, d satisfy<strong>in</strong>g (4.2.1) - (4.2.5) and abcd �= 0.<br />

Let supp[ζ (m:±)<br />

n ]={x ∈ Z : ζ (m:±)<br />

n (x) > 0}. Then, it is easily seen that<br />

for any n ∈ Z+, supp[ζ (0:±)<br />

n ] ⊂{−2, −1, 0, 1} <strong>in</strong> Type I, and supp[ζ (0:±)<br />

n ] ⊂<br />

{−1, 0, 1, 2} <strong>in</strong> Type II. So both Types I and II are also trivial cases. To<br />

<strong>in</strong>vestigate non-trivial Types III - V, we <strong>in</strong>troduce a quantum walk <strong>in</strong> the<br />

next section.<br />

We see that a direct computation implies that (a, b, c, d) satisfy<strong>in</strong>g (4.2.1)<br />

- (4.2.5) has the follow<strong>in</strong>g form:<br />

(a, b, c, d) =e iδ (cos θ cos φ, −i cos θ s<strong>in</strong> φ, s<strong>in</strong> θ s<strong>in</strong> φ, i s<strong>in</strong> θ cos φ), (4.2.6)<br />

where θ, φ, δ ∈ [0, 2π). From now on, we assume that (a, b, c, d) hastheabove<br />

form. Remark that the case studied by Patel, Raghunathan, and Rungta<br />

(2005a) is δ = π/2andθ = φ = π/4, that is, (a, b, c, d) =(i/2, 1/2,i/2, −1/2),<br />

which belongs to Type V.<br />

4.3 Def<strong>in</strong>ition of A-type and B-type Quantum Walks<br />

The time evolution of both one-dimensional A-type and B-type quantum<br />

walks is given by the follow<strong>in</strong>g unitary matrix:<br />

� �<br />

′ ′ a b<br />

U = ,<br />

c ′ d ′<br />

where a ′ ,b ′ ,c ′ ,d ′ ∈ C. Sowehave|a ′ | 2 + |b ′ | 2 = |c ′ | 2 + |d ′ | 2 =1,a ′ c ′ + b ′ d ′ =<br />

0, c ′ = −△b ′ ,d ′ = △a ′ , where △ =detU = a ′ d ′ − b ′ c ′ with |△| =1.<br />

Let |L〉 = T [1, 0] and |R〉 = T [0, 1]. For an A-type quantum walk, each<br />

co<strong>in</strong> performs the evolution:<br />

|L〉 → U|L〉 = a ′ |L〉 + c ′ |R〉, |R〉 → U|R〉 = b ′ |L〉 + d ′ |R〉,<br />

at each time step for which that co<strong>in</strong> is active, where L and R can be respectively<br />

thought of as the head and tail states of the co<strong>in</strong>, or equivalently<br />

as an <strong>in</strong>ternal chirality state of the particle. The value of the co<strong>in</strong> controls<br />

the direction <strong>in</strong> which the particle moves. When the co<strong>in</strong> shows L, the particle<br />

moves one unit to the left, when it shows R, it moves one unit to the<br />

right. Then a B-type quantum walk is also def<strong>in</strong>ed <strong>in</strong> a similar way as we will<br />

state later. Thus the quantum walk can be considered as a quantum version<br />

of a classical random walk with an additional degree of freedom called the<br />

chirality which takes values left and right.<br />

The amplitude of the location of the particle is def<strong>in</strong>ed by a 2-vector ∈ C 2<br />

at each location at any time n. The probability that the particle is at location<br />

x is given by the square of the modulus of the vector at x. Forthej-type


Quantum Walks 371<br />

quantum walk (j = A, B), let Ψj,n(x) denote the amplitude at time n at<br />

location x where<br />

� L Ψj,n (x)<br />

Ψj,n(x) =<br />

Ψ R j,n (x)<br />

�<br />

,<br />

with the chirality be<strong>in</strong>g left (upper component) or right (lower component).<br />

For each j = A and B, the dynamics of Ψj,n(x) forthej-type quantum<br />

walk start<strong>in</strong>g from the orig<strong>in</strong> with an <strong>in</strong>itial qubit state ϕ = T [α, β], (where<br />

α, β ∈ C and |α| 2 + |β| 2 = 1), is presented as the follow<strong>in</strong>g transformation:<br />

where<br />

�<br />

′ ′ a b<br />

PA =<br />

0 0<br />

Ψj,n+1(x) =PjΨj,n(x +1)+QjΨj,n(x − 1), (4.3.7)<br />

� �<br />

0 0<br />

, QA =<br />

c ′ d ′<br />

�<br />

and PB =<br />

�<br />

′ a 0<br />

c ′ � �<br />

′ 0 b<br />

, QB =<br />

0<br />

0 d ′<br />

�<br />

.<br />

It is noted that U = Pj + Qj (j = A, B). The unitarity of U ensures that<br />

the amplitude always def<strong>in</strong>es a probability distribution for the location. From<br />

(4.3.7), we see that the unitary matrix of the system is described as<br />

⎡<br />

⎤<br />

. .. .<br />

⎢ .<br />

.<br />

.<br />

.<br />

.<br />

. ⎥<br />

⎢ ... O Pj O O O ...<br />

⎥<br />

⎢<br />

⎥<br />

⎢ ... Qj O Pj O O ... ⎥<br />

⎢<br />

⎥<br />

⎢ ... O Qj O Pj O ... ⎥<br />

⎢ ... O O Qj O Pj ...<br />

⎥<br />

⎢<br />

⎥<br />

⎢ ... O O O Qj O ... ⎥<br />

⎣<br />

⎦<br />

.<br />

... . . . . . ..<br />

with O =<br />

� �<br />

00<br />

,<br />

00<br />

for j = A and B. Remark that the A-type (resp. B-type) quantum walk is<br />

called an A-type (resp. a G-type) quantum walk <strong>in</strong> Konno (2002b).<br />

4.4 Connection Between QCA and A-type<br />

Quantum Walk<br />

To beg<strong>in</strong> with, we <strong>in</strong>vestigate a relation between the QCA and the A-type<br />

quantum walk. To do so, the unitary matrix of the QCA


372 N. Konno<br />

U =<br />

is rewritten as<br />

where<br />

... −3 −2 −1 0 +1 +2 +3 +4 ...<br />

⎛ . ⎞<br />

. .. · · · · · · · · ...<br />

⎜<br />

⎟<br />

−3 ⎜ ... b a 0 0 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−2 ⎜ ... a b c d 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−1 ⎜ ... d c b a 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

0 ⎜ ... 0 0 a b c d 0 0 ... ⎟<br />

⎜<br />

⎟<br />

+1 ⎜ ... 0 0 d c b a 0 0 ... ⎟<br />

⎜<br />

⎟<br />

+2 ⎜ ... 0 0 0 0 a b c d ... ⎟<br />

⎜<br />

⎟<br />

+3 ⎜ ... 0 0 0 0 d c b a ... ⎟<br />

⎜<br />

⎟<br />

+4 ⎝ ... 0 0 0 0 0 0 a b ... ⎠<br />

.<br />

.<br />

. ... · · · · · · · ·<br />

..<br />

U =<br />

PA =<br />

... −1 0 +1 +2 ...<br />

⎛ . ⎞<br />

. .. . . . . .<br />

⎜<br />

⎟<br />

−1 ⎜ ... TA QA O O ... ⎟<br />

⎜<br />

⎟<br />

0 ⎜ ... PA TA QA O ... ⎟<br />

⎜<br />

⎟,<br />

+1 ⎜ ... O PA TA QA ... ⎟<br />

⎜<br />

⎟<br />

+2 ⎝ ... O O PA TA ... ⎠<br />

.<br />

. ... . . . . ..<br />

� �<br />

dc<br />

, TA =<br />

00<br />

� �<br />

ba<br />

, QA =<br />

ab<br />

� �<br />

00<br />

.<br />

cd<br />

We consider a pair (2x−1, 2x) <strong>in</strong>theQCAasasitex<strong>in</strong> the A-type quantum<br />

walk for any x ∈ Z. Moreover we observe that 2x−1(resp.2x)site<strong>in</strong>theQCA<br />

corresponds to right (resp. left) chirality at a site x <strong>in</strong> the A-type quantum<br />

walk.WecalltheQCAageneralized A-type quantum walk. WhenTAis not<br />

the zero matrix, the particle has non-zero amplitudes for ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g its<br />

position dur<strong>in</strong>g each time step. More precisely,<br />

� R ΨA,n (x)<br />

ΨA,n(x) =<br />

Ψ L A,n (x)<br />

�<br />

and<br />

ΨA,n+1(x) =QAΨA,n(x +1)+TAΨA,n(x)+PAΨA,n(x − 1).<br />

From the above observation, we see that “Type V QCA ←→ generalized Atype<br />

quantum walk”, where “X ←→ Y ” means that there is a one-to-one<br />

correspondence between X and Y ;thatis,


Quantum Walks 373<br />

Ψ R A,n(x) =βη (−1)<br />

n (2x − 1) + αη (0)<br />

n (2x − 1),<br />

Ψ L A,n(x) =βη (−1)<br />

n (2x)+αη (0)<br />

n (2x),<br />

ζ (0:−)<br />

n (2x − 1) = |Ψ R A,n(x)| 2 , ζ (0:−)<br />

n<br />

(2x) =|Ψ L A,n(x)| 2 .<br />

Here we recall Type III: |c| 2 + |d| 2 =1,cd + cd =0,cd�= 0,a= b =0. In<br />

this case, TA becomes zero matrix. So we see that Type III is noth<strong>in</strong>g but an<br />

A-type quantum walk by <strong>in</strong>terchang<strong>in</strong>g PA and QA, and the roles of left and<br />

right chiralities with c = b ′ = c ′ ,d= a ′ = d ′ . That is, “Type III QCA ←→<br />

A-type quantum walk”.<br />

We should remark that as tan φ <strong>in</strong>creases (see (4.2.6)), the relative weight<br />

of TA <strong>in</strong>creases and the particle has greater probability of ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g its<br />

position. This property also holds <strong>in</strong> a generalized B-type case <strong>in</strong>troduced <strong>in</strong><br />

the next section.<br />

4.5 Connection Between QCA and B-type<br />

Quantum Walk<br />

As <strong>in</strong> the case of the A-type quantum walk, we study a relation between the<br />

QCA and the B-type quantum walk; that is, “Type V QCA ←→ generalized<br />

B-type quantum walk”. To do this, the unitary matrix of the QCA<br />

U =<br />

is rewritten as<br />

... −2 −1 0 +1 +2 +3 +4 +5 ...<br />

⎛<br />

. . ⎞<br />

.<br />

..<br />

· · · · · · · · ...<br />

⎜<br />

⎟<br />

−2 ⎜ ... b c d 0 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−1 ⎜ ... c b a 0 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

0 ⎜ ... 0 a b c d 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

+1 ⎜ ... 0 d c b a 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

+2 ⎜ ... 0 0 0 a b c d 0 ... ⎟<br />

⎜<br />

⎟<br />

+3 ⎜ ... 0 0 0 d c b a 0 ... ⎟<br />

⎜<br />

⎟<br />

+4 ⎜ ... 0 0 0 0 0 a b c ... ⎟<br />

⎜<br />

⎟<br />

+5 ⎝ ... 0 0 0 0 0 d c b ... ⎠<br />

.<br />

.<br />

. ... · · · · · · · · ..<br />

U =<br />

... −1 0 +1 +2 ...<br />

⎛ . ⎞<br />

. .. . . . . .<br />

−1<br />

⎜ ... TB PB O O ...<br />

⎟<br />

0<br />

⎜ ... QB TB PB O ...<br />

⎟<br />

+1<br />

⎜ ... O QB TB PB ...<br />

⎟,<br />

⎟<br />

+2<br />

⎜<br />

⎝ ... O O QB TB ...<br />

⎟<br />

⎠<br />

.<br />

. ... . . . . ..


374 N. Konno<br />

where<br />

PB =<br />

� �<br />

d 0<br />

, TB =<br />

a 0<br />

� �<br />

bc<br />

, QB =<br />

cb<br />

� �<br />

0 a<br />

.<br />

0 d<br />

We consider a pair (2x, 2x+1) <strong>in</strong> the QCA as a site x <strong>in</strong> the B-type quantum<br />

walk for any x ∈ Z. Moreover we observe that 2x (resp. 2x +1) site <strong>in</strong> the<br />

QCA corresponds to left (resp. right) chirality at site x <strong>in</strong> the B-type quantum<br />

walk.WecalltheQCAageneralized B-type quantum walk. WhenTB is not<br />

the zero matrix, the particle has non-zero amplitudes for ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g its<br />

position dur<strong>in</strong>g each time step. As <strong>in</strong> the case of the A-type quantum walk,<br />

it is shown that “Type V QCA ←→ generalized B-type quantum walk”, that<br />

is,<br />

Ψ L B,n (x) =αη(0) x (2x)+βη(1) n (2x), ΨR B,n (x) =αη(0) n<br />

ζ (0:+)<br />

n (2x) =|Ψ L B,n(x)| 2 , ζ (0:+)<br />

n<br />

(2x +1)=|Ψ R B,n(x)| 2 .<br />

(2x +1)+βη(1) n (2x+1),<br />

We th<strong>in</strong>k of Type IV: |a| 2 + |d| 2 =1,ad + ad =0,ad�= 0,b= c =0. In this<br />

case, TB is zero matrix. So Type IV becomes a B-type quantum walk with<br />

d = a ′ = d ′ ,a= b ′ = c ′ ;thatis,“TypeIVQCA←→ B-type quantum walk”.<br />

Meyer (1996, 1997, 1998) has <strong>in</strong>vestigated the B-type quantum walks,<br />

which was called a quantum lattice gas automaton. His case (for example, (24)<br />

<strong>in</strong> his paper (1996)) can be obta<strong>in</strong>ed by δ → 3π/2, φ→ ρ, and θ → π/2+θ<br />

<strong>in</strong> (4.2.6).<br />

4.6 Connection Between Type V QCA and Two-step<br />

Quantum Walk<br />

In the previous two sections, we have clarified the follow<strong>in</strong>g relations: “Type<br />

III QCA ←→ A-type quantum walk”, “Type IV QCA ←→ B-type quantum<br />

walk”, moreover “Type V QCA ←→ generalized A-type quantum walk ←→<br />

generalized B-type quantum walk”. This section gives a relation between<br />

Type V QCA and two-step quantum walk. The mean<strong>in</strong>g of the “two-step”<br />

is that we identify the one-step transition of Type V QCA with the two-step<br />

transition of the quantum walk.<br />

First “Type V QCA ←→ two-step A-type quantum walk” is given. Next<br />

“Type V QCA ←→ two-step B-type quantum walk” is also presented. Comb<strong>in</strong><strong>in</strong>g<br />

them all, we f<strong>in</strong>ally obta<strong>in</strong> the follow<strong>in</strong>g relations:<br />

“Type V QCA ←→ generalized A-type quantum walk ←→ two-step A-type<br />

quantum walk”<br />

“Type V QCA ←→ generalized B-type quantum walk ←→ two-step B-type<br />

quantum walk”


Quantum Walks 375<br />

Now we present “generalized A-type quantum walk ←→ two-step A-type<br />

quantum walk” <strong>in</strong> the follow<strong>in</strong>g way. A direct computation implies that a<br />

generalized A-type quantum walk with<br />

PA =<br />

� dc<br />

00<br />

�<br />

, TA =<br />

� ba<br />

ab<br />

�<br />

, QA =<br />

� �<br />

00<br />

cd<br />

is equivalent to a two-step A-type quantum walk with<br />

� �<br />

iθ2 iθ2<br />

i cos φe s<strong>in</strong> φe<br />

PA(1) =<br />

, PA(2) = e<br />

0 0<br />

iδ<br />

�<br />

−iθ2 −iθ1<br />

s<strong>in</strong> θe −i cos θe<br />

0 0<br />

and<br />

�<br />

QA(1) =<br />

0 0<br />

s<strong>in</strong> φe iθ1 i cos φe iθ1<br />

for any θ1,θ2 ∈ [0, 2π) such that<br />

�<br />

, QA(2) = e iδ<br />

�<br />

0 0<br />

−i cos θe −iθ2 s<strong>in</strong> θe −iθ1<br />

PA = PA(2)PA(1), QA = QA(2)QA(1), TA = PA(2)QA(1) + QA(2)PA(1).<br />

Note that (a, b, c, d) has the form given <strong>in</strong> (4.2.6) and U(n) ≡ PA(n)+QA(n)<br />

is unitary for n =1, 2.<br />

In a similar fashion, we show that “generalized B-type quantum walk ←→<br />

two-step B-type quantum walk”; that is, a generalized B-type quantum walk<br />

with<br />

PB =<br />

� �<br />

d 0<br />

, TB =<br />

a 0<br />

� �<br />

bc<br />

, QB =<br />

cb<br />

corresponds to a two-step B-type quantum walk with<br />

and<br />

PB(1) =<br />

� i cos φe iθ2 0<br />

s<strong>in</strong> φe iθ1 0<br />

�<br />

iθ2 0 s<strong>in</strong>φe<br />

QB(1) =<br />

0 i cos φeiθ1 for any θ1,θ2 ∈ [0, 2π) such that<br />

�<br />

, PB(2) = e iδ<br />

�<br />

, QB(2) = e iδ<br />

� �<br />

0 a<br />

0 d<br />

� s<strong>in</strong> θe −iθ2 0<br />

−i cos θe −iθ2 0<br />

� 0 −i cos θe −iθ1<br />

0 s<strong>in</strong>θe −iθ1<br />

PB = PB(2)PB(1), QB = QB(2)QB(1), TB = PB(2)QB(1) + QB(2)PB(1).<br />

Remark that PB(n)+QB(n) =PA(n)+QA(n) forn =1, 2.<br />

F<strong>in</strong>ally we discuss the case given by Patel, Raghunathan, and Rungta<br />

(2005a). In their notation, we take general 2 × 2blocksofUeand Uo as<br />

� �<br />

� �<br />

cos φ1 i s<strong>in</strong> φ1<br />

cos φ2 i s<strong>in</strong> φ2<br />

Ue =<br />

, Uo =<br />

.<br />

i s<strong>in</strong> φ1 cos φ1<br />

i s<strong>in</strong> φ2 cos φ2<br />

�<br />

,<br />

�<br />

,<br />

�<br />

,<br />

�<br />

,


376 N. Konno<br />

Their special case is φ1 = φ2 = π/4, i.e.,<br />

Ue = Uo = 1 √ 2<br />

By us<strong>in</strong>g Ue and Uo, the follow<strong>in</strong>g matrices are def<strong>in</strong>ed:<br />

U e =<br />

and<br />

U o =<br />

� �<br />

1 i<br />

. (4.6.8)<br />

i 1<br />

... −2 −1 0 +1 +2 +3 ...<br />

. ⎛<br />

. .<br />

⎞<br />

.<br />

..<br />

⎜<br />

· · · · · · ...<br />

⎟<br />

−2 ⎜<br />

... cos φ1 i s<strong>in</strong> φ1 0 0 0 0 ... ⎟<br />

−1 ⎜<br />

... is<strong>in</strong> φ1 cos φ1 0 0 0 0 ... ⎟<br />

0 ⎜<br />

... 0 0 cosφ1is<strong>in</strong> φ1 0 0 ... ⎟<br />

+1 ⎜<br />

... 0 0 i s<strong>in</strong> φ1 cos φ1 0 0 ... ⎟,<br />

⎟<br />

+2 ⎜ ... 0 0 0 0 cosφ1is<strong>in</strong> φ1 ... ⎟<br />

+3 ⎜<br />

⎝ ... 0 0 0 0 i s<strong>in</strong> φ1 cos φ1 ... ⎟<br />

⎠<br />

.<br />

.<br />

. ... · · · · · ·<br />

..<br />

... −2 −1 0 +1 +2 +3 ...<br />

⎛ . ⎞<br />

. .. · · · · · · ...<br />

−2<br />

⎜ ... cos φ2 0 0 0 0 0 ...<br />

⎟<br />

−1<br />

⎜ ... 0 cosφ2is<strong>in</strong> φ2 0 0 0 ...<br />

⎟<br />

0<br />

⎜ ... 0 i s<strong>in</strong> φ2 cos φ2 0 0 0 ...<br />

⎟<br />

+1<br />

⎜ ... 0 0 0 cosφ2is<strong>in</strong> φ2 0 ...<br />

⎟.<br />

⎟<br />

+2<br />

⎜ ... 0 0 0 i s<strong>in</strong> φ2 cos φ2 0 ...<br />

⎟<br />

+3<br />

⎜<br />

⎝ ... 0 0 0 0 0 cosφ2 ...<br />

⎟<br />

⎠<br />

.<br />

. ... · · · · · · ..<br />

Not<strong>in</strong>g that U = U eU o, wehave<br />

(a, b, c, d) =(icos φ1 s<strong>in</strong> φ2, cos φ1 cos φ2,is<strong>in</strong> φ1 cos φ2, − s<strong>in</strong> φ1 s<strong>in</strong> φ2).<br />

(4.6.9)<br />

Therefore, by choos<strong>in</strong>g θ = φ1,φ= π/2 − φ2, and δ = π/2 <strong>in</strong> (4.2.6), we have<br />

(4.6.9).<br />

Furthermore, we see that if θ + φ = π/2, 3π/2, θ1 = θ2, and δ =2θ1 +<br />

π/2, then U(1) = U(2). To get the case of Patel, Raghunathan, and Rungta<br />

(2005a), we take θ = φ = π/4, θ1 = θ2 =0, and δ = π/2, so<br />

U(1) = U(2) = 1<br />

√ 2<br />

� i 1<br />

1 i<br />

�<br />

. (4.6.10)<br />

Remark that Ue = Uo is not equal to U(1) = U(2) <strong>in</strong> their case (see (4.6.8)<br />

and (4.6.10)). To obta<strong>in</strong> their asymptotic result, we def<strong>in</strong>e their walk at time n<br />

with the <strong>in</strong>itial qubit state ϕ = T [1/ √ 2, 1/ √ 2] by Xn.Notethatifϕ = T [α, β]


Quantum Walks 377<br />

satisfies αβ = αβ, then the distribution is symmetric at any time, (see<br />

Theorem 1.5). Then, Theorem 1.6 implies that<br />

P (a ≤ X ϕ � b<br />

n /n ≤ b) →<br />

a<br />

4<br />

π(4 − x2 ) √ dx (n →∞),<br />

4 − 2x2 for − √ 2 ≤ a


378 N. Konno<br />

fluctuation of quantum walk and show that the temporal standard deviation<br />

correspond<strong>in</strong>g to the Hadamard walk depends on the location of sites even if<br />

time-averaged density function is uniform. This result has a strik<strong>in</strong>g difference<br />

between quantum and classical cases, s<strong>in</strong>ce temporal standard deviation of<br />

the classical case is zero for any site. Furthermore we consider the dependence<br />

of the standard deviation on the <strong>in</strong>itial state. Results <strong>in</strong> this chapter are based<br />

on Inui, Konishi, Konno, and Soshi (2005).<br />

5.2 Def<strong>in</strong>ition<br />

The Hadamard walk on a cycle with N sites is given by the follow<strong>in</strong>g unitary<br />

matrix with 2N × 2N elements,<br />

⎡<br />

⎤<br />

0 P 0 ··· ··· 0 Q<br />

⎢ Q 0 P 0 ··· ··· 0 ⎥<br />

⎢ 0 Q 0 P 0 ··· 0 ⎥<br />

⎢<br />

U N = ⎢<br />

.<br />

⎢ . .. . .. . .. . .. . ..<br />

⎥<br />

. ⎥ ,<br />

⎥<br />

⎢ 0 ··· 0 Q 0 P 0 ⎥<br />

⎣ 0 ··· ··· 0 Q 0 P ⎦<br />

(5.2.1)<br />

P 0 ··· ··· 0 Q 0<br />

where<br />

P = 1 √ 2<br />

� �<br />

11<br />

, Q =<br />

00<br />

1<br />

� �<br />

0 0<br />

√ , 0=<br />

2 1 −1<br />

� �<br />

00<br />

. (5.2.2)<br />

00<br />

Thus the total state after n step, Ψn, isgivenbyU n<br />

N Ψ0 for the <strong>in</strong>itial state<br />

Ψ0.<br />

To express the total state as a function of position x and time n, weuse<br />

known results on eigenvalues (Lemma 5.1) and eigenvectors (Lemma 5.2) of<br />

U N obta<strong>in</strong>ed by Aharonov et al. (2001) and Bednarska et al. (2003).<br />

Lemma 5.1. For j =0, 1,...,N − 1 and k =0, 1, the(2j + k +1)-theigen value correspond<strong>in</strong>g to U N is given by<br />

cjk = 1<br />

�<br />

√ (−1)<br />

2<br />

k<br />

�<br />

1+cos2 � � � �<br />

2 jπ 2 jπ<br />

+ i s<strong>in</strong><br />

N<br />

N<br />

�<br />

. (5.2.3)<br />

We here def<strong>in</strong>e v o j,k,l and ve j,k,l<br />

to express eigenvectors by<br />

v o j,k,l = ajk bjk ω jl<br />

N , (5.2.4)<br />

v e j,k,l = ajk ω jl<br />

N , (5.2.5)


Quantum Walks 379<br />

where<br />

2 iπ<br />

ωN = e N , (5.2.6)<br />

1<br />

ajk = � � , (5.2.7)<br />

N 1+|bjk|<br />

2�<br />

bjk = ω j<br />

�<br />

N (−1) k<br />

�<br />

1+cos2 �<br />

ξj +cosξj , (5.2.8)<br />

ξj = 2jπ<br />

. (5.2.9)<br />

N<br />

Lemma 5.2. For l =1, 2,...,2N, thel-thelement of the eigenvectors correspond<strong>in</strong>g<br />

to cjk is given by<br />

⎧<br />

⎨<br />

vj,k,l =<br />

⎩<br />

v o j,k,(l+1)/2<br />

v e j,k,l/2<br />

(l = odd),<br />

(l = even).<br />

(5.2.10)<br />

For the convenience of readers, from now on here we check directly that<br />

eigenvectors vj,k = T [vj,k,1, ··· ,vj,k,2N ] satisfy the next equation:<br />

U N vj,k = cjkvj,k. (5.2.11)<br />

(a) Suppose l is odd, then the (2m − 1)-th element of the left-hand side of<br />

the (5.2.11) for m =1, 2, ··· ,N is given by<br />

(2m − 1)-th element of LHS<br />

= 1 �<br />

√ v<br />

2<br />

o<br />

�<br />

j,k,m+1 mod<br />

+ ve<br />

N j,k,m+1 mod N<br />

= ajk<br />

�<br />

√2 e 2j(m+1)iπ<br />

�<br />

N 1+e 2jiπ<br />

�<br />

N (−1) k<br />

�<br />

1+cos2 ���<br />

ξj +cosξj , (5.2.12)<br />

where ξj =2πj/N. On the other hand, the right-hand side of (5.2.11) is given<br />

by<br />

(2m − 1)-th element of RHS<br />

= cjkajkbjkω jm<br />

N<br />

= ajk<br />

�<br />

√2 e 2j(m+1)iπ<br />

N<br />

�<br />

(−1) k<br />

�<br />

×<br />

1+cos 2 ξj +cosξj<br />

�<br />

(−1) k<br />

�<br />

1+cos2 ��<br />

ξj + i s<strong>in</strong> ξj . (5.2.13)<br />

(b) Similarly suppose l is even, then the 2m-th element of the left-hand<br />

side of the (5.2.11) for m =1, 2, ··· ,N is given by<br />


380 N. Konno<br />

2m-th element of LHS<br />

= 1 �<br />

√ v<br />

2<br />

o<br />

j,k,m−1 mod N<br />

= ajk<br />

√2<br />

�<br />

e 2j(m−1)iπ<br />

N<br />

− ve<br />

j,k,m−1 mod N<br />

�<br />

−1+e 2jiπ<br />

N<br />

�<br />

(−1) k<br />

�<br />

�<br />

1+cos 2 ξj +cosξj<br />

On the other hand, the right-hand side of the (5.2.11) is given by<br />

2m-th element of RHS<br />

= cjkajkω jm<br />

N<br />

= ajk<br />

�<br />

√2 e 2jmiπ<br />

N<br />

���<br />

. (5.2.14)<br />

�<br />

(−1) k<br />

�<br />

1+cos2 ��<br />

ξj + i s<strong>in</strong> ξj . (5.2.15)<br />

Us<strong>in</strong>g the Euler’s formula, we can check that both sides of (5.2.11) co<strong>in</strong>cide<br />

for each case.<br />

5.3 Temporal Standard Deviation<br />

S<strong>in</strong>ce we obta<strong>in</strong>ed the eigenvalues and eigenvectors, matrix U N is transformed<br />

<strong>in</strong>to a diagonal matrix and the wave function after n steps is generally expressed<br />

by a l<strong>in</strong>er comb<strong>in</strong>ation of c n jk .Inrealsystemwecannotobservethe<br />

wave function but probability of be<strong>in</strong>g at position x at time n, Pn,N (x). In<br />

contrast to classical random walks, the probability Pn,N (x) dose not converge<br />

<strong>in</strong> the limit of n →∞. Thus we consider a time-averaged distribution<br />

def<strong>in</strong>ed by<br />

T −1<br />

1<br />

¯PN (x) = lim<br />

T →∞ T<br />

n=0<br />

�<br />

Pn,N (x), (5.3.16)<br />

if the right-hand side of (5.3.16) exists. Aharonov et al. (2001) showed that<br />

¯PN (x) exists, and it is <strong>in</strong>dependent of both an <strong>in</strong>itial state and x if all eigenvalues<br />

of U N are dist<strong>in</strong>ct. If the number of site is odd, then all eigenvalues of U N<br />

are dist<strong>in</strong>ct, so we immediately have ¯ PN (x) =1/N for any x =0, 1,...,N−1.<br />

On the other hand, if N is even, then there exist degenerate eigenvalues of<br />

U N and it is possible to derive non-uniform distributions. In the case of the<br />

classical random walk, the time-averaged distribution of a walker becomes<br />

uniform <strong>in</strong>dependently on the parity of the system size. Thus we can not dist<strong>in</strong>guish<br />

the classical random walk and the Hadamard walk on a cycle with<br />

odd sites only by the time-averaged distribution.<br />

As we mentioned above, the Pn,N (x) does not converge <strong>in</strong> the limit n →∞.<br />

It means that the probability always fluctuates to the upper and lower sides of<br />

PN (x). For this reason, we def<strong>in</strong>e the follow<strong>in</strong>g temporal standard deviation<br />

σN (x):


Quantum Walks 381<br />

�<br />

�<br />

� T<br />

σN (x) = � 1 �−1<br />

�<br />

lim Pn,N (x) −<br />

T →∞ T<br />

¯ PN (x) �2 , (5.3.17)<br />

n=0<br />

if the right-hand side of (5.3.17) exists.<br />

Here we consider the classical case. In the case of classical a random walk<br />

start<strong>in</strong>g from a site for an odd number of sites (i.e., aperiodic case), there<br />

exist a ∈ (0, 1) and C>0 (are <strong>in</strong>dependent of x and n) such that<br />

|Pn,N (x) − ¯ PN (x)| ≤Ca n ,<br />

where ¯ PN (x) =1/N for any x =0, 1,...,N − 1, (see page 63 of Sch<strong>in</strong>azi<br />

(1999), for example). Therefore we obta<strong>in</strong><br />

1<br />

T<br />

T� −1<br />

�<br />

Pn,N (x) − ¯ PN (x) �2 C<br />

≤ 2<br />

n=0<br />

T<br />

1 − a2T .<br />

1 − a2 The above <strong>in</strong>equality implies that for any x =0, 1,...,N − 1,<br />

σN (x) =0.<br />

<strong>in</strong> the classical case. As for N = even (i.e., periodic) case, we have the same<br />

conclusion σN (x) = 0 for any x by us<strong>in</strong>g a little modified argument.<br />

In this situation, we first ask the natural question whether the σN (x) is<br />

always zero or not <strong>in</strong> the quantum case. Moreover if σN (x) is not always zero,<br />

then we next ask the question whether σN (x) is uniform or not.<br />

From now on we concentrate our attention to the system with odd sites.<br />

Furthermore we assume that the <strong>in</strong>itial state is a fixed Ψ0 = T [1, 0, 0, ··· , 0].<br />

Let us express the standard deviation σN (x) start<strong>in</strong>gΨ0 = T [1, 0, 0, ··· , 0]<br />

by wave functions. To do so, we have<br />

σ 2 1<br />

N (x) = lim<br />

T →∞ T<br />

T� −1<br />

n=0<br />

�<br />

|Ψ L n (x)|2 + |Ψ R n (x)|2 − 1<br />

�2 , (5.3.18)<br />

N<br />

whereweusedtheresult ¯ P (x) =1/N for odd N. S<strong>in</strong>ce the matrix U N was<br />

already diagonalized <strong>in</strong> the previous section, the wave functions Ψ L n (x) and<br />

Ψ R n (x) are easily obta<strong>in</strong>ed as<br />

Ψ L n (x) =<br />

Ψ R n (x) =<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=0<br />

�<br />

αj0c n j0 + αj1cn �<br />

j1 , (5.3.19)<br />

�<br />

βj0c n j0 + βj1c n �<br />

j1 , (5.3.20)


382 N. Konno<br />

where<br />

αjk =<br />

2njπi<br />

e N<br />

2N<br />

⎛<br />

⎝1+(−1) k<br />

βjk = (−1)ke 2(n−1)jπi<br />

N<br />

�<br />

2N<br />

1+cos 2 ( 2jπ<br />

N )<br />

cos( 2jπ<br />

N )<br />

�<br />

1+cos 2 ( 2jπ<br />

N )<br />

⎞<br />

⎠ , (5.3.21)<br />

. (5.3.22)<br />

S<strong>in</strong>ce the matrix U N is unitary matrix, the eigenvalue cjk is written as e iθjk<br />

where θjk is the argument of cjk. Thus the probability |Ψ L n (x)| 2 is expressed<br />

by<br />

where<br />

�<br />

|Ψ L n (x)| 2 N−1<br />

= (|αj0| 2 + |αj1| 2 )<br />

j=0<br />

+<br />

N−1 �<br />

1�<br />

j0,j1=0 k0,k1=0<br />

δj0j1,k0k1 =<br />

δj0j1,k0k1αj0k0α ∗ j1k1 ei(θj 0 k 0 −θj 1 k 1 )n , (5.3.23)<br />

� 0 j0 = j1 and k0 = k1,<br />

1 otherwise.<br />

(5.3.24)<br />

The first term of (5.3.23) is constant and the second term vanishes by<br />

apply<strong>in</strong>g an operation limT →∞ 1 �T −1<br />

T n=0 . Similarly we can divide |Ψ R n (x)|2<br />

<strong>in</strong>to a constant term and a vanish<strong>in</strong>g term:<br />

�<br />

|Ψ R n (x)|2 N−1<br />

= (|βj0| 2 + |βj1| 2 )<br />

j=0<br />

+<br />

N−1 �<br />

As a result we have<br />

N−1 �<br />

¯PN (x) =<br />

1�<br />

j0,j1=0 k0,k1=0<br />

j=0<br />

δj0j1,k0k1βj0k0β ∗ j1k1 ei(θj 0 k 0 −θj 1 k 1 )n .(5.3.25)<br />

(|αj0| 2 + |αj1| 2 + |βj0| 2 + |βj1| 2 ). (5.3.26)<br />

Plugg<strong>in</strong>g (5.3.21)-(5.3.22) <strong>in</strong>to (5.3.26) gives confirm ¯ PN (x) =1/N for any<br />

x =0, 1,...,N − 1.<br />

Let express σ2 N (x) as a function of eigenvalues. By us<strong>in</strong>g (5.3.18), (5.3.23)<br />

and (5.3.25), we get


Quantum Walks 383<br />

where<br />

σ 2 T<br />

1 �−1<br />

N (x) = lim<br />

T →∞ T<br />

N−1 �<br />

1�<br />

n=0 j0,j1,j2,j3=0 k0,k1,k2,k3=0<br />

δj0j1,k0k1δj2j3,k2k3<br />

×(αj0k0α ∗ j1k1 + βj0k0β ∗ j1k1 )(αj2k2α ∗ j3k3 + βj2k2β ∗ j3k3 )ei∆θn , (5.3.27)<br />

∆θ ≡ ∆θ(j0,k0,j1,k1,j2,k2,j3,k3)<br />

= θj0k0 − θj1k1 + θj2k2 − θj3k3. (5.3.28)<br />

�T −1<br />

If ∆θ �= 0(mod2π), then limT →∞ n=0 ei∆θn /T converges to 0. Thus<br />

(x) is obta<strong>in</strong>ed by tak<strong>in</strong>g the sum of<br />

the variance σ 2 N<br />

(αj0k0α ∗ j1k1 + βj0k0β ∗ j1k1 )(αj2k2α ∗ j3k3 + βj2k2β ∗ j3k3 )<br />

over comb<strong>in</strong>ations j0,k0,...,j3,k3 satisfy<strong>in</strong>g ∆θ = 0 (mod 2π). For this<br />

reason, we consider the comb<strong>in</strong>ations which satisfy ∆θ =0(mod2π) fora<br />

given comb<strong>in</strong>ation (j0,k0,j1,k1). We def<strong>in</strong>e the follow<strong>in</strong>g four conditions:<br />

(a) ℜ(cj0k0) =ℜ(cj1k1) and ℑ(cj0k0) =−ℑ(cj1k1),<br />

(b) ℜ(cj0k0) =−ℜ(cj1k1) and ℑ(cj0k0) =ℑ(cj1k1),<br />

(c) ℜ(cj0k0) =−ℜ(cj1k1) and ℑ(cj0k0) =−ℑ(cj1k1),<br />

(d) ℜ(cj0k0) =ℜ(cj1k1) and ℑ(cj0k0) =ℑ(cj1k1), (5.3.29)<br />

where ℜ(z) (resp.ℑ(z)) is the real (resp. imag<strong>in</strong>ary) part of z ∈ C. Ifeach<br />

condition from (a) to(d) is not satisfied, there are four different comb<strong>in</strong>ations<br />

which satisfy ∆θ =0(mod2π) for a given comb<strong>in</strong>ation (j0,k0,j1,k1):<br />

(a) j2 = N − j0 mod N, k2 = k0, j3 = N − j1 mod N, k3 = k1,<br />

(b) j2 = j0, k2 =1− k0, j3 = j1, k3 =1− k1,<br />

(c) j2 = N − j1 mod N, k2 =1− k1, j3 = N − j0 mod N, k3 =1− k0,<br />

(d) j2 = j1, k2 = k1, j3 = j0, k3 = k0. (5.3.30)<br />

If one of conditions <strong>in</strong> (5.3.29) is satisfied, we f<strong>in</strong>d the same values <strong>in</strong> (5.3.30).<br />

We can <strong>in</strong>troduce a geometrical picture <strong>in</strong>to these comb<strong>in</strong>ations. All eigenvalues<br />

lie on the unit cycle <strong>in</strong> the complex plane and there is a symmetry <strong>in</strong><br />

these po<strong>in</strong>ts. For example, when we f<strong>in</strong>d cj0k0 at the po<strong>in</strong>t A, we can f<strong>in</strong>d<br />

always another eigenvalue on a po<strong>in</strong>t which is obta<strong>in</strong>ed by reflect<strong>in</strong>g po<strong>in</strong>t A<br />

<strong>in</strong> the x-axis and its eigenvalue is expressed by c N−j0 mod N,k0 . We also f<strong>in</strong>d<br />

always an eigenvalue on a po<strong>in</strong>t which is obta<strong>in</strong>ed by reflect<strong>in</strong>g po<strong>in</strong>t A <strong>in</strong><br />

the y-axis and its eigenvalues is expressed by cj0,1−k0. Additionally, we note<br />

that if the relation between cj0k0 and cj1k1 is symmetrical with respect to<br />

the orig<strong>in</strong> and the relation between cj2k2 and cj3k3 is also symmetrical with<br />

respect to the orig<strong>in</strong>, then ∆θ =0(mod2π).


384 N. Konno<br />

From the relation (5.3.30) we express j2,k2,j3 and k3 as a function of j0,<br />

k0, j1 and k1. Thusσ2 N (x) is obta<strong>in</strong>ed by summation over j0, k0, j1 and<br />

k1. Tak<strong>in</strong>g the degeneracy <strong>in</strong>to account correctly we obta<strong>in</strong> σ2 N (x) bysome<br />

computations. Before show<strong>in</strong>g the result we def<strong>in</strong>e next functions<br />

S0 =<br />

S1 =<br />

S+(x) =<br />

S−(x) =<br />

S2(x) =<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=1<br />

1<br />

, (5.3.31)<br />

3+cosθj<br />

cos θj<br />

, (5.3.32)<br />

3+cosθj<br />

cos ((x − 1)θj)+cos(xθj)<br />

, (5.3.33)<br />

3+cosθj<br />

cos ((x − 1)θj) − cos (xθj)<br />

,<br />

3+cosθj<br />

(5.3.34)<br />

7+cos2θj +8cosθjcos2 �� x − 1<br />

� �<br />

2<br />

θj<br />

(3 + cos θj) 2 , (5.3.35)<br />

where θj ≡ 4πj/N. Now we show σ2 N (x) as ma<strong>in</strong> result <strong>in</strong> this chapter:<br />

Theorem 5.3. When N is odd, we have<br />

σ 2 1<br />

N (x) =<br />

N 4<br />

� 2 � S 2 + (x)+S 2 − (x)� +11S 2 0 +10S0S1 +3S 2 1 − S2(x) � − 2<br />

for any x =0, 1,...,N − 1.<br />

5.4 Dependence of σN(x) on the Position<br />

and System Size<br />

,<br />

N 3<br />

(5.3.36)<br />

The formula obta<strong>in</strong>ed for σN (x) is written <strong>in</strong> the form of a s<strong>in</strong>gle summation,<br />

therefore, it is calculated very easily <strong>in</strong> comparison with (5.3.27). We consider<br />

σN (x) forN =3, 5, ··· , 11.<br />

First we f<strong>in</strong>d that σN (x) depends on the position and it has the maximum<br />

value at x = 0 and 1. The dependence of σN (n) on the position from 2 to<br />

N − 1 is not clear, however we f<strong>in</strong>d that σN (x)= σN (N +1− x) (mod N)<br />

for any x and all values are dist<strong>in</strong>ct except a pair x and N +1− x (mod N).<br />

Next we consider asymptotic behavior of σN (0) for large N to observe the<br />

dependence on the system size. Sett<strong>in</strong>g x = 0 <strong>in</strong> (5.3.36) gives


Quantum Walks 385<br />

σ 2 1<br />

N (0) =<br />

N 3<br />

⎡<br />

⎤<br />

N−1 � 7cosθj<br />

⎣<br />

− 2⎦<br />

3+cosθj<br />

j=0<br />

+ 1<br />

N 4<br />

⎡⎛<br />

⎞ ⎛<br />

⎞<br />

N−1 �<br />

N−1<br />

⎣⎝<br />

1 �<br />

⎠ ⎝<br />

15 − 11 cos θj ⎠<br />

3+cosθj 3+cosθj<br />

j=0<br />

j=0<br />

N−1 � 9+4cosθj +3cos2θj<br />

−<br />

(3 + cos θj)<br />

j=1<br />

2<br />

⎤<br />

⎦ . (5.4.37)<br />

On the other hand, we have<br />

N−1<br />

1 � 1<br />

lim<br />

N→∞ N 3+cosθj<br />

j=0<br />

lim<br />

N→∞<br />

lim<br />

N→∞<br />

lim<br />

N→∞<br />

lim<br />

N→∞<br />

N−1<br />

1 �<br />

N<br />

j=0<br />

N−1<br />

1 �<br />

N<br />

j=0<br />

N−1<br />

1 �<br />

N<br />

j=0<br />

N−1<br />

1 �<br />

N<br />

j=0<br />

cos θj<br />

3+cosθj<br />

= 1<br />

� 4π<br />

1 1<br />

dx =<br />

4π 0 3+cosx 2 √ 2 ,<br />

= 1<br />

� 4π<br />

cos x<br />

3<br />

dx =1−<br />

4π 0 3+cosx 2 √ 2 ,<br />

� 4π<br />

1 1<br />

=<br />

(3 + cos θj) 2 4π 0<br />

� 4π<br />

cos θj 1<br />

=<br />

(3 + cos θj) 2 4π 0<br />

� 4π<br />

cos(2θj) 1<br />

=<br />

(3 + cos θj) 2 4π 0<br />

1<br />

3<br />

dx =<br />

(3 + cos x) 2<br />

16 √ 2 ,<br />

cos x<br />

1<br />

dx = −<br />

(3 + cos x) 2 16 √ 2 ,<br />

cos 2x<br />

45<br />

dx =2−<br />

(3 + cos x) 2<br />

16 √ 2 .<br />

Us<strong>in</strong>g the above results, we obta<strong>in</strong> the asymptotic behavior of the variance<br />

σ2 N (0) for sufficiently large N as follows:<br />

Proposition 5.4. For N →∞, we have<br />

σ 2 N (0) = 13 − 8√ 2<br />

N 2<br />

+ 7√ 2 − 16<br />

2N 3<br />

�<br />

1<br />

+ o<br />

N 3<br />

�<br />

. (5.4.38)<br />

The above result implies that the fluctuation σN (0) decays <strong>in</strong> the form<br />

1/N as N <strong>in</strong>creases.<br />

5.5 Dependence of σN(x) on Initial States and Parity<br />

of System Size<br />

Dependence of σN (x) on Initial States<br />

In the previous sections, we have shown only results cover<strong>in</strong>g the case where<br />

the system size is odd and the <strong>in</strong>itial state is Ψ L 0 (0) = 1. The time averaged


386 N. Konno<br />

distribution on a cycle <strong>in</strong>clud<strong>in</strong>g odd sites is <strong>in</strong>dependent of the <strong>in</strong>itial state.<br />

Is also the temporal standard deviation <strong>in</strong>dependent of the <strong>in</strong>itial state? We<br />

consider here whether the temporal standard deviation depends on the <strong>in</strong>itial<br />

state even if the <strong>in</strong>itial probability distribution is the same.<br />

In order to observe the dependence of the temporal standard deviation<br />

on <strong>in</strong>itial state without chang<strong>in</strong>g the <strong>in</strong>itial probability distribution, we set<br />

Ψ L 0 (0)=α and Ψ R 0 (0)= √ 1 − α 2 i where α ∈ [0, 1]. As a particular case, we<br />

have a symmetrical <strong>in</strong>itial state by sett<strong>in</strong>g α =1/ √ 2. Thus we can observe<br />

the dependence of σN (x, α) on symmetry break down by chang<strong>in</strong>g α. Wefirst<br />

present the result for N = 3. The standard deviation σ3(x, α) fortheabove<br />

<strong>in</strong>itial state is obta<strong>in</strong>ed as follows:<br />

σ3(0,α)= 2 √ 46<br />

, (5.5.39)<br />

45<br />

σ3(1,α)= 2 �<br />

96 α4 − 75 α2 +25, (5.5.40)<br />

45<br />

σ3(2,α)= 2 �<br />

96 α4 − 117 α2 +46. (5.5.41)<br />

45<br />

One clearly f<strong>in</strong>ds that σ3(0,α) is <strong>in</strong>dependent of the <strong>in</strong>itial and both σ3(1,α)<br />

and σ3(2,α) depend on α. We further f<strong>in</strong>d that σ3(1,α)isequaltoσ3(2,α)<br />

at the symmetrical case α =1/ √ 2 and the m<strong>in</strong>imum values are located near<br />

α =1/ √ 2.<br />

Next we consider N =5andN =7.As<strong>in</strong>thecaseofN = 3, the temporal<br />

standard deviation at the position x = 0 is <strong>in</strong>dependent of the parameter α<br />

and the m<strong>in</strong>imum values are located near α =1/ √ 2. The dependence of<br />

σN (x, α) on the position is weak between 2 and N − 2, and they seem to be<br />

<strong>in</strong> close on a s<strong>in</strong>gle l<strong>in</strong>e.<br />

Temporal Standard Deviation on a Cycle with Even Sites<br />

We have concentrated our attention on the case where the number of the<br />

system size is odd. The ma<strong>in</strong> reason why we avoid even cases is that the<br />

eigenvalues of matrix U N are highly degenerate, and the time averaged distribution<br />

itself depends on the position of site and the <strong>in</strong>itial state. While<br />

the eigenvalues for even cases is exactly obta<strong>in</strong>ed by (5.2.3), there are many<br />

possible comb<strong>in</strong>ations of eigenvalues which satisfy ∆θ =0(mod2π) forthe<br />

even case. Thus we can not calculate σN (x) <strong>in</strong> the same way with odd cases.<br />

Therefore, we here try to carry out approximate calculations <strong>in</strong>stead of seek<strong>in</strong>g<br />

analytic results for N = 4, 6, ··· , 12 under the same <strong>in</strong>itial condition<br />

Ψ L 0 (0) = 1 and T =10 4 .<br />

First we f<strong>in</strong>d that the values of the temporal standard deviation for odd N<br />

are almost the same except x =0, 1, while the values for even N are strongly<br />

dependent on the location. Second we f<strong>in</strong>d that the maximum values of σN (x)


Quantum Walks 387<br />

for odd N exist only at x =0andx = 1, while the four same peaks are found<br />

for N =4, 8, 12. We confirm numerically that this behavior is always observed<br />

up to N =22whenN is a multiple of 4.<br />

5.6 Summary<br />

We expressed the temporal standard deviation of the Hadamard walk on a<br />

(0) = 1 <strong>in</strong> an analytical form.<br />

cycle with odd sites for a pure <strong>in</strong>itial state Ψ L 0<br />

The formula is obta<strong>in</strong>ed by calculat<strong>in</strong>g four Fourier coefficients which come<br />

from “L” and“R” states. Unlike the time-averaged distribution, the temporal<br />

standard deviation depends on the location of the site. The fluctuations<br />

decrease <strong>in</strong> <strong>in</strong>verse proportion to the system size for large N and its coefficient<br />

was calculated rigorously.<br />

When we set the <strong>in</strong>itial state <strong>in</strong> the form Ψ L 0 (0)=α and Ψ R 0 (0)=√1 − α2 i,<br />

the analytical result for N = 3 and numerical simulations <strong>in</strong>dicated that the<br />

temporal standard deviation depends on the <strong>in</strong>itial state except x =0.We<br />

would speculate that the temporal standard deviation takes the maximum<br />

value at x = 0 and its value is <strong>in</strong>dependent of α. The values of the temporal<br />

standard deviation at neighbor sites of the orig<strong>in</strong> is somewhat larger than<br />

other at sites and the rema<strong>in</strong>s are almost the same. The dependence of the<br />

temporal standard deviation on the position for even system size is more<br />

complex than that for odd system size due to degeneration of eigenvalues.<br />

Several peaks are observed when the system size is a multiple of 4.<br />

From the viewpo<strong>in</strong>t of technology, it seen to be of value to consider whether<br />

the temporal standard deviation is controlled or not. It is shown that the timeaveraged<br />

distribution can not be controlled for odd system size, however, for<br />

even system size, it is controlled thanks to the degeneracy of eigenvalues.<br />

Our results show that the standard deviation can be controlled. First the<br />

dependence of temporal standard deviation on location is used. Second the<br />

dependence on the <strong>in</strong>itial state is also used. However we note here that σN (0)<br />

for N = 3 is <strong>in</strong>dependent of the <strong>in</strong>itial state. In this sense, its control is<br />

limited.<br />

6 Absorption Problems<br />

6.1 Introduction<br />

In this chapter we consider absorption problems for quantum walks given<br />

by U ∈ U(2) located on the sets {0, 1,...,N} or {0, 1,...}. Results <strong>in</strong> this<br />

chapter appeared <strong>in</strong> Konno, Namiki, Soshi, and Sudbury (2003). Before we


388 N. Konno<br />

move to the quantum case, we first describe the classical random walk on a<br />

f<strong>in</strong>ite set {0, 1,...,N} with two absorb<strong>in</strong>g barriers at locations 0 and N (see<br />

Grimmett and Stirzaker (1992), Durrett (1999), for examples). The particle<br />

moves at each step either one unit to the left with probability p or one unit to<br />

the right with probability q =1−p until it hits one of the absorb<strong>in</strong>g barriers.<br />

The directions of different steps are <strong>in</strong>dependent of each other. The classical<br />

random walk start<strong>in</strong>g from m ∈{0, 1,...,N} at time n is denoted by Sm n<br />

here. Let<br />

Tℓ =m<strong>in</strong>{n≥0:S m n = ℓ}<br />

be the time of the first visit to ℓ ∈{0, 1,...,N}. Us<strong>in</strong>g the subscript m to<br />

<strong>in</strong>dicate Sm 0 = m, welet<br />

P (N,m) = P (T0


Quantum Walks 389<br />

Step 1. Initialize the system ϕ ∈ Φ at location m.<br />

Step 2. (a) Perform one step time evolution. (b) Measure the system to<br />

see where it is or is not at location 0.<br />

Step 3. If the result of measurement revealed that the system was at location<br />

0, then term<strong>in</strong>ate the process, otherwise repeat step 2.<br />

In this sett<strong>in</strong>g, let Ξ (∞,m)<br />

n be the sum over possible paths for which the<br />

particle first hits 0 at time n start<strong>in</strong>g from m. For example,<br />

Ξ (∞,1)<br />

5 = P 2 QP Q + P 3 Q 2 =(ab 2 c + a 2 bd)R. (6.2.3)<br />

The probability that the particle first hits 0 at time n start<strong>in</strong>g from m is<br />

P (∞,m)<br />

n (ϕ) =||Ξ (∞,m)<br />

n ϕ|| 2 . (6.2.4)<br />

So the probability that the particle hits 0 start<strong>in</strong>g from m is<br />

P (∞,m) (ϕ) =<br />

∞�<br />

n=0<br />

P (∞,m)<br />

n (ϕ).<br />

Next we consider N


390 N. Konno<br />

before it arrives at N. We may use conditional probabilities to see that P (N,m)<br />

satisfies the follow<strong>in</strong>g difference equation:<br />

P (N,m) = pP (N,m−1) + qP (N,m+1)<br />

with boundary conditions:<br />

(1 ≤ m ≤ N − 1), (6.3.5)<br />

P (N,0) =1, P (N,N) =0. (6.3.6)<br />

The solution of such a difference equation is given by<br />

P (N,m) =1− m<br />

N<br />

if p =1/2, (6.3.7)<br />

P (N,m) = (p/q)m − (p/q) N<br />

1 − (p/q) N if p �= 1/2, (6.3.8)<br />

for any 0 ≤ m ≤ N. Therefore, when N = ∞, weseethat<br />

Furthermore,<br />

P (∞,m) =1 if 1/2 ≤ p ≤ 1, (6.3.9)<br />

P (∞,m) =(p/q) m<br />

if 0 ≤ p


Quantum Walks 391<br />

for any <strong>in</strong>itial qubit state ϕ = T [α, β] ∈ Φ. Furthermore, <strong>in</strong> the case of<br />

U = � H(ρ), Bach et al. (2002) gave<br />

lim<br />

m→∞ P (∞,m) ( T [0, 1]) = ρ<br />

� �<br />

−1 cos (1 − 2ρ)<br />

2<br />

− 1 +<br />

1 − ρ π<br />

π � 1/ρ − 1 ,<br />

lim<br />

m→∞ P (∞,m) ( T [1, 0]) = cos−1 (1 − 2ρ)<br />

.<br />

π<br />

The second result was conjectured by Yamasaki, Kobayashi, and Imai (2002).<br />

When N is f<strong>in</strong>ite, the follow<strong>in</strong>g conjecture by Amba<strong>in</strong>is et al. (2001) is still<br />

open for the U = H case:<br />

P (N+1,1) ( T [0, 1]) = 2P (N,1) ( T [0, 1]) + 1<br />

2P (N,1) ( T [0, 1]) + 2<br />

Solv<strong>in</strong>g the above recurrence gives<br />

P (N,1) ( T [0, 1]) = 1<br />

√ 2 × (3 + 2√ 2) N−1 − 1<br />

(3 + 2 √ 2) N−1 +1<br />

(N ≥ 1), P (1,1) ( T [0, 1]) = 0.<br />

(N ≥ 1). (6.3.14)<br />

However <strong>in</strong> contrast with P (∞,1) ( T [0, 1]) = 2/π, Amba<strong>in</strong>is et al. (2001) proved<br />

lim<br />

N→∞ P (N,1) ( T [0, 1]) = 1/ √ 2.<br />

It should be noted that P (∞,m) = limN→∞ P (N,m) for any 0 ≤ m ≤ N <strong>in</strong> the<br />

classical case (see (6.3.7) - (6.3.10)).<br />

We are not aware of results concern<strong>in</strong>g E (∞,1) (T0|T0 < ∞) hav<strong>in</strong>g been<br />

published, however we will give such a result <strong>in</strong> the next section.<br />

6.4 Results<br />

In the first half of this section, we consider the general sett<strong>in</strong>g <strong>in</strong>clud<strong>in</strong>g a<br />

hitt<strong>in</strong>g time to N before it arrives at 0 or a hitt<strong>in</strong>g time to 0 before it arrives<br />

at N, whenN is f<strong>in</strong>ite. In the case of N = ∞, a similar argument holds, so<br />

we will omit it here.<br />

Not<strong>in</strong>g that {P, Q, R, S} is a basis of M2(C), Ξ (N,m)<br />

n can be written as<br />

Ξ (N,m)<br />

n<br />

= p (N,m)<br />

n<br />

Therefore (6.2.4) implies<br />

P (N,m)<br />

n<br />

(ϕ) =||Ξ (N,m)<br />

n<br />

P + q (N,m)<br />

Q + r (N,m)<br />

R + s (N,m)<br />

S.<br />

n<br />

n<br />

ϕ|| 2 = C1(n)|α| 2 + C2(n)|β| 2 +2ℜ(C3(n)αβ),<br />

n


392 N. Konno<br />

where ℜ(z) is the real part of z ∈ C, ϕ = T [α, β] ∈ Φ and<br />

C1(n) =|ap (N,m)<br />

n<br />

C2(n) =|bp (N,m)<br />

n<br />

C3(n) =(ap (N,m)<br />

n<br />

+ cr (N,m)<br />

n | 2 + |as (N,m)<br />

n<br />

+ dr (N,m)<br />

n<br />

| 2 + |bs (N,m)<br />

n<br />

+ cr (N,m)<br />

n )(bp (N,m)<br />

n<br />

+(as (N,m)<br />

n<br />

+ cq (N,m)<br />

n | 2 ,<br />

+ dq (N,m)<br />

n<br />

+ dr (N,m)<br />

n )<br />

+ cq (N,m)<br />

n )(bs (N,m)<br />

n<br />

| 2 ,<br />

+ dq (N,m)<br />

n ).<br />

From now on we assume N ≥ 3. Not<strong>in</strong>g the def<strong>in</strong>ition of Ξ (N,m)<br />

n ,wesee<br />

that for 1 ≤ m ≤ N − 1,<br />

Ξ (N,m)<br />

n<br />

= Ξ (N,m−1)<br />

n−1<br />

P + Ξ (N,m+1)<br />

Q.<br />

The above equation is a quantum version of the difference equation, i.e.,<br />

(6.3.5) for the classical random walk. Then we have<br />

p (N,m)<br />

n<br />

q (N,m)<br />

n<br />

r (N,m)<br />

n<br />

s (N,m)<br />

n<br />

= ap (N,m−1)<br />

n−1<br />

= dq (N,m+1)<br />

n−1<br />

= bp (N,m+1)<br />

n−1<br />

= cq (N,m−1)<br />

n−1<br />

n−1<br />

+ cr (N,m−1)<br />

n−1 ,<br />

+ bs (N,m+1)<br />

n−1 ,<br />

+ dr (N,m+1)<br />

n−1 ,<br />

+ as (N,m−1)<br />

n−1 .<br />

Next we consider boundary condition related to (6.3.6) <strong>in</strong> the classical<br />

case. When m = N,<br />

P (N,N)<br />

0<br />

for any ϕ ∈ Φ. SowetakeΞ (N,N)<br />

0<br />

If m =0,then<br />

p (N,N)<br />

0<br />

(ϕ) =||Ξ (N,N)<br />

ϕ|| 2 =0,<br />

= q (N,N)<br />

0<br />

P (N,0)<br />

0<br />

for any ϕ ∈ Φ. SowechooseΞ (N,0)<br />

0<br />

Let<br />

p (N,0)<br />

0<br />

v (N,m)<br />

n<br />

= a, q (N,0)<br />

0<br />

=<br />

�<br />

0<br />

= O2, i.e.,<br />

= r (N,N)<br />

0<br />

= s (N,N)<br />

0<br />

(ϕ) =||Ξ (N,0)<br />

ϕ|| 2 =1,<br />

p (N,m)<br />

n<br />

r (N,m)<br />

n<br />

0<br />

=0.<br />

= I2. From (6.2.2), we have<br />

= d, r (N,0)<br />

0<br />

�<br />

, w (N,m)<br />

n<br />

Then we see that for n ≥ 1and1≤ m ≤ N − 1,<br />

= c, s (N,0)<br />

0<br />

=<br />

�<br />

q (N,m)<br />

n<br />

s (N,m)<br />

n<br />

= b.<br />

�<br />

.


Quantum Walks 393<br />

v (N,m)<br />

� �<br />

ac<br />

n = v<br />

00<br />

(N,m−1)<br />

� �<br />

00<br />

n−1 + v<br />

bd<br />

(N,m+1)<br />

n−1 ,<br />

w<br />

(6.4.15)<br />

(N,m)<br />

� �<br />

00<br />

n = w<br />

ca<br />

(N,m−1)<br />

� �<br />

db<br />

n−1 + w<br />

00<br />

(N,m+1)<br />

n−1 , (6.4.16)<br />

and for 1 ≤ m ≤ N,<br />

Moreover,<br />

v (N,0)<br />

n<br />

v (N,0)<br />

0<br />

w (N,0)<br />

0<br />

= v (N,N)<br />

n<br />

� �<br />

a<br />

= , v<br />

c<br />

(N,m)<br />

� �<br />

0<br />

0 = ,<br />

0<br />

� �<br />

d<br />

= , w<br />

b<br />

(N,m)<br />

� �<br />

0<br />

0 = .<br />

0<br />

= w (N,0)<br />

n<br />

= w (N,N)<br />

n<br />

=<br />

� �<br />

0<br />

0<br />

(n ≥ 1).<br />

From now on we focus on 1 ≤ m ≤ N − 1 case. So we consider only n ≥ 1.<br />

, it is easily shown that there exist<br />

only two types of paths, that is, P ...P and P ...Q. Therefore we see that<br />

=0(n≥ 1). So we have<br />

Moreover from the def<strong>in</strong>ition of Ξ (N,m)<br />

n<br />

q (N,m)<br />

n<br />

= s (N,m)<br />

n<br />

Lemma 6.1.<br />

∞�<br />

P (N,m) (ϕ) = P<br />

n=1<br />

(N,m)<br />

n (ϕ),<br />

P (N,m)<br />

n (ϕ) =C1(n)|α| 2 + C2(n)|β| 2 +2ℜ(C3(n)αβ),<br />

where ϕ = T [α, β] ∈ Φ and<br />

C1(n) =|ap (N,m)<br />

n<br />

C2(n) =|bp (N,m)<br />

n<br />

C3(n) =(ap (N,m)<br />

n<br />

+ cr (N,m)<br />

n | 2 ,<br />

+ dr (N,m)<br />

n<br />

| 2 ,<br />

+ cr (N,m)<br />

n )(bp (N,m)<br />

n<br />

+ dr (N,m)<br />

n ).<br />

To solve P (N,m) (ϕ), we <strong>in</strong>troduce generat<strong>in</strong>g functions of p (N,m)<br />

n<br />

as follows:<br />

By (6.4.15), we have<br />

p (N,m) (z) =<br />

r (N,m) (z) =<br />

∞�<br />

n=1<br />

∞�<br />

n=1<br />

p (N,m)<br />

n z n ,<br />

r (N,m)<br />

n z n .<br />

p (N,m) (z) =azp (N,m−1) (z)+czr (N,m−1) (z),<br />

r (N,m) (z) =bzp (N,m+1) (z)+dzr (N,m+1) (z).<br />

and r (N,m)<br />

n


394 N. Konno<br />

Solv<strong>in</strong>g these, we see that both p (N,m) (z) andr (N,m) (z) satisfy the same<br />

recurrence:<br />

dp (N,m+2) �<br />

(z) − △z + 1<br />

�<br />

p<br />

z<br />

(N,m+1) (z)+ap (N,m) (z) =0,<br />

dr (N,m+2) �<br />

(z) − △z + 1<br />

�<br />

r<br />

z<br />

(N,m+1) (z)+ar (N,m) (z) =0.<br />

From the characteristic equations with respect to the above recurrences, we<br />

have the same roots: if a �= 0,then<br />

λ± = △z2 +1∓ � △2z4 +2△(1 − 2|a| 2 )z2 +1<br />

,<br />

2△az<br />

where △ =detU = ad − bc.<br />

From now on we consider ma<strong>in</strong>ly U = H (the Hadamard walk) with N =<br />

=<br />

∞. Remark that the def<strong>in</strong>ition of Ξ (∞,1)<br />

n<br />

gives p (∞,1)<br />

n<br />

=0(n ≥ 2) and p (∞,1)<br />

1<br />

1. So we have p (∞,1) (z) =z. Moreover not<strong>in</strong>g limm→∞ p (∞,m) (z) < ∞, the<br />

follow<strong>in</strong>g explicit form is obta<strong>in</strong>ed:<br />

where<br />

Therefore for m =1,<br />

p (∞,m) (z) =zλ m−1<br />

+ ,<br />

r (∞,m) (z) = −1+√ z 4 +1<br />

z<br />

λ± = z2 − 1 ± √ z4 +1<br />

√ .<br />

2z<br />

r (∞,1) (z) = −1+√z4 +1<br />

.<br />

z<br />

λ m−1<br />

+ ,<br />

From the above equation and the def<strong>in</strong>ition of the hypergeometric series<br />

2F1(a, b; c; z), we have<br />

∞�<br />

(r (∞,1)<br />

n=1<br />

n<br />

) 2 z n =<br />

∞�<br />

n=1<br />

� �2 1/2<br />

n<br />

On the other hand, it should be noted that<br />

2F1(a, b; c;1)=<br />

Γ (c)Γ (c − a − b)<br />

Γ (c − a)Γ (c − b)<br />

where Γ (z) is the gamma function def<strong>in</strong>ed by<br />

z 4n−1 = 2F1(−1/2, −1/2; 1; z4 ) − 1<br />

.<br />

z<br />

(ℜ(a + b − c) < 0),


Quantum Walks 395<br />

� ∞<br />

Γ (z) = e<br />

0<br />

−t t z−1 dt (ℜ(z) > 0).<br />

Therefore<br />

∞�<br />

(r<br />

n=1<br />

(∞,1)<br />

n ) 2 Γ (1)Γ (2) 4<br />

= − 1= − 1,<br />

Γ (3/2) 2 π<br />

s<strong>in</strong>ce Γ (1) = Γ (2) = 1 and Γ (3/2) = √ π/2. By Lemma 6.1,<br />

P (∞,1) (ϕ) =<br />

+p (∞,1)<br />

n<br />

∞�<br />

n=1<br />

�<br />

1<br />

�<br />

2<br />

r (∞,1)<br />

n<br />

(p (∞,1)<br />

n<br />

) 2 +(r (∞,1)<br />

n<br />

(|α| 2 −|β| 2 )+ 1<br />

�<br />

2<br />

) 2�<br />

(p (∞,1)<br />

n<br />

) 2 − (r (∞,1)<br />

n<br />

) 2�<br />

(α ¯ �<br />

β +¯αβ) .<br />

Note that p (∞,1)<br />

n r (∞,1)<br />

n =0(n ≥ 1), s<strong>in</strong>ce p (∞,1)<br />

n =0(n≥ 2), p (∞,1)<br />

1 =1and<br />

r (∞,1)<br />

1<br />

=0.Sowehave<br />

P (∞,1) (ϕ) = 2<br />

π +2<br />

�<br />

1 − 2<br />

�<br />

ℜ(αβ), (6.4.17)<br />

π<br />

for any <strong>in</strong>itial qubit state ϕ = T [α, β] ∈ Φ. This result is a generalization of<br />

(6.3.13) given by Amba<strong>in</strong>is et al. (2001). From (6.4.17), we get the range of<br />

P (∞,1) (ϕ):<br />

4 − π<br />

π ≤ P (∞,1) (ϕ) ≤ 1.<br />

The equality holds <strong>in</strong> the follow<strong>in</strong>g cases:<br />

P (∞,1) (ϕ) =1iffϕ = eiθ<br />

√ 2<br />

� �<br />

1<br />

, and P<br />

1<br />

(∞,1) 4 − π<br />

(ϕ) =<br />

π<br />

iff ϕ = eiθ<br />

� �<br />

1<br />

√ ,<br />

2 −1<br />

where 0 ≤ θ


396 N. Konno<br />

The above formula gives<br />

f ′ (z) = z42F1(1/2, 1/2; 2; z4 ) − 2F1(−1/2, −1/2; 1; z4 )+1<br />

z2 .<br />

Moreover not<strong>in</strong>g<br />

2F1(−1/2, −1/2; 1; 1) = 2F1(1/2, 1/2; 2; 1) = 4<br />

π ,<br />

we have f ′ (1) = 1. Therefore the desired conclusion is obta<strong>in</strong>ed:<br />

E (∞,1) (T0|T0 < ∞) =<br />

s<strong>in</strong>ce<br />

∞�<br />

nP (∞,1) (T0 = n)<br />

n=1<br />

=<br />

∞�<br />

�<br />

1<br />

�<br />

n<br />

2<br />

n=1<br />

(p (∞,1)<br />

n<br />

) 2 +(r (∞,1)<br />

n<br />

�∞ n=1 nP (∞,1) (T0 = n)<br />

�∞ n=1 P (∞,1) (T0 = n) =<br />

1<br />

P (∞,1) (ϕ) ,<br />

) 2�<br />

+ 1<br />

�<br />

2<br />

(p (∞,1)<br />

n<br />

= 1<br />

2 {1+f ′ (1)} + 1<br />

2 {1 − f ′ (1)}(α ¯ β +¯αβ) =1.<br />

In a similar way we see that f ′′ (1) = ∞ implies<br />

E (∞,1) ((T0) 2 |T0 < ∞) =∞,<br />

) 2 − (r (∞,1)<br />

n ) 2�<br />

(α ¯ �<br />

β +¯αβ)<br />

so the ℓ-th moment E (∞,1) ((T0) ℓ |T0 < ∞) diverges for ℓ ≥ 2.<br />

Next we consider the f<strong>in</strong>ite N case. Then p (N,m) (z) andr (N,m) (z) satisfy<br />

p (N,m) (z) =Azλ m−1<br />

+<br />

r (N,m) (z) =Czλ m−N+1<br />

+<br />

+ Bzλ m−1<br />

− ,<br />

+ Dzλ m−N+1<br />

− ,<br />

s<strong>in</strong>ce λ+λ− = −1.<br />

Allwehavetodoistodeterm<strong>in</strong>ethecoefficientsAz,Bz,Cz,Dz by us<strong>in</strong>g<br />

the boundary conditions: p (N,1) (z) =z and r (N,N−1) (z) =0comefromthe<br />

def<strong>in</strong>ition of Ξ (N,m)<br />

n . The boundary conditions imply Cz + Dz =0andAz +<br />

Bz = z, sowesee<br />

p (N,m) �<br />

z<br />

�<br />

(z) = + Ez λ<br />

2 m−1<br />

�<br />

z<br />

�<br />

+ + − Ez λ<br />

2 m−1<br />

− , (6.4.18)<br />

r (N,m) (z) =Cz(λ m−N+1<br />

+<br />

where Ez = Az − z/2=z/2 − Bz.<br />

− λ m−N+1<br />

− ), (6.4.19)


Quantum Walks 397<br />

To obta<strong>in</strong> Ez and Cz, we use r (N,1) (z) = (p (N,2) (z) − r (N,2) (z))z/ √ 2<br />

and r (N,N−2) (z) = (p (N,N−1) (z) − r (N,N−1) (z))z/ √ 2 = p (N,N−1) (z)z/ √ 2.<br />

Therefore<br />

Cz(λ+ − λ−) = z ��<br />

z<br />

�<br />

√ + Ez λ<br />

2 2 N−2<br />

�<br />

z<br />

�<br />

+ + − Ez λ<br />

2 N−2<br />

�<br />

− ,<br />

Cz(λ N−2<br />

+ − λ N−2<br />

− )= z � �z �<br />

√ + Ez (−1)<br />

2 2 N−1 �<br />

z<br />

�<br />

λ+ + − Ez (−1)<br />

2 N−1 λ−<br />

− λN−3)<br />

�<br />

.<br />

Solv<strong>in</strong>g the above equations gives<br />

Cz = z2<br />

√ (−1)<br />

2 N−2 (λ N−3<br />

+<br />

�<br />

×<br />

(λ N−2<br />

+<br />

Ez = −<br />

�<br />

× (λ N−2<br />

+<br />

− λN−3)<br />

−<br />

− λN−2 − )2 − z √ (λ<br />

2 N−2<br />

+<br />

2(λ N−2<br />

+<br />

By Lemma 6.1, we obta<strong>in</strong><br />

Theorem 6.2.<br />

where ϕ = T [α, β] ∈ Φ and<br />

C1 = 1<br />

2π<br />

C2 = 1<br />

2π<br />

C3 = 1<br />

2π<br />

−<br />

+Cz(λ N−3<br />

+<br />

− λN−2 − )(λN−3 + − λN−3 − )<br />

−(−1) N−3 (λ+ − λ−) 2<br />

�−1 , (6.4.20)<br />

z<br />

− λN−2)<br />

�<br />

2(−1) N−3 (λ+ − λ−)(λ N−3<br />

+ − λN−3 − )<br />

− λ N−2<br />

− ) 2 − z √ (λ<br />

2 N−2<br />

+<br />

− λ N−2<br />

−<br />

−(−1) N−3 (λ+ − λ−) 2<br />

�−1 )(λ N−3<br />

+<br />

+(λ N−2<br />

+<br />

P (N,m) (ϕ) =C1|α| 2 + C2|β| 2 +2ℜ(C3αβ),<br />

− λ N−3<br />

− )<br />

+ λN−2<br />

−<br />

)<br />

�<br />

−<br />

. (6.4.21)<br />

� 2π<br />

|ap<br />

0<br />

(N,m) (e iθ )+cr (N,m) (e iθ )| 2 dθ,<br />

� 2π<br />

|bp<br />

0<br />

(N,m) (e iθ )+dr (N,m) (e iθ )| 2 dθ,<br />

� 2π<br />

(ap<br />

0<br />

(N,m) (eiθ )+cr (N,m) (eiθ ))(bp (N,m) (e iθ )+dr (N,m) (e iθ ))dθ,<br />

with a = b = c = −d =1/ √ 2,herep (N,m) (z) and r (N,m) (z) satisfy (6.4.18)<br />

and (6.4.19), and Cz and Ez satisfy (6.4.20) and (6.4.21).<br />

Therefore to compute P (N,m) (ϕ), we need to obta<strong>in</strong> both p (N,m) (z) and<br />

r (N,m) (z), and calculate the above Ci(i =1, 2, 3) explicitely.


398 N. Konno<br />

Here we consider U = H (the Hadamard walk), ϕ = T [α, β] andm =1.<br />

From Theorem 6.2, not<strong>in</strong>g that p (N,1) (z) =z for any N ≥ 2, we have<br />

Corollary 6.3. For N ≥ 2,<br />

P (N,1) (ϕ) = 1<br />

�<br />

1+<br />

2<br />

1<br />

2π<br />

� 2π<br />

where r (2,1) (z) =0,r (3,1) (z) =z 3 /(2 − z 2 ),<br />

and <strong>in</strong> general for N ≥ 4<br />

with<br />

r (N,1) (z) =−<br />

Jn(z) =<br />

n�<br />

ℓ=0<br />

0<br />

|r (N,1) (e iθ )| 2 �<br />

dθ (1 + 2ℜ(αβ)),<br />

r (4,1) (z) = z3 (1 − z2 )<br />

2 − 2z2 ,<br />

+ z4 r (5,1) (z) = z3 (2 − 3z2 +2z4 )<br />

4 − 6z2 +5z4 ,<br />

− 2z6 r (6,1) (z) = 2z3 (1 − z2 )(1 − z2 + z4 )<br />

4 − 8z2 +9z4 − 6z6 ,<br />

+2z8 z 2 JN−3(z)JN−4(z)<br />

√<br />

2(JN−3(z)) 2 − zJN−3(z)JN−4(z) − √ ,<br />

2(−1) N−3<br />

λ ℓ + λn−ℓ − , λ+ + λ− = √ �<br />

2 z − 1<br />

�<br />

, λ+λ− = −1.<br />

z<br />

In particular, when ϕ = T [0, 1] = |R〉, m=1andN =2,...,6, Corollary 6.3<br />

gives<br />

P (2,1) ( T [0, 1]) = 1<br />

2 , P(3,1) ( T [0, 1]) = 2<br />

3 , P(4,1) ( T [0, 1]) = 7<br />

10 ,<br />

P (5,1) ( T [0, 1]) = 12<br />

17 , P(6,1) ( T [0, 1]) = 41<br />

58 .<br />

It is easily checked that the above values P (N,1) ( T [0, 1])(N =2,...,6) satisfy<br />

the conjecture given by (6.3.14).<br />

6.5 Summary<br />

In this chapter we consider absorption problems for quantum walks on<br />

{0, 1,...,N} for both N < ∞ and N = ∞ cases. Here we summarize the<br />

results.


Quantum Walks 399<br />

and<br />

First we describe N = ∞ case. In this case, we have<br />

P (∞,1) (ϕ) = 2<br />

π +2<br />

E (∞,1) (T0|T0 < ∞) =<br />

�<br />

1 − 2<br />

π<br />

�<br />

ℜ(αβ),<br />

1<br />

P (∞,1) (ϕ) ,<br />

E (∞,1) ((T0) ℓ |T0 < ∞) =∞ (ℓ ≥ 2),<br />

for any <strong>in</strong>itial qubit state ϕ = T [α, β] ∈ Φ, whereP (∞,1) (ϕ) is the probability<br />

that the particle first hits location 0 start<strong>in</strong>g from location 1 and<br />

E (∞,1) ((T0) ℓ |T0 < ∞) is the conditional ℓ-th moment of T0 start<strong>in</strong>g from<br />

location 1 given {T0 < ∞} where T0 is the first hitt<strong>in</strong>g time to location 0.<br />

Next we describe N


Part II: Cont<strong>in</strong>uous-Time<br />

Quantum Walks<br />

7 One-Dimensional Lattice<br />

7.1 Introduction<br />

The quantum walk can be considered as a quantum analog of the classical<br />

random walk. However there are some differences between them. For the<br />

discrete-time symmetric classical random walk Y o<br />

t start<strong>in</strong>g from the orig<strong>in</strong>,<br />

the central limit theorem shows that Y o<br />

t /√t → e−x2 √<br />

/2dx/ 2π as t →∞. The<br />

same weak limit theorem holds for the cont<strong>in</strong>uous-time classical symmetric<br />

random walk. On the other hand, concern<strong>in</strong>g a discrete-time quantum walk<br />

X (d)<br />

t with a symmetric distribution on the l<strong>in</strong>e, whose evolution is described<br />

by the Hadamard transformation, it was shown by the author (Konno (2002a,<br />

/t → dx/π(1 −<br />

2005a)) that the follow<strong>in</strong>g weak limit theorem holds: X (d)<br />

t<br />

x 2 ) √ 1 − 2x 2 as t →∞. This chapter presents a similar type of weak limit<br />

theorem for a cont<strong>in</strong>uous-time quantum walk X (c)<br />

t<br />

on the l<strong>in</strong>e, X (c)<br />

t /t →<br />

dx/π √ 1 − x 2 as t →∞. Both limit density f unctions for quantum walks have<br />

two peaks at the two end po<strong>in</strong>ts of the supports. This chapter deals also with<br />

the issue of the relationship between discrete and cont<strong>in</strong>uous-time quantum<br />

walks. This topic, subject of a long debate <strong>in</strong> the previous literature, is treated<br />

with<strong>in</strong> the formalism of matrix representation and the limit distributions are<br />

exhaustively compared <strong>in</strong> the two cases.<br />

As a corollary, we have the follow<strong>in</strong>g result. Let σ c (d) (t) (resp. σc (c) (t))<br />

be the standard deviation of the probability distribution for a discrete-time<br />

(resp. cont<strong>in</strong>uous-time) classical random walk on the l<strong>in</strong>e start<strong>in</strong>g from the<br />

orig<strong>in</strong> at time t. Similarly, σ q<br />

(d) (t) (resp.σq (c) (t)) denotes the standard deviation<br />

for a discrete-time (resp. cont<strong>in</strong>uous-time) quantum walk. Then, it<br />

is well known that σc (d) (t), σc (c) (t) ≍ √ t, where f(t) ≍ g(t) <strong>in</strong>dicates that<br />

f(t)/g(t) → c∗(�=0) as t →∞. In contrast, σ q<br />

(d) (t), σq<br />

(c) (t) ≍ t holds. That is,<br />

U. Franz, M. Schürmann (eds.) Quantum <strong>Potential</strong> <strong>Theory</strong>. 401<br />

Lecture Notes <strong>in</strong> Mathematics 1954.<br />

c○ Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg 2008


402 N. Konno<br />

the quantum walks spread over the l<strong>in</strong>e faster than the classical walks <strong>in</strong> the<br />

both discrete- and cont<strong>in</strong>uous-time cases. Results <strong>in</strong> this chapter appeared <strong>in</strong><br />

Konno (2005c).<br />

7.2 Model and Results<br />

To def<strong>in</strong>e the cont<strong>in</strong>uous-time quantum walk on Z, we <strong>in</strong>troduce an ∞×∞<br />

adjacency matrix of Z denoted by A as follows:<br />

A =<br />

... −3 −2 −1 0 +1 +2 ...<br />

⎛ . ⎞<br />

. .. · · · · · · ...<br />

⎜<br />

⎟<br />

−3 ⎜ ... 0 1 0 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−2 ⎜ ... 1 0 1 0 0 0 ... ⎟<br />

⎜<br />

⎟<br />

−1 ⎜ ... 0 1 0 1 0 0 ... ⎟<br />

⎜<br />

⎟.<br />

0 ⎜ ... 0 0 1 0 1 0 ... ⎟<br />

⎜<br />

⎟<br />

+1 ⎜ ... 0 0 0 1 0 1 ... ⎟<br />

⎜<br />

⎟<br />

+2 ⎝ ... 0 0 0 0 1 0 ... ⎠<br />

.<br />

. ... · · · · · · ..<br />

The amplitude wave function of the walk at time t, Ψt, is def<strong>in</strong>ed by<br />

where<br />

Ψt = UtΨ0,<br />

Ut = e itA/2 .<br />

Note that Ut is a unitary matrix. As <strong>in</strong>itial state, we take<br />

Ψ0 = T [...,0, 0, 0, 1, 0, 0, 0,...].<br />

Let Ψt(x) be the amplitude wave function at location x at time t. The probability<br />

that the particle is at location x at time t, Pt(x), is given by<br />

Pt(x) =|Ψt(x)| 2 .<br />

There are various results for a walk on a cycle CN ,whereCN = {0, 1,...,<br />

N − 1} (see Adamczak et al. (2003), Ahmadi et al. (2003), Inui, Kasahara,<br />

Konishi, and Konno (2005)).<br />

We will obta<strong>in</strong> an explicit form of Ut first. Our approach is based on a<br />

direct computation of the ∞×∞ matrix A without us<strong>in</strong>g its eigenvalues and<br />

eigenvectors <strong>in</strong> order to clarify between the cont<strong>in</strong>uous-time and discrete-time<br />

quantum walks, (another approach and the same explicit form can be found<br />

<strong>in</strong> Section III C of Childs et al. (2003)). Let Jk(x) denote the Bessel function


Quantum Walks 403<br />

of the first k<strong>in</strong>d of order k. As for the Bessel function, see Watson (1944) and<br />

Chapter 4 <strong>in</strong> Andrews, Askey, and Roy (1999).<br />

Proposition 7.1. In our sett<strong>in</strong>g, we have<br />

Ut =<br />

... −3 −2 −1 0 +1 +2 ...<br />

. ⎛<br />

. .<br />

.<br />

..<br />

· · · · · · ...<br />

⎜<br />

−3 ⎜<br />

... J0(t) iJ1(t) i<br />

⎜<br />

⎝<br />

2 J2(t) i 3 J3(t) i 4 J4(t) i 5 J5(t) ...<br />

−2 ... iJ1(t) J0(t) iJ1(t) i 2 J2(t) i 3 J3(t) i 4 J4(t) ...<br />

−1 ... i 2 J2(t) iJ1(t) J0(t) iJ1(t) i 2 J2(t) i 3 J3(t) ...<br />

0 ... i 3 J3(t) i 2 J2(t) iJ1(t) J0(t) iJ1(t) i 2 J2(t) ...<br />

+1 ... i 4 J4(t) i 3 J3(t) i 2 J2(t) iJ1(t) J0(t) iJ1(t) ...<br />

+2 ... i 5 J5(t) i 4 J4(t) i 3 J3(t) i 2 ⎞<br />

⎟<br />

⎟.<br />

⎟<br />

J2(t) iJ1(t) J0(t) ... ⎟<br />

⎠<br />

.<br />

.<br />

. ... · · · · · ·<br />

..<br />

That is, the (l, m) component of Ut is given by i |l−m| J |l−m|(t).<br />

From Proposition 7.1, look<strong>in</strong>g at a column of Ut, we have immediately<br />

Corollary 7.2. The amplitude wave function of our model is given by<br />

Ψt = T [...,i 3 J3(t),i 2 J2(t),iJ1(t),J0(t),iJ1(t),i 2 J2(t),i 3 J3(t),...],<br />

that is, Ψt(x) =i |x| J |x|(t) for any location x ∈ Z and time t ≥ 0.<br />

Moreover, not<strong>in</strong>g that J−x(t) =(−1) x Jx(t) ((4.5.4) <strong>in</strong> Andrews, Askey, and<br />

Roy (1999)), we have<br />

Corollary 7.3. The probability distribution is<br />

Pt(x) =J 2 |x| (t) =J 2 x (t),<br />

for any location x ∈ Z and time t ≥ 0.<br />

In fact, the follow<strong>in</strong>g result (see (4.9.5) <strong>in</strong> Andrews, Askey, and Roy (1999)):<br />

J 2 0 (t)+2<br />

∞�<br />

J 2 x(t) =1<br />

x=1<br />

ensures that � ∞<br />

x=−∞ Pt(x) = 1 for any t ≥ 0. Remark that the distribution<br />

is symmetric for any time, i.e., Pt(x) =Pt(−x).<br />

One of the <strong>in</strong>terest<strong>in</strong>g po<strong>in</strong>ts of the result is as follows. Fix a positive<br />

<strong>in</strong>teger r. We suppose that the unitary matrix U (r) has the follow<strong>in</strong>g form:


404 N. Konno<br />

U (r) =<br />

... −3 −2 −1 0 +1 +2 ...<br />

⎛<br />

⎞<br />

.<br />

. ..<br />

⎜<br />

· · · · · · ...<br />

⎟<br />

−3 ⎜<br />

... w0 w1 w2 w3 w4 w5 ... ⎟<br />

−2 ⎜<br />

... w−1 w0 w1 w2 w3 w4 ... ⎟<br />

−1 ⎜<br />

... w−2 w−1 w0 w1 w2 w3 ... ⎟<br />

0 ⎜<br />

... w−3 w−2 w−1 w0 w1 w2 ... ⎟,<br />

⎟<br />

+1 ⎜<br />

... w−4 w−3 w−2 w−1 w0 w1 ... ⎟<br />

+2 ⎜<br />

⎝<br />

... w−5 w−4 w−3 w−2 w−1 w0 ... ⎟<br />

⎠<br />

.<br />

.<br />

. ... · · · · · · ..<br />

with wx ∈ C (x ∈ Z) andws = 0 for |s| > r.ThentheNo-Go lemma<br />

(see Meyer (1996)) shows that the only non-zero w∗ have |w∗| =1;that<br />

is, there exists no non-trivial, homogeneous f<strong>in</strong>ite-range model governed by<br />

the U (r) . For example, when r =1,wehave“|w−1| =1,w0 = w1 = 0”,<br />

“|w0| =1,w−1 = w1 = 0”, or “|w1| =1,w0 = w−1 = 0”. Thus, the model<br />

has a trivial probability distribution. However, our homogeneous, but <strong>in</strong>f<strong>in</strong>iterange<br />

model has a non-trivial probability distribution given by squared Bessel<br />

functions (Corollary 7.3).<br />

One of the open problems of quantum walks is to clarify a relation between<br />

discrete-time and cont<strong>in</strong>uous-time quantum walks (see Amba<strong>in</strong>is (2003), for<br />

example). To expla<strong>in</strong> the reason, we <strong>in</strong>troduce the follow<strong>in</strong>g matrices as <strong>in</strong><br />

Chapter 4 (see also Hamada, Konno, and Segawa (2005)):<br />

PA =<br />

� ab<br />

00<br />

�<br />

, QA =<br />

� 00<br />

cd<br />

�<br />

and PB =<br />

� a 0<br />

c 0<br />

�<br />

, QB =<br />

� �<br />

0 b<br />

,<br />

0 d<br />

whereweassumethatU = Pj + Qj (j = A, B) isa2× 2 unitary matrix. Here<br />

we consider two types of the discrete-time case; one is A-type, the other is<br />

B-type. The precise def<strong>in</strong>ition is given <strong>in</strong> Chapter 4. Then the unitary matrix<br />

of the discrete-time quantum walk on the l<strong>in</strong>e is described as<br />

⎡<br />

⎤<br />

. ..<br />

⎢ . . . . . . ⎥<br />

U (d) ⎢ ... O Pj O O O ... ⎥<br />

⎢ ... Qj O Pj O O ... ⎥<br />

= ⎢ ... O Qj O Pj O ... ⎥<br />

⎢ ... O O Qj O Pj ... ⎥<br />

⎢<br />

⎣ ... O O O Qj O ... ⎥<br />

⎦<br />

.<br />

... . . . . . ..<br />

with O =<br />

� �<br />

00<br />

,<br />

00<br />

for j = A and B. The unitary matrix U (d) for the discrete-time case corresponds<br />

to Ut for our cont<strong>in</strong>uous-time case at time t = 1. More generally,<br />

(U (d) ) n corresponds to Un for n =0, 1,.... Once an explicit formula of Ut is<br />

obta<strong>in</strong>ed, the difference between cont<strong>in</strong>uous and discrete walks becomes clear.<br />

As we stated before, Ut has an <strong>in</strong>f<strong>in</strong>ite-range form. On the other hand, U (d)


Quantum Walks 405<br />

has a f<strong>in</strong>ite-range form. Moreover, we see that U (d) is not homogeneous. It is<br />

believed that the difference seems to be derived from the fact that discrete<br />

quantum walk has a co<strong>in</strong> but cont<strong>in</strong>uous quantum walk does not. However,<br />

the situation is not so simple, s<strong>in</strong>ce the discrete-time case also does not necessarily<br />

need the co<strong>in</strong> (see Chapter 4 for more detailed discussion).<br />

We def<strong>in</strong>e a cont<strong>in</strong>uous-time quantum walk on Z by Xt whose probability<br />

distribution is def<strong>in</strong>ed by P (Xt = x) =Pt(x) for any location x ∈ Z and time<br />

t ≥ 0. Note that it follows from Corollary 7.3 that Pt(x) =J2 x (t). Then we<br />

obta<strong>in</strong> a weak limit theorem for a cont<strong>in</strong>uous-time quantum walk on the l<strong>in</strong>e:<br />

Theorem 7.4. If t →∞,then<br />

Xt<br />

t ⇒ Z(c) ,<br />

where Z (c) has the follow<strong>in</strong>g density:<br />

1<br />

π √ 1 − x2 I (−1,1)(x) (x∈R). For a more general sett<strong>in</strong>g, Gottlieb (2005) obta<strong>in</strong>ed weak limit theorems for<br />

cont<strong>in</strong>uous-time quantum walks on Z by a different method. Note that<br />

� 1<br />

−1<br />

x2m π √ 2<br />

dx =<br />

1 − x2 π<br />

� π/2<br />

0<br />

s<strong>in</strong> 2m ϕdϕ=<br />

(2m − 1)!!<br />

, (7.2.1)<br />

(2m)!!<br />

for m =1, 2,...,where n!! = n(n−2)·····5·3·1, if n is odd, = n(n−2)·····6·4·2,<br />

if n is even. From Proposition 7.1 and (7.2.1), we have<br />

Corollary 7.5. For m =1, 2,...,<br />

E((Xt/t) 2m ) → (2m − 1)!!/(2m)!! (t →∞).<br />

By this corollary, for the standard deviation of our walk, σ q<br />

(c) (t), we see that<br />

σ q<br />

(c) (t)/t → 1/√ 2=0.70710 ... (t →∞).<br />

We consider a discrete-time quantum walk X (d)<br />

n with a symmetric distribution<br />

on the l<strong>in</strong>e, whose evolution is described by the Hadamard transfor-<br />

mation, that is,<br />

U = 1 √ 2<br />

� �<br />

1 1<br />

.<br />

1 −1<br />

The walk is often called the Hadamard walk. In contrast with the cont<strong>in</strong>uoustime<br />

case, for the Hadamard walk, the follow<strong>in</strong>g weak limit theorem holds<br />

(see Theorem 1.6 or Konno (2002a, 2005a)): If n →∞,then<br />

X (d)<br />

n<br />

n<br />

⇒ Z (d) ,


406 N. Konno<br />

where Z (d) has the follow<strong>in</strong>g density:<br />

As a corollary, we have<br />

1<br />

π(1 − x2 ) √ 1 − 2x2 I (−1/ √ 2,1/ √ 2) (x) (x∈R). σ q<br />

(d) (n)/n →<br />

�<br />

(2 − √ 2)/2 =0.54119 ... (n →∞).<br />

Compar<strong>in</strong>g with the discrete-time case, the scal<strong>in</strong>g <strong>in</strong> our cont<strong>in</strong>uous-time<br />

case is same, but the limit density function is different. However, both density<br />

functions have some similar properties, for example, they have two peaks at<br />

the end po<strong>in</strong>ts of the support.<br />

7.3 Proof of Proposition 7.1<br />

To beg<strong>in</strong> with, A is rewritten as<br />

where<br />

A =<br />

P =<br />

... −1 0 +1 +2 ...<br />

⎛<br />

. . ⎞<br />

.<br />

.. .<br />

.<br />

.<br />

.<br />

.<br />

−1<br />

⎜ ... T P O O ...<br />

⎟<br />

0<br />

⎜ ... Q T P O ...<br />

⎟<br />

+1<br />

⎜ ... O Q T P ...<br />

⎟,<br />

(7.3.2)<br />

⎟<br />

+2<br />

⎜<br />

⎝ ... O O Q T ...<br />

⎟<br />

⎠<br />

. ...<br />

.<br />

.<br />

.<br />

. .<br />

.<br />

..<br />

� �<br />

00<br />

, T =<br />

10<br />

� �<br />

01<br />

, Q =<br />

10<br />

� �<br />

01<br />

.<br />

00<br />

The follow<strong>in</strong>g algebraic relations are useful for some computations:<br />

P 2 = Q 2 = O, PT + TP = QT + TQ= I, PQ + QP = T, (7.3.3)<br />

where O is 2 × 2 zero matrix and I is 2 × 2 unit matrix. From now on, for<br />

simplicity, (7.3.2) is written as<br />

A direct computation gives<br />

A =[...,O,O,O,O,Q,T,P,O,O,O,O,O,...].<br />

A 2 =[...,O,O,O,O,I,2I,I,O,O,O,O,...],<br />

A 4 =[...,O,O,O,I,4I,6I,4I,I,O,O,O,...],<br />

A 6 =[...,O,O,I,6I,15I,20I,15I,6I,I,O,O,...].


Quantum Walks 407<br />

It follows by <strong>in</strong>duction that<br />

A 2n =[...,O,O,A (2n)<br />

−n ,...,A(2n) −1 ,A(2n) 0<br />

where A (2n)<br />

k<br />

= A(2n)<br />

−k<br />

= a(2n)<br />

k I and<br />

a (2n)<br />

k<br />

=<br />

� �<br />

2n<br />

,<br />

n − k<br />

,A (2n)<br />

,...,A (2n)<br />

,O,O,...],<br />

for any k =0, 1,...,n. On the other hand, by us<strong>in</strong>g A 2n+1 = A 2n × A and<br />

(7.3.3), we have<br />

A 2n+1 =[...,O,A (2n+1)<br />

−(n+1) ,...,A(2n+1) −1 ,A (2n+1)<br />

0<br />

where<br />

A (2n+1)<br />

−(n+1)<br />

A (2n+1)<br />

−(n−1) =(a(2n)<br />

n−1<br />

A (2n+1)<br />

−1 =(a (2n)<br />

1<br />

A (2n+1)<br />

1<br />

A (2n+1)<br />

n<br />

= a(2n) n Q, A(2n+1) −n<br />

=(a (2n)<br />

1<br />

= a (2n)<br />

n<br />

= a (2n)<br />

n<br />

+ a(2n) n )T +(a (2n)<br />

n−2<br />

+ a(2n) )T +(a (2n)<br />

− a(2n)<br />

2<br />

0<br />

1<br />

n<br />

,A (2n+1)<br />

,...,A (2n+1)<br />

,O,...],<br />

1<br />

T + a(2n) n−1Q, − a(2n) n )Q,...,<br />

2 )Q, A (2n+1)<br />

0<br />

+ a(2n) )T +(a (2n)<br />

− a(2n) )P,...,<br />

2<br />

T + a (2n)<br />

The def<strong>in</strong>ition of Ut gives<br />

Ut = e itA/2 =<br />

=<br />

n=0<br />

0<br />

n−1P, A(2n+1)<br />

(n+1)<br />

∞�<br />

n=0<br />

� �<br />

t 2n<br />

2<br />

� �<br />

it n<br />

2<br />

n! An<br />

2<br />

= a(2n) n P.<br />

n+1<br />

=(a (2n)<br />

0<br />

∞�<br />

(−1) n<br />

(2n)! A2n ∞�<br />

+ i (−1) n<br />

� �<br />

t 2n+1<br />

2<br />

(2n +1)! A2n+1 .<br />

n=0<br />

Therefore we obta<strong>in</strong> Ut = B(t)+iC(t), where<br />

with<br />

B(t) =[...,B−k,...,B−1,B0,B1,...,Bk,...],<br />

C(t) =[...,C−k,...,C−1,C0,C1,...,Ck,...],<br />

Bk = B−k = bkI, bk =<br />

∞�<br />

(−1) n<br />

(2n)!<br />

n=0<br />

� t<br />

2<br />

� 2n<br />

� �<br />

2n<br />

n − k<br />

+ a(2n) )T,<br />

1<br />

(k =0, 1,...),


408 N. Konno<br />

and<br />

C−k = ckI + dkQ, Ck = ckI + dkP (k =0, 1,...),<br />

∞�<br />

ck = (−1)<br />

n=0<br />

n<br />

� �<br />

t 2n+1 �� � � ��<br />

2<br />

2n<br />

2n<br />

+<br />

(2n +1)! n − k n − (k +1)<br />

(k =0, 1,...),<br />

d0 =0,<br />

∞�<br />

dk = (−1) n<br />

� �<br />

t 2n+1 �� � � ��<br />

2<br />

2n<br />

2n<br />

+<br />

(2n +1)! n − (k − 1) n − (k +1)<br />

(k =1, 2,...).<br />

n=0<br />

Remark that<br />

Jk(x) =<br />

�<br />

x<br />

�k ∞ � �<br />

� ix 2m � �<br />

2 k +2m<br />

, (7.3.4)<br />

2 (k +2m)! m<br />

m=0<br />

(see (4.9.5) <strong>in</strong> Andrews, Askey, and Roy (1999)). F<strong>in</strong>ally, by us<strong>in</strong>g (7.3.4) and<br />

the follow<strong>in</strong>g relation:<br />

� � � � � �<br />

2n 2n 2n +1<br />

+ = ,<br />

k k − 1 k<br />

we have the desired conclusion.<br />

7.4 Proof of Theorem 7.4<br />

We beg<strong>in</strong> by stat<strong>in</strong>g the follow<strong>in</strong>g result (see page 214 <strong>in</strong> Andrews, Askey,<br />

and Roy (1999)): suppose that a, b, and c are lengths of sides of a triangle<br />

and c 2 = a 2 + b 2 − 2ab cos ξ. Then<br />

J0(c) =<br />

∞�<br />

k=−∞<br />

If we set t = a = b <strong>in</strong> (7.4.5), then<br />

J0(t � 2(1 − cos ξ)) =<br />

Jk(a)Jk(b)e ikξ . (7.4.5)<br />

∞�<br />

k=−∞<br />

J 2 k (t)eikξ . (7.4.6)<br />

From (7.4.6), we see that the characteristic function of a cont<strong>in</strong>uous-time<br />

quantum walk on the l<strong>in</strong>e is given by<br />

E(e iξXt )=<br />

∞�<br />

k=−∞<br />

e ikξ J 2 k (t) =J0(t � 2(1 − cos ξ)). (7.4.7)


Quantum Walks 409<br />

First we consider that t is a positive <strong>in</strong>teger case, that is, n(= t) =1, 2,....<br />

By us<strong>in</strong>g (7.4.7), we have<br />

E(e iξXn/n )=J0(n � 2(1 − cos(ξ/n))) → J0(ξ),<br />

as n →∞. To know the limit density function, we use the follow<strong>in</strong>g expression<br />

of J0(x) (see (4.9.11) <strong>in</strong> Andrews, Askey, and Roy (1999));<br />

J0(ξ) = 1<br />

π<br />

� π<br />

Tak<strong>in</strong>g x =s<strong>in</strong>ϕ, wehave<br />

J0(ξ) =<br />

� 1<br />

−1<br />

By us<strong>in</strong>g (7.4.8), we get<br />

0<br />

cos(ξ s<strong>in</strong> ϕ) dϕ = 2<br />

π<br />

1<br />

cos(ξx)<br />

π √ dx =<br />

1 − x2 |J0(ξ) − J0(0)| ≤ 2<br />

π<br />

� π/2<br />

0<br />

� π/2<br />

0<br />

� 1<br />

−1<br />

cos(ξ s<strong>in</strong> ϕ) dϕ. (7.4.8)<br />

e iξx 1<br />

π √ dx. (7.4.9)<br />

1 − x2 | cos(ξ s<strong>in</strong> ϕ) − 1| dϕ.<br />

From the bounded convergence theorem, we see that the limit J0(ξ) iscont<strong>in</strong>uous<br />

at ξ = 0, s<strong>in</strong>ce cos(ξ s<strong>in</strong> ϕ) → 1asξ → 0. Therefore, by the cont<strong>in</strong>uity<br />

theorem and (7.4.9), we conclude that if n →∞,thenXn/n converges<br />

weakly to a random variable whose density function is given by 1/π √ 1 − x 2<br />

for x ∈ (−1, 1). That is, if −1 ≤ a


410 N. Konno<br />

of discrete-time one, the limit theorem resembles that of the discrete-time<br />

case, for example, the symmetric Hadamard walk: X (d)<br />

n /n → dx/π(1 − x2 √ )<br />

1 − 2x2 (n →∞).<br />

Romanelli et al. (2004) <strong>in</strong>vestigated the cont<strong>in</strong>uum time limit for a discretetime<br />

quantum walk on Z and obta<strong>in</strong>ed the position probability distribution.<br />

When the <strong>in</strong>itial condition is given by ãl(0) = δl,0, ˜bl(0) ≡ 0 <strong>in</strong> their notation<br />

for the Hadamard walk, the distribution becomes the follow<strong>in</strong>g <strong>in</strong> our no-<br />

tation: P (R)<br />

t<br />

(x) =J 2 x (t/√ 2). More generally, we consider the time evolution<br />

given by the follow<strong>in</strong>g unitary matrix:<br />

U(θ) =<br />

� �<br />

cos θ s<strong>in</strong> θ<br />

,<br />

s<strong>in</strong> θ − cos θ<br />

where θ ∈ (0,π/2). Note that θ = π/4 case is equivalent to the Hadamard<br />

walk. Then we have P (R,θ)<br />

t (x) =J 2 x (t cos θ). In this case, a similar argument<br />

<strong>in</strong> the proof of Theorem 7.4 implies that if t →∞,then<br />

X (R,θ)<br />

t<br />

t<br />

⇒ Z (R,θ) ,<br />

where Z (R,θ) has the follow<strong>in</strong>g density function:<br />

Here X (R,θ)<br />

t<br />

tribution is given by P (R,θ)<br />

t<br />

1<br />

π √ cos 2 θ − x 2 I (− cos θ,cos θ)(x) (x ∈ R).<br />

denotes a cont<strong>in</strong>uous-time quantum walk whose probability dis-<br />

(x). As a consequence, we obta<strong>in</strong><br />

E((X (R,θ)<br />

t /t) 2m ) → cos 2m θ × (2m − 1)!!/(2m)!! (t →∞).<br />

In particular, when m = 1, the limit cos 2 θ/2 is consistent with (30) <strong>in</strong><br />

Romanelli et al. (2004).<br />

8 Tree<br />

8.1 Introduction<br />

In this chapter we consider a cont<strong>in</strong>uous-time quantum walk on a homogeneous<br />

tree <strong>in</strong> quantum probability theory. Results here appeared <strong>in</strong> Konno<br />

(2006a). The walk is def<strong>in</strong>ed by identify<strong>in</strong>g the Hamiltonian of the system<br />

with a matrix related to the adjacency matrix of the tree.<br />

Let T (p)<br />

M denote a homogeneous tree of degree p with M-generation. After<br />

we fix a root o ∈ T (p)<br />

M , a stratification (distance partition) is <strong>in</strong>troduced by


Quantum Walks 411<br />

the natural distance function <strong>in</strong> the follow<strong>in</strong>g way:<br />

T (p)<br />

M =<br />

M�<br />

k=0<br />

V (p)<br />

k , V (p)<br />

k = {x ∈ T(p)<br />

M : ∂(o, x) =k}.<br />

Here ∂(x, y) stands for the length of the shortest path connect<strong>in</strong>g x and y.<br />

Then<br />

|V (p)<br />

0 | =1, |V (p)<br />

1 | = p, |V (p)<br />

2 | = p(p − 1),...,|V (p)<br />

k | = p(p − 1) k−1 ,...,<br />

where |A| is the number of elements <strong>in</strong> a set A. The total number of po<strong>in</strong>ts<br />

<strong>in</strong> M-generation, |T (p)<br />

M |, is p(p − 1)M − (p − 1).<br />

be a |T(p)<br />

M |×|T(p) M | symmetric matrix given by the adjacency<br />

matrix of the tree T (p)<br />

M . The matrix is treated as the Hamiltonian of the<br />

quantum system. The (i, j) component of H (p)<br />

M denotes H(p)<br />

M (i, j) fori, j ∈<br />

{0, 1,...,|T (p)<br />

M |−1}. In our case, the diagonal component of H(p)<br />

M is always<br />

zero, i.e., H (p)<br />

M (i, i) = 0 for any i. On the other hand, the diagonal component<br />

of correspond<strong>in</strong>g matrix H (p)<br />

M,MB <strong>in</strong>vestigated <strong>in</strong> Mülken, Bierbaum,<br />

and Blumen (2006) for p = 3 is not zero. For example, H (3)<br />

1,MB (0, 0) = −3,<br />

H (3)<br />

1,MB (1, 1) = H(3)<br />

1,MB (2, 2) = H(3)<br />

1,MB (3, 3) = −1.<br />

The evolution of the cont<strong>in</strong>uous-time quantum walk on the tree of Mgeneration,<br />

T (p)<br />

M , is governed by the follow<strong>in</strong>g unitary matrix:<br />

Let H (p)<br />

M<br />

U (p)<br />

M (t) =eitH(p) M .<br />

The amplitude wave function at time t, Ψ (p)<br />

M (t), is def<strong>in</strong>ed by<br />

Ψ (p)<br />

M<br />

(p) (p)<br />

(t) =U M (t)Ψ M (0).<br />

In this chapter we take Ψ (p)<br />

M (0) = T [1, 0, 0,...,0] as <strong>in</strong>itial state.<br />

The (n + 1)-th coord<strong>in</strong>ate of Ψ (p)<br />

(p)<br />

M (t) isdenotedbyΨ M (n, t) whichisthe<br />

amplitude wave function at site n at time t for n =0, 1,...,p(p − 1) M − p.<br />

The probability f<strong>in</strong>d<strong>in</strong>g the walker at site n at time t on T (p)<br />

M is given by<br />

P (p)<br />

(p)<br />

M (n, t) =|Ψ M (n, t)|2 .<br />

Then we def<strong>in</strong>e the cont<strong>in</strong>uous-time quantum walk X (p)<br />

M (t)attimeton T(p)<br />

M by<br />

P (X (p)<br />

M<br />

(p)<br />

(t) =n) =P (n, t).<br />

In a similar way, let X (p)<br />

M,MB (t) be a quantum walk given by H(p)<br />

M,MB .Aswe<br />

stated before, H (p)<br />

M,MB (i, i) depends on i for any f<strong>in</strong>ite M. However <strong>in</strong> M →∞<br />

M


412 N. Konno<br />

limit, the (i, i) component of the matrix becomes −p for any i. Remarkthat<br />

the probability distribution of the cont<strong>in</strong>uous-time walk does not depend<br />

on the value of the diagonal component of the scalar matrix. Therefore the<br />

def<strong>in</strong>itions of the walks imply that both quantum walks co<strong>in</strong>cide <strong>in</strong> M →∞<br />

limit, i.e.,<br />

lim<br />

M→∞<br />

for any t and n.<br />

P (X(p)<br />

M (t) =n) = lim P (X(p)<br />

M→∞<br />

M,MB (t) =n),<br />

8.2 Quantum Probabilistic Approach<br />

F<strong>in</strong>ite M case<br />

Let µ (p)<br />

M denote the spectral distribution of our adjacency matrix H(p)<br />

M .From<br />

the general theory of an <strong>in</strong>teract<strong>in</strong>g Fock space (see Jafarizadeh and Salimi<br />

(2007), Accardi and Bo˙zejko (1998), Hashimoto (2001), Obata (2004), for<br />

examples), the orthogonal polynomials {Q (p)<br />

n } and {Q (p,∗)<br />

n } associated with<br />

µ (p)<br />

M satisfy the follow<strong>in</strong>g three-term recurrence relations with a Szegö-Jacobi<br />

parameter ({ωn}, {αn}) respectively:<br />

and<br />

Q (p)<br />

0<br />

(x) =1, Q(p) (x) =x − α1,<br />

1<br />

xQ (p)<br />

n (x) =Q (p)<br />

(p)<br />

n+1 (x)+αn+1Q n (x)+ωnQ (p)<br />

n−1 (x) (n≥ 1),<br />

Q (p,∗)<br />

0<br />

xQ (p,∗)<br />

n<br />

In our tree case,<br />

(x) =1, Q (p,∗)<br />

(x) =x − α2,<br />

1<br />

(x) =Q (p,∗)<br />

(p,∗)<br />

(x)+αn+2Q (x)+ωn+1Q (p,∗)<br />

(x) (n≥ 1).<br />

n+1<br />

n<br />

n−1<br />

ω1 = p, ω2 = ω3 = ···= ωM = p − 1, ωM+1 = ωM+2 = ···=0,<br />

Then the Stieltjes transform G µ (p)<br />

M<br />

α1 = α2 = ···=0.<br />

G µ (p)<br />

M<br />

of µ (p)<br />

M<br />

where n = |T (p)<br />

M | = p(p − 1)M − (p − 1).<br />

is given by<br />

(x) = Q(p,∗) n−1 (x)<br />

Q (p)<br />

n (x) ,


Quantum Walks 413<br />

The follow<strong>in</strong>g result was shown <strong>in</strong> Jafarizadeh and Salimi (2007):<br />

Ψ (p) (p)<br />

M (V k ,t)= �<br />

Ψ (p)<br />

1<br />

M (n, t) = �<br />

|V (p)<br />

k |<br />

�<br />

R<br />

exp(itx) Q (p)<br />

k (x) µ(p)<br />

M (x) dx,<br />

n∈V (p)<br />

k<br />

for k =0, 1, 2,.... Remark that |V (p)<br />

k | = ω1ω2 ···ωk = p(p − 1) k−1 (1 ≤ k ≤<br />

M) and|V (p)<br />

0 | =1. It is important to note that<br />

Ψ (p)<br />

M<br />

(n, t) = 1<br />

|V (p)<br />

k |<br />

�<br />

R<br />

exp(itx) Q (p)<br />

k (x)µ(p) M (x) dx,<br />

if n ∈ V (p)<br />

k (k = 0, 1,...,M). The proof appeared <strong>in</strong> Appendix A <strong>in</strong><br />

Jafarizadeh and Salimi (2007).<br />

p =3andM =2case<br />

Here we consider p =3andM =2case.Thenwehaven =10,ω1 =3,<br />

ω2 =2,ω3 = ω4 = ···=0,α1 = α2 = ···=0. The def<strong>in</strong>itions of Q (3)<br />

n (x) and<br />

Q (3,∗)<br />

n (x) imply<br />

Q (3)<br />

0<br />

and<br />

(x) =1, Q(3) (x) =x, Q(3)<br />

Q (3,∗)<br />

0<br />

1<br />

(x) =1, Q (3,∗)<br />

(x) =x, Q (3,∗)<br />

Therefore we obta<strong>in</strong> the Stieltjes transform:<br />

G µ (3)<br />

2<br />

1<br />

(x) = Q(3,∗) 9 (x)<br />

Q (3)<br />

10<br />

From this, we see that<br />

Then<br />

Ψ (3) (3)<br />

2 (V 0 ,t)=<br />

�<br />

µ (3)<br />

2<br />

R<br />

Ψ (3) (3)<br />

2 (V 1 ,t)= 1<br />

√<br />

ω1<br />

Ψ (3) (3)<br />

2 (V 2 ,t)=<br />

2 (x) =x2− 3, Q (3)<br />

k (x) =xk−2 (x 2 − 5) (k ≥ 3),<br />

2<br />

=<br />

(x) 5<br />

k<br />

1 3<br />

· +<br />

x 10 ·<br />

(x) =x k−2 (x 2 − 2) (k ≥ 2).<br />

1<br />

x + √ 5<br />

2 3<br />

= δ0(x)+<br />

5 10 δ− √ 3<br />

5 (x)+<br />

10 δ√ 5 (x).<br />

exp(itx) µ (3)<br />

2 (dx) =1<br />

�<br />

1<br />

√ ω1ω2<br />

R<br />

�<br />

5 (2 + 3 cos(√ 5t)),<br />

3<br />

+<br />

10 ·<br />

1<br />

x − √ 5 .<br />

exp(itx)Q (3)<br />

1 (x) µ(3) 2 (dx) =i√ 3<br />

√5 s<strong>in</strong>( √ 5t),<br />

R<br />

exp(itx)Q (3)<br />

2<br />

(x) µ(3) 2 (dx) =<br />

√ 6<br />

5<br />

�<br />

−1+cos( √ �<br />

5t) .


414 N. Konno<br />

Not<strong>in</strong>g that Ψ (3)<br />

�<br />

(3) (3)<br />

2 (n, t) =Ψ 2 (V k ,t)/ |V (3)<br />

k | for any k =0, 1, 2, we obta<strong>in</strong><br />

the same conclusion as the result given by the eigenvalues and the eigenvectors<br />

of H (3)<br />

2 .<br />

M →∞case<br />

The quantum probabilistic approach implies that<br />

Ψ (p) (p)<br />

(p) (p) 1<br />

∞ (V k ,t) = lim Ψ M (V<br />

M→∞<br />

k ,t)= �<br />

|V (p)<br />

k |<br />

�<br />

R<br />

exp(itx) Q (p)<br />

k (x)µ(p) M (x) dx,<br />

for k =0, 1, 2,..., where the limit spectral distribution µ (p)<br />

M (x) isgivenby<br />

p � 4(p − 1) − x 2<br />

2π(p 2 − x 2 )<br />

I (−2 √ p−1,2 √ p−1)(x) (x ∈ R).<br />

This type of measure was first obta<strong>in</strong>ed by Kesten (1959) <strong>in</strong> a classical random<br />

walk with a different method. An immediate consequence is<br />

P (p) (p)<br />

∞ (V k ,t)= 1<br />

|V (p)<br />

��� cos(tx) Q<br />

k | R<br />

(p)<br />

k (x)µ(p)<br />

�2 M (x) dx<br />

��<br />

+ s<strong>in</strong>(tx) Q (p)<br />

k (x)µ(p)<br />

� �<br />

2<br />

M (x) dx ,<br />

for k =0, 1, 2,... Furthermore, as <strong>in</strong> the case of f<strong>in</strong>ite M, weseethat<br />

�<br />

exp(itx) Q (p)<br />

k (x)µ(p) M (x) dx, (8.2.1)<br />

if n ∈ V (p)<br />

k<br />

Ψ (p)<br />

∞ (n, t) = 1<br />

|V (p)<br />

k |<br />

R<br />

R<br />

(k =0, 1, 2,...). From (8.2.1) and the Riemann-Lebesgue lemma,<br />

we have limt→∞ Ψ (p)<br />

∞ (n, t) = 0, for any n, s<strong>in</strong>ceQ (p)<br />

k (x)µ(p) ∞ (x) ∈ L1 (R).<br />

Therefore we see that limt→∞ P (p)<br />

∞ (n, t) =0. So we conclude that ¯ P (p)<br />

∞ (n) =0,<br />

where ¯ P (p)<br />

∞ (n) is the time-averaged distribution of P (p)<br />

∞ (n, t).<br />

p =2andM →∞case<br />

In this subsection, we consider p =2andM→∞, i.e., Z1 case. Then we have<br />

Proposition 8.1.<br />

Ψ (2) (2)<br />

∞ (V<br />

0 ,t)=J0(2t), Ψ (2)<br />

∞<br />

(V (2)<br />

k ,t)= √ 2 i k Jk(2t) (k =1, 2,...),<br />

where Jn(x) is the Bessel function of the first k<strong>in</strong>d of order n.


Quantum Walks 415<br />

Proof. Induction on k. Fork = 0 case, we use the follow<strong>in</strong>g result (see (4) <strong>in</strong><br />

page 48 <strong>in</strong> Watson (1944)):<br />

� 1<br />

exp (isx) (1−x −1<br />

2 ) ν−1/2 dx =<br />

Γ (1/2)Γ (ν +1/2)<br />

(s/2) ν Jν(s), (8.2.2)<br />

where Γ (x) is the Gamma function. Comb<strong>in</strong><strong>in</strong>g Γ (3/2) = √ π/2,Γ(1/2) = √ π<br />

with Q (2)<br />

0 (x) =1andν =0gives<br />

Ψ (2)<br />

∞ (V (2)<br />

� 2<br />

dx<br />

0 ,t)= exp (itx)<br />

−2 π √ = J0(2t).<br />

4 − x2 In a similar fashion, we verify that the result holds for k =1, 2.<br />

Next we suppose that the result is true for all values up to k, wherek≥2. Then we see that<br />

Ψ (2) (2) 1<br />

∞ (V k+1 ,t)= √<br />

2<br />

= 1<br />

√ 2<br />

= 1<br />

i<br />

d<br />

dt<br />

� 2<br />

−2<br />

� 2<br />

−2<br />

� 1√2<br />

exp (itx) Q (2)<br />

k+1 (x)<br />

�<br />

exp (itx)<br />

� 2<br />

−2<br />

xQ (2)<br />

k<br />

dx<br />

π √ 4 − x 2<br />

(x) − Q(2)<br />

k−1 (x)<br />

exp (itx) Q (2)<br />

k (x)<br />

− 1 √ 2<br />

= 1<br />

i<br />

= √ 2 i k+1 Jk+1(2t).<br />

� 2<br />

−2<br />

dx<br />

�<br />

π √ 4 − x 2<br />

exp (itx) Q (2)<br />

k−1 (x)<br />

d<br />

dt (√ 2 i k Jk(2t)) − √ 2 i k−1 Jk−1(2t)<br />

dx<br />

π √ 4 − x2 �<br />

dx<br />

π √ 4 − x 2<br />

The second equality follows from the def<strong>in</strong>ition of Q (2)<br />

(x). By <strong>in</strong>duction, we<br />

k<br />

have the fourth equality. For the last equality, we use a recurrence formula<br />

for the Bessel coefficients: 2J ′ k (2t) =Jk−1(2t) − Jk+1(2t) (see (2) <strong>in</strong> page 17<br />

of Watson (1944)).<br />

As a consequence, we have<br />

Corollary 8.2.<br />

P (2) (2)<br />

∞ (V 0 ,t)=J 2 (2)<br />

0 (2t), P(2) ∞ (V k ,t)=2J 2 k (2t) (k =1, 2,...).<br />

We confirm that<br />

∞�<br />

k=0<br />

P (2) (2)<br />

∞ (V k ,t)=1,


416 N. Konno<br />

s<strong>in</strong>ce it follows from J 2 0 (2t)+2 �∞ k=1 J 2 k (2t) = 1 (see (3) <strong>in</strong> page 31 <strong>in</strong> Watson<br />

(1944)). Not<strong>in</strong>g that V (2)<br />

k = {−k, k} for any k ≥ 0, we have an equivalent<br />

result given by Corollary 7.3:<br />

for any n ∈ Z and t ≥ 0.<br />

P (2)<br />

∞ (n, t) =J 2 n (2t),<br />

8.3 Quantum Central Limit Theorem<br />

To state a quantum central limit theorem <strong>in</strong> our case, it is convenient to<br />

rewrite as<br />

� � �<br />

�<br />

�exp<br />

�� �<br />

�� (p)<br />

Φ<br />

where<br />

Φ (p)<br />

k<br />

Φ (p)<br />

k =<br />

itH (p)<br />

∞<br />

1<br />

0<br />

�<br />

|V (p)<br />

k |<br />

�<br />

n∈V (p)<br />

k<br />

= Ψ (p)<br />

∞ (V (p)<br />

k ,t),<br />

I {n},<br />

and I {n} denotes the <strong>in</strong>dicator function of the s<strong>in</strong>gleton {n}. It is easily<br />

obta<strong>in</strong>ed that<br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

� �<br />

�<br />

k �exp itH (p)<br />

�� �<br />

�� (p)<br />

∞ Φ 0 =0,<br />

for any k ≥ 0. Then we have the follow<strong>in</strong>g quantum central limit theorem:<br />

Theorem 8.3.<br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

k �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

k Jk+1(2t)<br />

0 =(k +1)i ,<br />

t<br />

for k =0, 1, 2,....<br />

Proof. Induction on k. First we consider k = 0 case. We see that<br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

0 �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

0<br />

� �<br />

= lim exp it<br />

p→∞<br />

R<br />

x �<br />

√ µ<br />

p<br />

(p)<br />

M (x) dx<br />

�<br />

√ �<br />

2 (p−1)/p<br />

(2(p − 1)/p) 2 − x2 = lim<br />

exp (itx)<br />

p→∞<br />

2π(1 − x2 dx<br />

/p)<br />

=<br />

� 1<br />

−1<br />

−2 √ (p−1)/p<br />

exp (2itx) 2√ 1 − x 2<br />

π<br />

dx.


Quantum Walks 417<br />

Then (8.2.2) with ν =1yields<br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

0 �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

0<br />

= J1(2t)<br />

.<br />

t<br />

So the result holds for k = 0. Similarly we obta<strong>in</strong><br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

1 �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

0 = 2iJ2(2t)<br />

,<br />

t<br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

2 �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

0 = − 3J3(2t)<br />

.<br />

t<br />

Next we suppose that the result holds for all values up to k, wherek≥2. Then we have<br />

�<br />

lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

k+1 �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

0<br />

1<br />

= lim �<br />

p→∞<br />

|V (p)<br />

k+1 |<br />

� �<br />

exp it<br />

R<br />

x �<br />

√ Q<br />

p<br />

(p)<br />

k+1 (x) µ(p)<br />

M (x) dx<br />

�<br />

√<br />

2 (p−1)/p<br />

1<br />

= lim � exp (itx) Q<br />

p→∞ p(p − 1) k<br />

(p)<br />

k+1 (√ �<br />

(2(p − 1)/p) 2 − x2 px)<br />

2π(1 − x2 dx<br />

/p)<br />

=<br />

=<br />

� 2<br />

−2<br />

� 2<br />

= 1<br />

i<br />

−2<br />

−2 √ (p−1)/p<br />

exp (itx) Q (∞)<br />

k+1 (x)<br />

√<br />

22 − x2 dx<br />

2π<br />

�<br />

exp (itx) xQ (∞)<br />

k (x) − Q (∞)<br />

k−1 (x)<br />

� √ 22 − x2 dx<br />

2π<br />

�� 1<br />

d<br />

exp (2itx) Q<br />

dt<br />

(∞)<br />

k (2x) 2√1 − x2 �<br />

dx<br />

π<br />

−1<br />

= i k−1<br />

�<br />

(k +1) d<br />

dt<br />

� 1<br />

−<br />

−1<br />

exp (2itx) Q (∞)<br />

k−1 (2x) 2√ 1 − x 2<br />

� Jk+1(2t)<br />

t<br />

�<br />

− k Jk(2t)<br />

�<br />

,<br />

t<br />

where the last equality is given by the <strong>in</strong>duction and<br />

�<br />

p(p − 1) k−1 ,<br />

Q (∞)<br />

k (x) = lim<br />

p→∞ Q(p)<br />

k (√ px)/<br />

if the right-hand side exists. We confirm that the limit exists for any k ≥ 1.<br />

(x)=x, Q (∞)<br />

(x)=x2 − 1, Q (∞)<br />

(x)=x3 − 2x,<br />

For example, we compute Q (∞)<br />

1<br />

2<br />

π<br />

dx<br />

3


418 N. Konno<br />

Q (∞)<br />

4 (x) =x4 − 3x2 +1,... In order to prove the result, it suffices to check<br />

the follow<strong>in</strong>g relation:<br />

(k +1) d<br />

� �<br />

Jk+1(2t)<br />

− k<br />

dt t<br />

Jk(2t)<br />

= −(k +2)Jk+2(2t) .<br />

t<br />

t<br />

The left-hand side of this equation becomes<br />

(k +1) 2Jk+1(2t)<br />

t<br />

= Jk(2t)<br />

t<br />

− (k +1) Jk+1(2t)<br />

t 2<br />

− k Jk(2t)<br />

t<br />

− (k +1)Jk+1(2t)<br />

t2 − (k +1) Jk(2t)<br />

t<br />

= −(k +2) Jk+2(2t)<br />

,<br />

t<br />

s<strong>in</strong>ce the first and second equalities are obta<strong>in</strong>ed from recurrence formulas for<br />

the Bessel coefficients: 2J ′ k+1 (2t) =Jk(2t)−Jk+2(2t) andJk(2t)+Jk+2(2t) =<br />

(k +1)Jk+1(2t)/t (see (1) <strong>in</strong> page 17 of Watson (1944)), respectively. This<br />

f<strong>in</strong>ishes the proof of the theorem.<br />

Remark that an extension of this theorem for a wider family of graphs and<br />

another proof by us<strong>in</strong>g Gegenbauer’s <strong>in</strong>tegral formula for the Bessel function<br />

can be found <strong>in</strong> Obata (2006).<br />

8.4 A New Type of Limit Theorems<br />

We can now state the ma<strong>in</strong> result of this chapter. To do so, let def<strong>in</strong>e<br />

˜Ψ (∞)<br />

�<br />

(∞)<br />

∞ (V k ,t) = lim Φ<br />

p→∞<br />

(p)<br />

�<br />

�<br />

�<br />

k �<br />

� exp<br />

�<br />

it H(p)<br />

��<br />

����<br />

∞<br />

√ Φ<br />

p<br />

(p)<br />

�<br />

0 ,<br />

and<br />

˜P (∞)<br />

∞ (V (∞)<br />

k<br />

,t)=| ˜ Ψ (∞)<br />

By Theorem 8.3 and the def<strong>in</strong>ition of ˜ Ψ (∞)<br />

∞�<br />

˜P (∞)<br />

k=0<br />

∞ (V (∞)<br />

k<br />

,t)=<br />

∞ (V (∞)<br />

k<br />

∞ (V (∞)<br />

k<br />

∞�<br />

k=1<br />

,t)| 2 .<br />

,t), we see that<br />

k 2 J 2 k (2t)<br />

=1. (8.4.3)<br />

t2 The second equality comes from an expansion of z2 as a series of squares of<br />

Bessel coefficients (see page 37 <strong>in</strong> Watson (1944)):<br />

z 2 ∞�<br />

=4 k 2 J 2 k(z).<br />

k=1


Quantum Walks 419<br />

Not<strong>in</strong>g the result (8.4.3), here we def<strong>in</strong>e another cont<strong>in</strong>uous-time quantum<br />

walk Y (t) start<strong>in</strong>g from the root def<strong>in</strong>ed by<br />

Then we obta<strong>in</strong><br />

P (Y (t) =k) = ˜ P (∞)<br />

Theorem 8.4. As t →∞,<br />

∞ (V (∞)<br />

k ,t)=(k +1) 2 J 2 k+1 (2t)<br />

Y (t)<br />

t<br />

⇒ Z∗,<br />

where Z∗ has the follow<strong>in</strong>g density function:<br />

x2 π √ 4 − x2 I (0,2)(x) (x∈R). Proof. From Theorem 8.3, we beg<strong>in</strong> with comput<strong>in</strong>g<br />

� � ��<br />

Y (t)<br />

E exp iξ =<br />

t<br />

exp(−iξ/t)<br />

t2 ∞�<br />

k=1<br />

t 2<br />

�<br />

exp iξ k<br />

�<br />

k<br />

t<br />

2 J 2 k+1 (2t),<br />

for ξ ∈ R. By Neumann’s addition theorem (see p.358 <strong>in</strong> Watson (1944)), we<br />

have<br />

J0( � a 2 + b 2 − 2ab cos(ξ)) =<br />

Tak<strong>in</strong>g t = a = b <strong>in</strong> this equation gives<br />

J0(4t � s<strong>in</strong>(ξ/2)) =<br />

∞�<br />

k=−∞<br />

∞�<br />

k=−∞<br />

Jk(a)Jk(b)exp(ikξ).<br />

J 2 k(t)exp(ikξ).<br />

By differentiat<strong>in</strong>g both sides of the equation twice with respect to t, wesee<br />

= t<br />

4 s<strong>in</strong><br />

∞�<br />

k=1<br />

� ξ<br />

2<br />

k 2 J 2 1<br />

k (t)exp(ikξ) =<br />

�<br />

J ′ 0<br />

Therefore we obta<strong>in</strong><br />

� � ��<br />

Y (t)<br />

E exp iξ<br />

t<br />

�<br />

2t s<strong>in</strong><br />

2<br />

∞�<br />

k=−∞<br />

� ��<br />

ξ<br />

−<br />

2<br />

t2<br />

2 cos2<br />

�<br />

=exp − iξ<br />

��<br />

1<br />

t 2t s<strong>in</strong><br />

�<br />

ξ<br />

2t<br />

−2cos 2<br />

k 2 J 2 k (t)exp(ikξ)<br />

� �<br />

ξ<br />

J<br />

2<br />

′′<br />

�<br />

0 2t s<strong>in</strong><br />

�<br />

� ξ<br />

2t<br />

.<br />

� ��<br />

ξ<br />

.<br />

2<br />

J ′ � � ��<br />

ξ<br />

0 4t s<strong>in</strong><br />

2t<br />

�<br />

J ′′<br />

� �<br />

ξ<br />

0 4t s<strong>in</strong><br />

2t<br />

���<br />

.


420 N. Konno<br />

Then a similar argument as <strong>in</strong> Konno (2005c) yields<br />

lim<br />

t→∞ E<br />

� � ��<br />

Y (t)<br />

exp iξ<br />

t<br />

On the other hand, (8.2.2) with ν =0gives<br />

J ′′<br />

� 1<br />

0 (2ξ) =−<br />

−1<br />

0<br />

= −2J ′′<br />

0 (2ξ).<br />

exp (2iξx)<br />

x2 π √ dx.<br />

1 − x2 From the last two equations, we conclude that<br />

lim<br />

t→∞ E<br />

� � �� � 2<br />

Y (t)<br />

exp iξ = exp (iξx)<br />

t<br />

x2 π √ dx.<br />

4 − x2 It is <strong>in</strong>terest<strong>in</strong>g to remark that when p = 2 case, i.e., Z1 , a similar type of<br />

density function can be derived from Theorem 7.4:<br />

2<br />

π √ 1 − x2 I (0,1)(x) (x∈R). 9 Ultrametric Space<br />

9.1 Introduction<br />

Cont<strong>in</strong>uous-time classical random walks on ultrametric spaces have been<br />

<strong>in</strong>vestigated by some groups to describe the relaxation processes <strong>in</strong> complex<br />

systems, such as glasses, clusters and prote<strong>in</strong>s. On the other hand, a<br />

cont<strong>in</strong>uous-time quantum walk has been widely studied by various researchers<br />

for some important graphs, such as cycle graph, l<strong>in</strong>e, hypercube and complete<br />

graph, with the hope of f<strong>in</strong>d<strong>in</strong>g a new quantum algorithmic technique. However<br />

no result is known for a quantum walk on the ultrametric space. Results<br />

here appeared <strong>in</strong> Konno (2006b).<br />

Let X be a set and let ρ be a metric on X. The pair (X, ρ) is called a<br />

metric space. Moreover if ρ satisfies the strong triangle <strong>in</strong>equality:<br />

ρ(x, z) ≤ max(ρ(x, y),ρ(y, z)),<br />

for any x, y, z ∈ X, thenρ is said to be an ultrametric. A set endowed with<br />

an ultrametric is called an ultrametric space. Let p be a prime number and<br />

Qp be the field of p-adic numbers. Zp denotes the set of p-adic <strong>in</strong>tegers. Then<br />

Zp is a subr<strong>in</strong>g of Qp and Zp = {x ∈ Qp : |x|p ≤ 1}, where |·|p is the p-adic<br />

absolute value. Remark that the metric ρp(x, y) =|x − y|p is an ultrametric.<br />

Every x ∈ Zp can be expanded <strong>in</strong> the follow<strong>in</strong>g way:<br />

x = y0 + y1p + y2p 2 + ···+ ynp n + ··· ,


Quantum Walks 421<br />

where yn ∈{0, 1,...,p− 1} for any n. See Khrennikov and Nilsson (2004)<br />

for more details. It is important that Zp can be represented by the bottom<br />

<strong>in</strong>f<strong>in</strong>ite regular Cayley tree of degree p, Tp, <strong>in</strong> which every branch at each<br />

level splits <strong>in</strong>to p other branches. So we consider a cont<strong>in</strong>uous-time quantum<br />

walk on the bottom of Tp at level (or depth) M, denotedbyT (p)<br />

M .<br />

We calculate the probability distribution of the quantum walk on T (p)<br />

M .<br />

From this, we obta<strong>in</strong> the time-averaged probability distribution on T (p)<br />

M and<br />

then take a limit as M →∞. On the other hand, we can first take the limit<br />

as M →∞and then derive the time-averaged probability distribution. It<br />

is shown that the first limit result co<strong>in</strong>cides with the second one, that is,<br />

time-averaged operation and M-limit one are commutative. By us<strong>in</strong>g these<br />

results, we clarify a difference between classical and quantum cases. A strik<strong>in</strong>g<br />

difference is that localization occurs at any site for a wide class of quantum<br />

walks. Furthermore we compare the results of ultrametric space with those<br />

of other graphs, such as cyclic graph, l<strong>in</strong>e, hypercube and complete graph <strong>in</strong><br />

the quantum case. Some applications of p-adic analysis <strong>in</strong> physics and biology<br />

were reported <strong>in</strong> Khrennikov and Nilsson (2004). Social network models<br />

based on the ultrametric distance were present ed and studied (see Watts,<br />

Dodds, and Newman (2002), Dodds, Watts, and Sabel (2003)) The disease<br />

spread<strong>in</strong>g on a hierarchical metapopulation model was <strong>in</strong>vestigated <strong>in</strong> Watts<br />

et al. (2005). Our quantum walk on the ultrametric space may be useful<br />

for design<strong>in</strong>g a new quantum search algorithm on graphs with a hierarchical<br />

structure.<br />

9.2 Def<strong>in</strong>ition<br />

For M ≥ 0andx ∈ Zp, the closed p-adic ball B (p)<br />

M (x) ofradiusp−M with<br />

center x is def<strong>in</strong>ed by<br />

B (p)<br />

M (x) ={y ∈ Zp : ρp(x, y) ≤ p −M }.<br />

Each ball B (p)<br />

M (x)ofradiusp−M can be represented as a f<strong>in</strong>ite union of disjo<strong>in</strong>t<br />

balls B (p)<br />

M+1 (xm) ofradiusp −(M+1) :<br />

B (p)<br />

p−1 �<br />

M (x) = B<br />

m=0<br />

(p)<br />

M+1 (xm),<br />

for suitable x0,x1,...,xp−1 ∈ Zp, so we have<br />

Zp =<br />

p M �−1<br />

m=0<br />

B (p)<br />

M (m).


422 N. Konno<br />

When p =3,<br />

and<br />

Zp =<br />

B (3)<br />

1<br />

B (3)<br />

1<br />

B (3)<br />

1<br />

3 1 �−1<br />

m=0<br />

(0) = B(3)<br />

2<br />

(1) = B(3)<br />

2<br />

(2) = B(3)<br />

2<br />

B (3)<br />

3<br />

1 (m) =<br />

2 �−1<br />

k=0<br />

B (3)<br />

2 (k),<br />

(0) ∪ B(3) 2 (3) ∪ B(3) 2 (6),<br />

(1) ∪ B(3) 2 (4) ∪ B(3) 2 (7),<br />

(2) ∪ B(3) 2 (5) ∪ B(3) 2 (8).<br />

We def<strong>in</strong>e the distance between two balls B1 and B2 as<br />

ρp(B1,B2) = <strong>in</strong>f{ρp(x, y) :x ∈ B1,y ∈ B2}.<br />

We observe that end po<strong>in</strong>ts of a regular Cayley tree with p degree at level<br />

M may be represented as a set of disconnected balls<br />

{B (p)<br />

M (0),B(p)<br />

M (1),...,B(p)<br />

M (pM − 1)}<br />

with radius p −M cover<strong>in</strong>g Zp. So we can consider p M balls<br />

{B (p)<br />

M (0),B(p)<br />

M (1),...,B(p)<br />

M (pM − 1)}<br />

as p M po<strong>in</strong>ts {0, 1,...,p M − 1}, denotedbyT (p)<br />

M .<br />

As <strong>in</strong> the case of classical random walk, let ψ (p)<br />

M<br />

(i, j) be the amplitude of<br />

(i) to the ball B(p)<br />

M (j) separated by p-adic distance<br />

p−(M−k) . Here we put ɛk = ɛ (p)<br />

M (k) =ψ(p)<br />

M (i, j). Remark that ψ(p)<br />

M (i, j) =<br />

ψ (p)<br />

M (j, i).<br />

We consider p =3andM =2.Inthiscase,ɛkis given by<br />

a jump from the ball B (p)<br />

M<br />

ɛ1 = ψ (3)<br />

2<br />

= ψ (3)<br />

2<br />

ɛ2 = ψ (3)<br />

2<br />

= ψ (3)<br />

2<br />

······ .<br />

(0, 3) = ψ(3)<br />

2<br />

(6, 0) = ψ(3)<br />

2<br />

(0, 1) = ψ(3)<br />

2<br />

(3, 1) = ψ(3)<br />

2<br />

(0, 6) = ψ(3)<br />

2<br />

(6, 3) = ψ(3)<br />

2<br />

(3, 0) = ψ(3) 2 (3, 6)<br />

(0, 4) = ···= ψ(3) 2 (0, 8)<br />

(1, 4) = ψ(3) 2 (1, 7) = ··· ,<br />

(3, 4) = ···= ψ(3) 2 (3, 8) = ··· ,<br />

Let In and Jn denote the n×n identity matrix and the all-one n×n matrix.<br />

We def<strong>in</strong>e a pM × pM symmetric matrix H (p) (p)<br />

M on T M which is treated as the<br />

Hamiltonian of the quantum system as follows: for any M =1, 2,...,


Quantum Walks 423<br />

H (p)<br />

M+1 = Ip ⊗ H (p)<br />

M +(Jp − Ip) ⊗ ɛM+1IpM ,<br />

H (p)<br />

1 = ɛ0Ip + ɛ1(Jp − Ip),<br />

where ɛj ∈ R (j =0, 1, 2,...)andRis the set of real numbers.<br />

The evolution of the cont<strong>in</strong>uous-time quantum walk on T (p)<br />

M is governed<br />

by the follow<strong>in</strong>g unitary matrix:<br />

U (p)<br />

M (t) =eitH(p) M .<br />

The amplitude wave function at time t, Ψ (p)<br />

M (t), is def<strong>in</strong>ed by<br />

Ψ (p) (p) (p)<br />

M (t) =U M (t)Ψ M (0).<br />

In this chapter we take Ψ (p)<br />

M (0) = T [1, 0, 0,...,0] as <strong>in</strong>itial state.<br />

The (n + 1)-th coord<strong>in</strong>ate of Ψ (p)<br />

(p)<br />

M (t) isdenotedbyΨ M (n, t) whichisthe<br />

amplitude wave function at site n at time t for n =0, 1,...,pM − 1. The<br />

is given by<br />

probability f<strong>in</strong>d<strong>in</strong>g the walker is at site n at time t on T (p)<br />

M<br />

9.3 Results<br />

P (p)<br />

(p)<br />

M (n, t) =|Ψ M (n, t)|2 .<br />

Let {ηm : m =0, 1,...,M} be def<strong>in</strong>ed by<br />

⎧<br />

⎨ ɛ0 − ɛ1<br />

if m =0,<br />

ηm = ɛ0 +(p− 1)<br />

⎩<br />

�m k=1 pk−1ɛk − pmɛm+1 if m =1, 2,...,M − 1,<br />

0 if m = M.<br />

Direct computation yields that {ηm : m =0, 1,...,M} is the set of eigen-<br />

values of H (p)<br />

M and eigenvectors for ηm are given <strong>in</strong> the follow<strong>in</strong>g way, see<br />

Ogielski and Ste<strong>in</strong> (1985) for p = 2 case. Let ω = ωp =exp(2πi/p). Put<br />

0n =[<br />

The p M−1 (p − 1) eigenvectors for η0 are<br />

n<br />

� �� � � �� �<br />

0, 0,...,0], 1n =[ 1, 1,...,1].<br />

u0(1) = [1,ω,ω 2 ,...,ω p−1 , 0pM −p] T ,<br />

u0(2) = [1,ω 2 ,ω 4 ,...,ω 2(p−1) , 0pM −p] T ,<br />

...<br />

u0(p − 1) = [1,ω (p−1) ,ω 2(p−1) ,...,ω (p−1)2<br />

, 0pM −p] T ,<br />

u0(p) = [0p, 1,ω,ω 2 ,...,ω p−1 , 0pM −2p] T ,<br />

n


424 N. Konno<br />

u0(p +1) = [0p, 1,ω 2 ,ω 4 ,...,ω 2(p−1) , 0pM −2p] T ,<br />

...<br />

u0(2(p − 1)) = [0p, 1,ω (p−1) ,ω 2(p−1) ,...,ω (p−1)2<br />

, 0pM −2p] T ,<br />

...<br />

u0(p M−1 (p − 1)) = [0pM −p, 1,ω (p−1) ,ω 2(p−1) ,...,ω (p−1)2<br />

] T .<br />

The p M−2 (p − 1) eigenvectors for η1 are<br />

u1(1) = [1p,ω1p,ω 2 1p,...,ω p−1 1p, 0pM −p2] T ,<br />

u1(2) = [1p,ω 2 1p,ω 4 1p,...,ω 2(p−1) 1p, 0pM −p2] T ,<br />

...<br />

u1(p − 1) = [1p,ω (p−1) 1p,ω 2(p−1) 1p,...,ω (p−1)2<br />

1p, 0pM −p2] T ,<br />

u1(p) = [0p2, 1p,ω1p,ω 2 1p,...,ω p−1 1p, 0pM −2p2] T ,<br />

u1(p +1) = [0p2, 1p,ω 2 1p,ω 4 1p,...,ω 2(p−1) 1p, 0pM −2p2] T ,<br />

...<br />

u1(2(p − 1)) = [0p2, 1p,ω (p−1) 1p,ω 2(p−1) 1p,...,ω (p−1)2<br />

1p, 0pM −p2] T ,<br />

...<br />

u1(p M−2 (p − 1)) = [0pM −p2, 1p,ω (p−1) 1p,ω 2(p−1) 1p,...,ω (p−1)2<br />

1p] T .<br />

We have p M−(k+1) (p − 1) eigenvectors for ηk (k =2, 3,...,M − 1) similarly.<br />

So the p − 1 eigenvectors for ηM−1 are<br />

uM−1(1) = [1pM−1,ω1pM−1,ω 2 1pM−1,...,ω p−1 1pM−1] T ,<br />

uM−1(2) = [1pM−1,ω 2 1pM−1,ω 4 1pM−1,...,ω 2(p−1) 1pM−1] T ,<br />

...<br />

uM−1(p − 1) = [1pM−1,ω (p−1) 1pM−1,ω 2(p−1) 1pM−1,...,ω (p−1)2<br />

1pM−1] T .<br />

F<strong>in</strong>ally the only eigenvector for ηM is [1 p M ] T .<br />

Throughout this chapter, we assume that<br />

0


Quantum Walks 425<br />

as <strong>in</strong> the classical case. We should note that ɛ0 < 0


426 N. Konno<br />

P (3)<br />

⎧<br />

{41 + 24 cos(t(η0 − η1)) + 12 cos(tη0)+4cos(tη1)} /3<br />

⎪⎨<br />

2 (n, t) =<br />

⎪⎩<br />

4<br />

if n =0,<br />

{14 − 12 cos(t(η0 − η1)) − 6cos(tη0)+4cos(tη1)} /34 if n =1, 2,<br />

2 {1 − cos(tη1)} /34 if n =3, 4,...,8.<br />

In general, P (p)<br />

M (n, t) does not converge as t →∞for any fixed n. Sowe<br />

<strong>in</strong>troduce the time-averaged distribution of P (p)<br />

M (n, t) as follows:<br />

¯P (p) 1<br />

M (n) = lim<br />

t→∞ t<br />

� t<br />

0<br />

P (p)<br />

M (n, s)ds, (9.3.1)<br />

if the right-hand side of (9.3.1) exists. Then Proposition 9.2 gives<br />

Theorem 9.3.<br />

¯P (p)<br />

⎧<br />

p − 1<br />

⎪⎨<br />

p +1<br />

M (n) =<br />

⎪⎩<br />

+<br />

2<br />

(p +1)p2M if n ∈ V (p)<br />

0 ,<br />

�<br />

2 1 1<br />

+<br />

p +1 p2k−1 p2M �<br />

if n ∈ V (p)<br />

k (k =1, 2,...,M − 1),<br />

2<br />

p2M if n ∈ V (p)<br />

M .<br />

It is <strong>in</strong>terest<strong>in</strong>g to note that ¯ P (p)<br />

M (n) does not depend on {ɛk : k =<br />

0, 1,...,M}. In the case of p =3andM =2,wehave<br />

¯P (3)<br />

2 (0) = 41<br />

3 4 , ¯ P (3)<br />

2 (1) = ¯ P (3)<br />

2 (2) = 14<br />

3 4 , ¯ P (3)<br />

2 (3) = ···= ¯ P (3)<br />

The follow<strong>in</strong>g result is immediate from Theorem 9.3.<br />

Corollary 9.4.<br />

lim<br />

M→∞<br />

¯P (p)<br />

M (n) =<br />

⎧<br />

⎪⎨<br />

p − 1<br />

p +1<br />

⎪⎩<br />

2<br />

(p +1)p2k−1 if n ∈ V (p)<br />

0 ,<br />

if n ∈ V (p)<br />

k<br />

2 (8) = 2<br />

.<br />

34 (k =1, 2,...).<br />

Here we consider the mean distance from 0 at time t def<strong>in</strong>ed by<br />

d (p)<br />

M (t) =<br />

M�<br />

k=1<br />

�<br />

n∈V (p)<br />

k<br />

p −(M−k) P (p)<br />

M (n, t).<br />

Then its time-averaged mean distance is given by<br />

M� �<br />

¯d (p)<br />

M =<br />

k=1<br />

n∈V (p)<br />

k<br />

p −(M−k) ¯ P (p)<br />

M (n).


Quantum Walks 427<br />

From Theorem 9.3, we get<br />

¯d (p)<br />

M<br />

This gives<br />

= 2(p − 1)(M − 1)<br />

(p +1)p M<br />

+ 2 �� (p − 1)(p +1) 2 +1 � p2M−2 − 1 �<br />

(p +1) 2p3M−1 .<br />

p<br />

lim<br />

M→∞<br />

M<br />

M ¯ d (p) 2(p − 1)<br />

M =<br />

p +1 .<br />

Next we take the limit as M →∞first. We def<strong>in</strong>e P (p)<br />

∞ (n, t)byP (p)<br />

∞ (n, t) =<br />

limM→∞ P (p)<br />

M (n, t), if the right-hand side of the equation exists. By<br />

Proposition 9.2, we obta<strong>in</strong><br />

Proposition 9.5.<br />

⎧<br />

(p − 1) 2<br />

⎡�<br />

∞�<br />

⎣ p −(m+1) cos(tηm)<br />

P (p)<br />

∞ (n, t) =<br />

⎪⎨<br />

⎪⎩<br />

m=0<br />

if n ∈ V (p)<br />

0 ,<br />

� 2<br />

�<br />

∞�<br />

+ p −(m+1) � ⎤<br />

2<br />

s<strong>in</strong>(tηm) ⎦<br />

m=0<br />

�<br />

−p −k ∞�<br />

cos(tηk−1)+(p− 1) p −(m+1) cos(tηm)<br />

� 2<br />

�<br />

m=k<br />

+ −p −k ∞�<br />

s<strong>in</strong>(tηk−1)+(p− 1) p −(m+1) s<strong>in</strong>(tηm)<br />

if n ∈ V (p)<br />

k<br />

(k =1, 2,...).<br />

m=k<br />

In a similar way, the time-averaged distribution P (p)<br />

∞ (n, t) isgivenby<br />

¯P (p) 1<br />

∞ (n) = lim<br />

t→∞ t<br />

� t<br />

0<br />

� 2<br />

P (p)<br />

∞ (n, s)ds, (9.3.2)<br />

if the right-hand side of (9.3.2) exists. Comb<strong>in</strong><strong>in</strong>g Theorem 9.3 with<br />

Proposition 9.5 yields<br />

Corollary 9.6. For any n =0, 1, 2,..., we have<br />

Furthermore,<br />

¯P (p)<br />

∞ (n) = lim<br />

M→∞<br />

¯P (p)<br />

M (n) > 0.<br />

lim ¯P<br />

p→∞<br />

(p)<br />

∞ (n) =δ0(n),<br />

where δm(n) =1, if n = m, =0, if n �= m.


428 N. Konno<br />

In general, we say that localization occurs at site n if the time-averaged<br />

probability at the site is positive. Therefore the localization occurs at any<br />

site for both any f<strong>in</strong>ite M and M →∞limit cases.<br />

9.4 <strong>Classical</strong> Case<br />

In this section we review three classical examples and clarify a difference<br />

between classical and quantum walks. Let P (p)<br />

c (n, t) be the probability that<br />

a classical random walker start<strong>in</strong>g from 0 is located at site n at time t on Zp.<br />

has the form<br />

In the case of a l<strong>in</strong>ear landscape, the transition rate on T (p)<br />

M<br />

ɛk = w0 p −(1+α)(k−M) ,<br />

for w0 > 0andα>0. Then Avetisov et al. (2002) showed a power decay law<br />

<strong>in</strong> M →∞limit:<br />

P (p)<br />

c (0,t) ∼ 1<br />

t 1/α<br />

(t →∞),<br />

where f(t) ∼ g(t)(t →∞) means there exist positive constants C1 and C2<br />

such that<br />

C1 ≤ lim <strong>in</strong>f<br />

t→∞<br />

f(t) f(t)<br />

≤ lim sup ≤ C2.<br />

g(t) t→∞ g(t)<br />

For a logarithmic landscape case, the transition rate on T (p)<br />

M is<br />

ɛk = w0 p −(k−M)<br />

1<br />

(log(1 + p−(k−M) ,<br />

)) α<br />

for w0 > 0 and α > 1. The follow<strong>in</strong>g stretched exponential decay law<br />

(the Kohlrausch-Williams-Watts law) was proved by Avetisov, Bikulov, and<br />

Osipov (2003) <strong>in</strong> M →∞limit:<br />

log(P (p)<br />

c (0,t)) ∼−t 1/α<br />

(t →∞).<br />

In the case of an exponential landscape, the transition rate on T (p)<br />

M has<br />

the form<br />

ɛk = w0 p −(k−M) exp(−αp (k−M) ),<br />

for w0 > 0andα>0. Then Avetisov, Bikulov, and Osipov (2003) obta<strong>in</strong>ed<br />

a logarithmic decay law tak<strong>in</strong>g a limit as M →∞:<br />

P (p)<br />

c (0,t) ∼ 1<br />

log t<br />

(t →∞).


Quantum Walks 429<br />

The above three facts imply that the localization does not occur at position<br />

0. In contrast to the classical case, the localization occurs at any position for<br />

our quantum case.<br />

9.5 Quantum Case for Other Graphs<br />

We consider the time-averaged probability distribution for cont<strong>in</strong>uous-time<br />

quantum walk start<strong>in</strong>g from a site on other graphs, such as cycle graph, l<strong>in</strong>e,<br />

hypercube and complete graph. Then the Hamiltonian of the walk is given<br />

by the adjacency matrix of the graph.<br />

In the case of a cycle graph CN with N sites, we have obta<strong>in</strong>ed the follow<strong>in</strong>g<br />

result (see Chapter 10 or Inui, Kasahara, Konishi, and Konno (2005)):<br />

¯PN (n) = 1<br />

N<br />

for any n =0, 1,...,N − 1, where<br />

⎧<br />

2RN(n)<br />

+<br />

N 2 ,<br />

⎨ −1/2 if N =odd, ξ2n�= 0 (mod2π),<br />

RN (n) = −1<br />

⎩ ¯N<br />

if N =even, ξ2n�= 0<br />

if ξ2n =0 (mod2π).<br />

(mod2π),<br />

Here ξj =2πj/N, ¯ N =[(N − 1)/2], and [x] is the smallest <strong>in</strong>teger greater<br />

than x. WhenN = odd (i.e., ¯ N =(N − 1)/2),<br />

¯PN =<br />

�<br />

1 N − 1<br />

+<br />

N N 2 ,<br />

N−1<br />

� �� �<br />

1<br />

N<br />

when N = even (i.e., ¯ N =(N − 2)/2),<br />

¯PN =<br />

� 1<br />

N<br />

Then we have<br />

for any n.<br />

+ N − 2<br />

N 2 ,<br />

1 1<br />

− ,...,<br />

N 2 N<br />

(N−2)/2<br />

� �� �<br />

1<br />

N<br />

2 1<br />

− ,...,<br />

N 2 N<br />

2<br />

− ,<br />

N 2<br />

1 N − 2<br />

+<br />

N N 2 ,<br />

lim ¯PN (n) =0,<br />

N→∞<br />

1<br />

−<br />

N 2<br />

�<br />

,<br />

(N−2)/2<br />

� �� �<br />

1<br />

N<br />

2 1<br />

− ,...,<br />

N 2 N<br />

2<br />

−<br />

N 2<br />

�<br />

.


430 N. Konno<br />

In Z case, the probability distribution P (n, t) of the walk start<strong>in</strong>g from<br />

the orig<strong>in</strong> at location n and time t is given by<br />

P (n, t) =J 2 n (t),<br />

where Jn(t) is the Bessel function of the first k<strong>in</strong>d of order n, see Chapter 7,<br />

for example. The asymptotic behavior of Jn(t) at <strong>in</strong>f<strong>in</strong>ity is as follows:<br />

� �<br />

2<br />

Jn(t) = cos(t − θ(n)) − s<strong>in</strong>(t − θ(n))<br />

πt<br />

4n2 − 1<br />

8t<br />

+ O(t−2 �<br />

) , t →∞,<br />

where θ(n) =(2n+1)π/4, see page 195 <strong>in</strong> Watson (1944). From this fact and<br />

|Jn(t)| ≤1 for any t and n (see page 31 <strong>in</strong> Watson (1944)), we obta<strong>in</strong><br />

¯P<br />

1<br />

(n) = lim<br />

t→∞ t<br />

� t<br />

0<br />

J 2 n(s) ds =0,<br />

for any n ∈ Z.<br />

Next we consider the quantum walk start<strong>in</strong>g from a site, denoted by 0, on<br />

a hypercube with 2 N sites, WN . Then the probability f<strong>in</strong>d<strong>in</strong>g the walker at<br />

site n at time t is<br />

PN (n, t) =cos(t/N) 2(N−k) s<strong>in</strong>(t/N) 2k<br />

if n ∈ Vk (0 ≤ k ≤ N),<br />

where Vk = {x ∈ WN : ||x|| = k} and ||x|| is the path length from 0 to x <strong>in</strong><br />

WN . A derivation is shown <strong>in</strong> Jafarizadeh and Salimi (2007). From the result<br />

we see that<br />

¯PN (n) = 1<br />

2 2N<br />

� N<br />

k<br />

� −1� 2k<br />

k<br />

�� 2(N − k)<br />

N − k<br />

if n ∈ Vk (0 ≤ k ≤ N). For example, when N =4,<br />

⎧<br />

⎨ 35/128 if n ∈ V0 ∪ V4,<br />

¯P4(n) = 5/128 if n ∈ V1 ∪ V3,<br />

⎩<br />

3/128 if n ∈ V2.<br />

For general N, wehave<br />

lim ¯PN (n) =0,<br />

N→∞<br />

for any n =0, 1,.... Moreover if we let<br />

¯PN (Vk) = �<br />

¯PN (n),<br />

for k =0, 1,...,N, then we get<br />

n∈Vk<br />

�<br />

,


Quantum Walks 431<br />

¯PN (Vk) = 1<br />

2 2N<br />

� 2k<br />

k<br />

�� 2(N − k)<br />

N − k<br />

s<strong>in</strong>ce |Vk| = N!/(N − k)!k!. It is <strong>in</strong>terest<strong>in</strong>g to note that this probability<br />

corresponds to the well-known arcs<strong>in</strong>e law for the classical random walk.<br />

Therefore, <strong>in</strong> three cases, CN (cycle graph) and WN (N-cube) as N →∞<br />

and Z, the localization does not occur at any location.<br />

In the case of the complete graph with N sites, KN, the Hamiltonian HN<br />

is def<strong>in</strong>ed by HN = IN −NJN . Then eigenvalues of HN are η0 =0,η1 = η2 =<br />

··· = ηN−1 = −N. The eigenvectors are given by the Vandermonde matrix,<br />

see Ahmadi et al. (2003). Then direct computation yields<br />

⎧<br />

⎪⎨<br />

(N − 1)<br />

PN (n, t) =<br />

⎪⎩<br />

2 +1+2(N− 1) cos(Nt)<br />

N 2<br />

if n =0,<br />

2(1 − cos(Nt))<br />

N 2<br />

if n =1, 2,...,N − 1.<br />

Therefore<br />

⎧<br />

⎪⎨ (N − 1)<br />

¯PN (n) =<br />

⎪⎩<br />

2 +1<br />

N 2<br />

2<br />

N 2<br />

So we have<br />

if n =0,<br />

lim ¯PN (n) =δ0(n),<br />

N→∞<br />

�<br />

,<br />

if n =1, 2,...,N − 1.<br />

for any n =0, 1,.... Therefore the localization occurs only at n =0.This<br />

corresponds to our case for p →∞.<br />

9.6 Conclusion<br />

We have derived the expression of the probability distribution of a cont<strong>in</strong>uous-<br />

time quantum walk on T (p)<br />

M correspond<strong>in</strong>g to Zp <strong>in</strong> the M →∞limit. As a<br />

result, we obta<strong>in</strong>ed<br />

¯P (p)<br />

∞ (n) = lim<br />

M→∞<br />

for a class of ɛk satisfy<strong>in</strong>g<br />

¯P (p)<br />

M (n) > 0 (n =0, 1, 2,...),<br />

ɛ0 < 0


432 N. Konno<br />

any location. For KN (complete graph) as N →∞case, the localization<br />

occurs only at 0. In three typical classical cases, the localization does not<br />

occur even at 0 site. We hope that this property of our quantum walk can be<br />

useful <strong>in</strong> search problems on a tree-like hierarchical structure.<br />

10 Cycle<br />

10.1 Introduction<br />

In this chapter we treat the cont<strong>in</strong>uous-time quantum walk on CN which is<br />

the cycle with N vertices, i.e., CN = {0, 1,...,N− 1}. Results here appeared<br />

<strong>in</strong> Inui, Kasahara, Konishi, and Konno (2005). Let A be the N ×N adjacency<br />

matrix of CN . The cont<strong>in</strong>uous-time quantum walk considered here is given<br />

by the follow<strong>in</strong>g unitary matrix:<br />

U(t) =e −itA/2 .<br />

Follow<strong>in</strong>g the paper of Ahmadi et al. (2003), we take −itA/2 <strong>in</strong>stead of itA/2.<br />

The amplitude wave function at time t, ΨN (t), is def<strong>in</strong>ed by<br />

ΨN(t) =U(t)ΨN (0).<br />

The (n + 1)-th coord<strong>in</strong>ate of ΨN (t) is denoted by ΨN (n, t) whichistheamplitude<br />

wave function at vertex n at time t for n =0, 1,...,N − 1. In this<br />

chapter we take ΨN (0) = T [1, 0, 0,...,0] as <strong>in</strong>itial state. The probability that<br />

the particle is at vertex n at time t, PN (n, t), is given by<br />

PN (n, t) =|ΨN (n, t)| 2 .<br />

Here we give a connection between a cont<strong>in</strong>uum time limit model for<br />

discrete-time quantum walk on Z given by Romanelli et al. (2003) and<br />

cont<strong>in</strong>uous-time quantum walk on circles. Ahmadi et al. (2003) showed<br />

ΨN (n, t) = 1<br />

N−1 �<br />

N<br />

j=0<br />

e −i(t cos ξj −nξj) , (10.1.1)<br />

for any t ≥ 0andn = 0, 1,...,N − 1, where ξj = 2πj/N. As Ahmadi<br />

et al. (2003) po<strong>in</strong>ted out, ΨN (n, t) can be also expressed by us<strong>in</strong>g the Bessel<br />

function, that is,<br />

ΨN (n, t) =<br />

�<br />

k=n (modN)<br />

(−i) k Jk(t) = 1<br />

2<br />

�<br />

k=±n (modN)<br />

(−i) k Jk(t), (10.1.2)


Quantum Walks 433<br />

where Jν(z) is the Bessel function. In fact, the generat<strong>in</strong>g function of the<br />

Bessel function:<br />

� �<br />

t<br />

exp z −<br />

2<br />

1<br />

��<br />

=<br />

z<br />

�<br />

z ν Jν(t),<br />

gives<br />

e −it cos ξj = �<br />

ν∈Z<br />

ν∈Z<br />

(−i) ν e iνξj Jν(t).<br />

By us<strong>in</strong>g (10.1.1), the amplitude wave function is then calculated as<br />

ΨN (n, t) = 1<br />

N<br />

In a similar way, we have<br />

ΨN (n, t) = 1<br />

N<br />

N−1 �<br />

�<br />

j=0 ν∈Z<br />

N−1 �<br />

�<br />

j=0 ν∈Z<br />

(−i) ν e i(n+ν)ξj Jν(t). (10.1.3)<br />

(−i) ν e i(n−ν)ξj Jν(t). (10.1.4)<br />

From (10.1.3) and (10.1.4), we obta<strong>in</strong> (10.1.2). Moreover, (10.1.2) gives<br />

PN (n, t) = � �<br />

π<br />

�<br />

cos (j − k)N JjN+n(t)JkN+n(t)<br />

2<br />

j,k∈Z<br />

= �<br />

JkN+n(t) 2<br />

k∈Z<br />

+ �<br />

j�=k:j,k∈Z<br />

cos<br />

�<br />

π<br />

�<br />

(j − k)N JjN+n(t)JkN+n(t). (10.1.5)<br />

2<br />

The follow<strong>in</strong>g formula is well known (see page 213 <strong>in</strong> Andrews, Askey, and<br />

Roy (1999)):<br />

�<br />

Jk(t) 2 =1, (10.1.6)<br />

k∈Z<br />

for t ≥ 0. That is, {Jk(t) 2 : k ∈ Z} is a probability distribution on Z for any<br />

time t. Then (10.1.6) can be rewritten as<br />

N−1 �<br />

�<br />

n=0 k∈Z<br />

for t ≥ 0. Thus (10.1.5) and (10.1.7) give<br />

JkN+n(t) 2 =1, (10.1.7)


434 N. Konno<br />

N−1 �<br />

�<br />

n=0 j�=k:j,k∈Z<br />

cos<br />

�<br />

π<br />

�<br />

(j − k)N JjN+n(t)JkN+n(t) =0.<br />

2<br />

Romanelli et al. (2003) studied a cont<strong>in</strong>uum time limit for a discrete-time<br />

quantum walk on Z and obta<strong>in</strong>ed the position probability distribution. When<br />

the <strong>in</strong>itial condition is given by ãl(0) = δl,0, ˜ bl(0) ≡ 0 <strong>in</strong> their notation for<br />

the Hadamard walk, the distribution becomes the follow<strong>in</strong>g <strong>in</strong> our notation:<br />

P (n, t) =Jn(t/ √ 2) 2 .<br />

More generally, we consider a discrete-time quantum walk whose co<strong>in</strong> flip<br />

transformation is given by the follow<strong>in</strong>g unitary matrix:<br />

� �<br />

ab<br />

U = .<br />

cd<br />

Note that a = b = c = −d =1/ √ 2 case is equivalent to the Hadamard walk.<br />

In a similar fashion, we have<br />

P (n, t) =Jn(at) 2 . (10.1.8)<br />

In fact, (10.1.6) guarantees �<br />

n∈Z P (n, t) = 1 for any t ≥ 0. From the probability<br />

theoretical po<strong>in</strong>t of view, it is <strong>in</strong>terest<strong>in</strong>g that a cont<strong>in</strong>uum time limit<br />

model for a discrete-time quantum walk on Z gives a simple model hav<strong>in</strong>g<br />

a squared Bessel distribution <strong>in</strong> the above mean<strong>in</strong>g. For large N, (10.1.5)<br />

is similar to (10.1.8) except for the second term of the right-hand side of<br />

(10.1.5).<br />

To <strong>in</strong>vestigate fluctuations for cont<strong>in</strong>uous-time quantum walks, we consider<br />

both <strong>in</strong>stantaneous uniform mix<strong>in</strong>g property (see below) and temporal<br />

standard deviation (see Section 10.4). The former corresponds to a spatial<br />

fluctuation and the latter corresponds a temporal one.<br />

A walk start<strong>in</strong>g from 0 on CN has the <strong>in</strong>stantaneous uniform mix<strong>in</strong>g<br />

property (IUMP) if there exists t>0 such that PN (n, t) =1/N for any<br />

n =0, 1,...,N − 1. To restate this def<strong>in</strong>ition, we <strong>in</strong>troduce the total variation<br />

distance which is the notion of distance between probability distributions<br />

def<strong>in</strong>ed as<br />

||P − Q|| = max |P (A) − Q(A)| =<br />

A⊂CN<br />

1 �<br />

|P (n) − Q(n)|. (10.1.9)<br />

2<br />

n∈CN<br />

By us<strong>in</strong>g this, the IUMP can be rewritten as follows: if there exists t>0<br />

such that ||PN (·,t) − πN || = 0, where πN is the uniform distribution on<br />

CN . As we will mention <strong>in</strong> Section 10.3, cont<strong>in</strong>uous-time quantum walks on<br />

CN for N =3, 4 have the IUMP. On the other hand, Ahmadi et al. (2003)<br />

conjectured that cont<strong>in</strong>uous-time quantum walks on CN for N ≥ 5donot<br />

have the IUMP. Carlson et al. (2006) showed that the conjecture is true for


Quantum Walks 435<br />

N = 5. It is easily checked that the conjecture is true for N = 6 by a direct<br />

computation (see Section 10.3). Remark that the cont<strong>in</strong>uous-time classical<br />

random walk on CN does not have the IUMP for any N ≥ 3.<br />

10.2 <strong>Classical</strong> Case<br />

First we review the classical case on CN , (see Sch<strong>in</strong>azi (1999) and references<br />

<strong>in</strong> his book, for example). In the cont<strong>in</strong>uous-time classical random walk,<br />

PN (n, t) = 1<br />

N−1 �<br />

N<br />

j=0<br />

cos(ξjn)e t(cos ξj−1) ,<br />

for any t ≥ 0andn =0, 1,...,N − 1. In this walk, the distribution for the<br />

random time between two jumps is the exponential distribution. As for the<br />

discrete-time classical random walk,<br />

PN (n, t) = 1<br />

N−1 �<br />

cos(ξjn)(cos ξj)<br />

N<br />

t ,<br />

j=0<br />

for any t =0, 1,...,and n =0, 1,...,N−1. Here a time-averaged distribution<br />

<strong>in</strong> the cont<strong>in</strong>uous-time case is given by<br />

� T<br />

1<br />

¯PN (n) = lim PN (n, t)dt, (10.2.10)<br />

T →∞ T 0<br />

if the right-hand side of (10.2.10) exists. As for the discrete-time case,<br />

T −1<br />

1<br />

¯PN (n) = lim<br />

T →∞ T<br />

t=0<br />

�<br />

PN (n, t), (10.2.11)<br />

if the right-hand side of (10.2.11) exists. Concern<strong>in</strong>g the cont<strong>in</strong>uous-time<br />

classical case, it is easily shown that PN (n, t) → 1/N (t → ∞) for any<br />

n = 0, 1,...,N − 1. Then we have immediately ¯ PN (n) = 1/N for any<br />

n =0, 1,...,N − 1. On the other hand, <strong>in</strong> the discrete-time classical case, if<br />

N is odd, then PN (n, t) → 1/N as t →∞, and, if N is even, then it does not<br />

converge <strong>in</strong> the limit of t →∞. However, an important property is that ¯ PN<br />

is a uniform distribution <strong>in</strong> both the cont<strong>in</strong>uous- and discrete-time classical<br />

cases.


436 N. Konno<br />

10.3 Quantum Case<br />

Now we consider cont<strong>in</strong>uous-time quantum walks. As we will see later, <strong>in</strong><br />

contrast to classical case, ¯ PN does not become uniform distribution for any<br />

N ≥ 3 <strong>in</strong> quantum case. From (10.1.1), we have<br />

where<br />

RN (n, t) =<br />

�<br />

0≤j


Quantum Walks 437<br />

Now we consider the time-averaged distribution for cont<strong>in</strong>uous-time case.<br />

By us<strong>in</strong>g (10.3.12), we have<br />

Theorem 10.1.<br />

¯PN (n) = 1<br />

N<br />

for any n =0, 1,...,N − 1, where<br />

⎧<br />

+ 2RN(n)<br />

N 2 , (10.3.14)<br />

⎨ −1/2 if N = odd, ξ2n�= 0 (mod 2π),<br />

RN (n) = −1<br />

⎩ ¯N<br />

if N = even, ξ2n�= 0<br />

if ξ2n =0 (mod 2π).<br />

(mod 2π),<br />

This result gives<br />

¯PN (0) = 1<br />

N + 2 ¯ N<br />

. (10.3.15)<br />

N 2<br />

As for discrete-time quantum case given by the Hadamard transformation,<br />

Aharonov et al. (2001) and Bednarska et al. (2003) showed that ¯ PN (n) ≡ 1/N<br />

if N is odd or N = 4, and ¯ PN (n) �≡ 1/N ,otherwise.So<strong>in</strong>contrastto<br />

classical cont<strong>in</strong>uous- and discrete-time and quantum discrete-time cases, ¯ PN<br />

<strong>in</strong> quantum cont<strong>in</strong>uous-time case is not uniform distribution on CN for any<br />

N ≥ 3. In fact, when N = odd (i.e., ¯ N =(N − 1)/2),<br />

¯PN =<br />

�<br />

1 N − 1<br />

+<br />

N N 2 ,<br />

N−1<br />

� �� �<br />

1<br />

N<br />

when N = even (i.e., ¯ N =(N − 2)/2),<br />

¯PN =<br />

For examples,<br />

� 1<br />

N<br />

N − 2<br />

+ ,<br />

N 2<br />

¯P3 =<br />

1 1<br />

− ,...,<br />

N 2 N<br />

(N−2)/2<br />

� �� �<br />

1<br />

N<br />

� �<br />

5 2 2<br />

, , ,<br />

9 9 9<br />

2 1<br />

− ,...,<br />

N 2 N<br />

1 N − 2<br />

+<br />

N N 2 ,<br />

¯ P4 =<br />

2<br />

− ,<br />

N 2<br />

1<br />

−<br />

N 2<br />

�<br />

,<br />

(N−2)/2<br />

� �� �<br />

1<br />

N<br />

� 3<br />

8<br />

2 1<br />

− ,...,<br />

N 2 N<br />

, 1<br />

8<br />

�<br />

3 1<br />

, , .<br />

8 8<br />

2<br />

−<br />

N 2<br />

�<br />

.


438 N. Konno<br />

10.4 Temporal Standard Deviation<br />

We consider the follow<strong>in</strong>g temporal standard deviation σN(n) <strong>in</strong> the cont<strong>in</strong>uoustime<br />

case as <strong>in</strong> the discrete-time case:<br />

�<br />

� T<br />

1 �<br />

σN (n) = lim PN (n, t) −<br />

T →∞ T 0<br />

¯ PN (n) �2 dt, (10.4.16)<br />

if the right-hand side of (10.4.16) exists. In the discrete-time case,<br />

�<br />

�<br />

�<br />

σN (n) = lim �<br />

T →∞<br />

1<br />

T� −1<br />

�<br />

PN (n, t) −<br />

T<br />

¯ PN (n) �2 , (10.4.17)<br />

t=0<br />

if the right-hand side of (10.4.17) exists. In both cont<strong>in</strong>uous- and discretetime<br />

classical cases,<br />

σN (n) =0,<br />

for N ≥ 3andn =0, 1,...,N − 1.Thereasonisasfollows.Inthecase<br />

of classical random walk start<strong>in</strong>g from a site for N = odd (i.e., aperiodic<br />

case), a coupl<strong>in</strong>g method implies that there exist a ∈ (0, 1) and C>0(are<br />

<strong>in</strong>dependent of n and t) such that<br />

|PN (n, t) − ¯ PN (n)| ≤Ca t ,<br />

for any n and t (see page 63 of Sch<strong>in</strong>azi (1999), for example). Therefore we<br />

obta<strong>in</strong><br />

1<br />

T<br />

T� −1<br />

�<br />

PN (n, t) − ¯ PN (n) �2 C<br />

≤ 2<br />

t=0<br />

T<br />

1 − a2T .<br />

1 − a2 The above <strong>in</strong>equality implies that σN (n) =0(n =0, 1,...,N − 1). As for<br />

N = even (i.e., periodic) case, we have the same conclusion σN (n) = 0 for any<br />

n by us<strong>in</strong>g a little modified argument. The same conclusion can be extended<br />

to the cont<strong>in</strong>uous-time classical case <strong>in</strong> a standard fashion.<br />

For the discrete-time quantum case, Theorem 5.3 gives<br />

σ 2 1<br />

N (n) =<br />

N 4<br />

� 2 � S 2 + (n)+S 2 − (n)� +11S 2 0 +10S0S1 +3S 2 1 − S2(n) � − 2<br />

for any n =0, 1,...,N − 1, where N(≥ 3) is odd and<br />

,<br />

N 3<br />

(10.4.18)


Quantum Walks 439<br />

S0 =<br />

S+(n) =<br />

S−(n) =<br />

S2(n) =<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=0<br />

N−1 �<br />

j=1<br />

1<br />

, S1 =<br />

3+cosθj<br />

with θj = ξ2j =4πj/N. For example,<br />

N−1 �<br />

j=0<br />

cos ((n − 1)θj)+cos(nθj)<br />

,<br />

3+cosθj<br />

cos ((n − 1)θj) − cos (nθj)<br />

,<br />

3+cosθj<br />

cos θj<br />

,<br />

3+cosθj<br />

7+cos(2θj)+8cosθjcos2 ((n − 1/2) θj)<br />

(3 + cos θj) 2<br />

,<br />

σ3(0) = σ3(1) = 2√46 45 , σ3(2) = 2<br />

Furthermore from Proposition 5.4 we have<br />

.<br />

9<br />

(10.4.19)<br />

� √ � �<br />

13 − 8 2 1<br />

σN (0) =<br />

+ o ,<br />

N<br />

N<br />

(10.4.20)<br />

as N →∞. The above result implies that the temporal standard deviation<br />

σN (0) <strong>in</strong> the discrete-time case decays <strong>in</strong> the form 1/N as N <strong>in</strong>creases.<br />

Now we consider the cont<strong>in</strong>uous-time quantum case. For example, direct<br />

computation gives<br />

σ3(0) = 2√2 9 , σ3(1)<br />

√<br />

2<br />

= σ3(2) =<br />

9 ,<br />

√<br />

34<br />

σ4(0) = σ4(2) =<br />

16 , σ4(1)<br />

√<br />

2<br />

= σ4(3) =<br />

16 ,<br />

σ5(0) = 4√3 25 , σ5(1) = σ5(2) = σ5(3) = σ5(4) = 2√2 25 ,<br />

σ6(0) = σ6(3) = 7√2 36 , σ6(1)<br />

√<br />

5<br />

= σ6(2) = σ5(4) = σ5(5) =<br />

18 .<br />

From now on we focus on general N = odd case to obta<strong>in</strong> results correspond<strong>in</strong>g<br />

to the discrete-time case (i.e., (10.4.18) and (10.4.20)). The def<strong>in</strong>ition<br />

of σN (n) implies<br />

σ 2 N (n) = 4<br />

× lim<br />

N 4 T →∞<br />

= 4<br />

×<br />

N 4<br />

�<br />

lim<br />

T →∞<br />

� T<br />

1<br />

T<br />

1<br />

T<br />

(RN (n, t) − RN (n)) 2 dt<br />

�<br />

.<br />

0<br />

� T<br />

RN (n, t)<br />

0<br />

2 dt − RN (n) 2


440 N. Konno<br />

For N =oddcase,wehave<br />

�<br />

(N − 1)/2 n =0,<br />

RN (n) =<br />

−1/2 n =1, 2,...,N − 1.<br />

From (10.3.13), we have<br />

� T<br />

1<br />

lim R<br />

T →∞ T 0<br />

2 N (n, t)dt<br />

⎡<br />

� T<br />

1<br />

= lim ⎣<br />

T →∞ T 0<br />

�<br />

⎡<br />

× ⎣<br />

�<br />

= lim<br />

T →∞<br />

� T<br />

1<br />

2T 0<br />

0≤j1


Quantum Walks 441<br />

Theorem 10.2. When N(≥ 3) is odd,<br />

σN (0) = 2√N 2 − 3N +2<br />

N 2 ,<br />

σN (n) =<br />

√ N 2 − 5N +8<br />

N 2 (n =1,...,N − 1). (10.4.21)<br />

In particular, σN (0) >σN (n) for any n =1,...,N − 1.<br />

We should remark that there is a difference between the cont<strong>in</strong>uous-time<br />

case and the discrete-time one, s<strong>in</strong>ce, for example, (10.4.19) gives σ3(0) =<br />

σ3(1) >σ3(2) <strong>in</strong> the discrete-time case. Furthermore (10.4.21) implies<br />

Corollary 10.3.<br />

as N →∞.<br />

σN (0) = 2<br />

N<br />

σN (n) = 1<br />

+ o<br />

N<br />

�<br />

1<br />

+ o<br />

�<br />

,<br />

N<br />

� �<br />

1<br />

N<br />

(n =1,...,N − 1), (10.4.22)<br />

The above result implies that the temporal fluctuation σN (0) <strong>in</strong> the<br />

cont<strong>in</strong>uous-time case decays <strong>in</strong> the form 1/N as N <strong>in</strong>creases as <strong>in</strong> the discretetime<br />

case. However, from � 13 − 8 √ 2 < 2 (see (10.4.20) and (10.4.22)), it<br />

follows that the temporal standard deviation of the cont<strong>in</strong>uous-time case at<br />

site 0 is greater than that of the discrete-time case for large N.<br />

References<br />

Abal, G., Donangelo, R., and Fort, H. (2006). Conditional quantum walk and iterated<br />

quantum games, <strong>in</strong> Annais do 1st Workshop-Escola de Computacao e Informacao<br />

Cuántica, WECIQ06, PPGINF-UCPel, 2006, quant-ph/0607143.<br />

Abal, G., Donangelo, R., and Fort, H. (2007). Asymptotic entanglement <strong>in</strong> the discrete-time<br />

quantum walk, <strong>in</strong> Annals of the 1st Workshop on Quantum Computation and Information,<br />

pp. 189-200, UCPel, 9-11 october 2006, Pelotas, RS, Brazil, arXiv:0709.3279.<br />

Abal, G., Donangelo, R., Romanelli, A., and Siri, R. (2006). Effects of non-local <strong>in</strong>itial<br />

conditions <strong>in</strong> the quantum walk on the l<strong>in</strong>e, Physica A, 371, 1–4, quant-ph/0602188.<br />

Abal, G., Donangelo, R., and Siri, R. (2007). Decoherent quantum walks driven by a generic<br />

co<strong>in</strong> operation, arXiv:0708.1297.<br />

Abal, G., Siri, R., Romanelli, A., and Donangelo, R. (2006). Quantum walk on the l<strong>in</strong>e:<br />

Entanglement and non-local <strong>in</strong>itial conditions, Phys. Rev. A, 73, 042302, quant-ph/<br />

0507264. Erratum, Phys. Rev. A, 73, 069905 (2006).<br />

Accardi, L., and Bo˙zejko, M. (1998). Interact<strong>in</strong>g Fock spaces and Gaussianization of probability<br />

measures, Inf<strong>in</strong>ite Dimensional Analysis, Quantum <strong>Probability</strong> and Related<br />

Topics, 1, 663–670.<br />

Acevedo, O. L., and Gobron, T. (2006). Quantum walks on Cayley graphs, J. Phys. A:<br />

Math. Gen., 39, 585–599, quant-ph/0503078.


442 N. Konno<br />

Acevedo, O. L., Roland, J., and Cerf, N. J. (2006). Explor<strong>in</strong>g scalar quantum walks on<br />

Cayley graphs, quant-ph/0609234.<br />

Adamczak, W., Andrew, K., Bergen, L., Ethier, D., Hernberg, P., L<strong>in</strong>, J., and Tamon, C.<br />

(2007). Non-uniform mix<strong>in</strong>g of quantum walk on cycles, arXiv:0708.2096.<br />

Adamczak, W., Andrew, K., Hernberg, P., and Tamon, C. (2003). A note on graphs resistant<br />

to quantum uniform mix<strong>in</strong>g, quant-ph/0308073.<br />

Agarwal, G. S., and Pathak, P. K. (2005). Quantum random walk of the field <strong>in</strong> an externally<br />

driven cavity Phys. Rev. A, 72, 033815, quant-ph/0504135.<br />

Aharonov, D., Amba<strong>in</strong>is, A., Kempe, J., and Vazirani, U. V. (2001). Quantum walks on<br />

graphs, Proc. of the 33rd Annual ACM Symposium on <strong>Theory</strong> of Comput<strong>in</strong>g, 50-59,<br />

quant-ph/0012090.<br />

Aharonov, Y., Davidovich, L., and Zagury, N. (1993). Quantum random walks, Phys.<br />

Rev. A, 48, 1687–1690.<br />

Ahmadi, A., Belk, R., Tamon, C., and Wendler, C. (2003). On mix<strong>in</strong>g <strong>in</strong> cont<strong>in</strong>uous-time<br />

quantum walks on some circulant graphs, Quantum Information and Computation,<br />

3, 611-618, quant-ph/0209106.<br />

Alagic, G., and Russell, A. (2005). Decoherence <strong>in</strong> quantum walks on the hypercube, Phys.<br />

Rev. A, 72, 062304, quant-ph/0501169.<br />

Amba<strong>in</strong>is, A. (2003). Quantum walks and their algorithmic applications, International<br />

Journal of Quantum Information, 1, 507–518, quant-ph/0403120.<br />

Amba<strong>in</strong>is, A. (2004a). Quantum walk algorithm for element dist<strong>in</strong>ctness, Proceed<strong>in</strong>gs of<br />

the 45th Symposium on Foundations of Computer Science, 22–31, quant-ph/0311001.<br />

Amba<strong>in</strong>is, A. (2004b). Quantum search algorithms, SIGACT News, 35, 22–35.<br />

Amba<strong>in</strong>is, A., Bach, E., Nayak, A., Vishwanath, A., and Watrous, J. (2001). Onedimensional<br />

quantum walks, Proc. of the 33rd Annual ACM Symposium on <strong>Theory</strong><br />

of Comput<strong>in</strong>g, 37–49.<br />

Amba<strong>in</strong>is, A., Kempe, J., and Rivosh, A. (2005). Co<strong>in</strong>s make quantum walks faster, Proceed<strong>in</strong>gs<br />

of the 16th Annual ACM-SIAM Symposium on Discrete Algorithms, 1099–<br />

1108, quant-ph/0402107.<br />

Andrews, G. E., Askey, R., and Roy, R. (1999). Special Functions, Cambridge University<br />

Press.<br />

Aoun, B., and Tarifi, M. (2004). Introduction to quantum cellular automata, quant-ph/<br />

0401123.<br />

Aslangul, C. (2004). Quantum dynamics of a particle with a sp<strong>in</strong>-dependent velocity, quantph/0406057.<br />

Avetisov, V. A., Bikulov, A. H., Kozyrev, S. V., and Osipov, V. A. (2002). p-adic models<br />

of ultrametric diffusion constra<strong>in</strong>ed by hierarchical energy landscapes, J. Phys. A :<br />

Math. Gen., 35, 177–189.<br />

Avetisov, V. A., Bikulov, A. Kh., and Osipov, V. Al. (2003). p-adic description of characteristic<br />

relaxation <strong>in</strong> complex systems, J.Phys.A:Math.Gen., 36, 4239–7246.<br />

ben-Avraham, D., Bollt, E., and Tamon, C. (2004). One-dimensional cont<strong>in</strong>uous-time quantum<br />

walks, Quantum Information Process<strong>in</strong>g, 3, 295–308, cond-mat/0409514.<br />

Bach, E., Coppersmith, S., Goldschen, M. P., Joynt, R., and Watrous, J. (2004). Onedimensional<br />

quantum walks with absorption boundaries, Journal of Computer and<br />

System Sciences, 69, 562–592, quant-ph/0207008.<br />

Banuls, M. C., Navarrete, C., Perez, A,. Roldan, E., and Soriano, J. C. (2006). Quantum<br />

walk with a time-dependent co<strong>in</strong>, Phys. Rev. A, 73, 062304, quant-ph/0510046.<br />

Bartlett, S. D., Rudolph, T., Sanders, B. C., and Turner, P. S. (2006). Degradation of a<br />

quantum directional reference frame as a random walk, quant-ph/0607107.<br />

Bednarska, M., Grudka, A., Kurzyński, P., Luczak, T., and Wójcik, A. (2003). Quantum<br />

walks on cycles, Phys. Lett. A, 317, 21–25, quant-ph/0304113.<br />

Bednarska, M., Grudka, A., Kurzyński, P., Luczak, T., and Wójcik, A. (2004). Examples of<br />

nonuniform limit<strong>in</strong>g distributions for the quantum walk on even cycles, International<br />

Journal of Quantum Information, 2, 453–459, quant-ph/0403154.


Quantum Walks 443<br />

Bessen, A. J. (2006). Distributions of cont<strong>in</strong>uous-time quantum walks, quant-ph/0609128.<br />

Biane, P. (1991). Quantum random walk on the dual of SU(n), Probab. <strong>Theory</strong> Related<br />

Fields, 89, 117–129.<br />

Blanchard, Ph., and Hongler, M.-O. (2004). Quantum random walks and piecewise determ<strong>in</strong>istic<br />

evolutions, Phys. Rev. Lett., 92, 120601.<br />

Blumen, A., Bierbaum, V., and Muelken, O. (2006). Coherent dynamics on hierarchical<br />

systems, Physica A, 371, 10–15, cond-mat/0610686.<br />

Bracken, A. J., Ell<strong>in</strong>as, D., and Smyrnakis, I. (2007). Free Dirac evolution as a quantum<br />

random walk, Phys. Rev. A, 75, 022322, quant-ph/0605195.<br />

Bracken, A. J., Ell<strong>in</strong>as, D., and Tsohantjis, I. (2004). Pseudo memory effects, majorization<br />

and entropy <strong>in</strong> quantum random walks, J. Phys. A : Math. Gen., 37, L91–L97, quantph/0402187.<br />

Bressler, A., and Pemantle, R. (2007). Quantum random walks <strong>in</strong> one dimension via generat<strong>in</strong>g<br />

functions.<br />

Brun, T. A., Carteret, H. A., and Amba<strong>in</strong>is, A. (2003a). Quantum to classical transition<br />

for random walks, Phys. Rev. Lett., 91, 130602, quant-ph/0208195.<br />

Brun, T. A., Carteret, H. A., and Amba<strong>in</strong>is, A. (2003b). Quantum walks driven by many<br />

co<strong>in</strong>s, Phys. Rev. A, 67, 052317, quant-ph/0210161.<br />

Brun, T. A., Carteret, H. A., and Amba<strong>in</strong>is, A. (2003c). Quantum random walks with<br />

decoherent co<strong>in</strong>s, Phys. Rev. A, 67, 032304, quant-ph/0210180.<br />

Buerschaper, O., and Burnett, K. (2004). Stroboscopic quantum walks, quant-ph/0406039.<br />

Carlson, W., Ford, A., Harris, E., Rosen, J., Tamon, C., and Wrobel, K. (2006). Universal<br />

mix<strong>in</strong>g of quantum walk on graphs, quant-ph/0608044.<br />

Carneiro, I., Loo, M., Xu, X., Girerd, M., Kendon V., and Knight, P. L. (2005). Entanglement<br />

<strong>in</strong> co<strong>in</strong>ed quantum walks on regular graphs, New J. Phys., 7, 156, quantph/0504042.<br />

Carteret, H. A., Ismail, M. E. H., and Richmond, B. (2003). Three routes to the exact<br />

asymptotics for the one-dimensional quantum walk, J. Phys. A: Math. Gen., 36,<br />

8775–8795, quant-ph/0303105.<br />

Carteret, H. A., Richmond, B., and Temme, N. (2005). Evanescence <strong>in</strong> co<strong>in</strong>ed quantum<br />

walks, J. Phys. A: Math. Gen., 38, 8641–8665, quant-ph/0506048.<br />

Chandrashekar, C. M. (2006a). Implement<strong>in</strong>g the one-dimensional quantum (Hadamard)<br />

walk us<strong>in</strong>g a Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev. A, 74, 032307, quant-ph/0603156.<br />

Chandrashekar, C. M. (2006b). Discrete time quantum walk model for s<strong>in</strong>gle and entangled<br />

particles to reta<strong>in</strong> entanglement <strong>in</strong> co<strong>in</strong> space, quant-ph/0609113.<br />

Chandrashekar, C. M., and Laflamme, R. (2007). Quantum walk and quantum phase transition<br />

<strong>in</strong> optical lattice, arXiv:0709.1986.<br />

Chandrashekar, C. M., and Srikanth, R. (2006). Quantum walk with a bit flip, quant-ph/<br />

0607188.<br />

Chandrashekar, C. M., Srikanth, R., and Banerjee, S. (2007). Symmetries and noise <strong>in</strong><br />

quantum walk, Phys. Rev. A, 76, 022316, quant-ph/0607188.<br />

Chandrashekar, C. M., Srikanth, R., and Laflamme, R. (2007). Optimiz<strong>in</strong>g the discrete<br />

time quantum walk us<strong>in</strong>g a SU(2) co<strong>in</strong>, arXiv:0711.1882.<br />

Childs, A. M., Cleve, R., Deotto, E., Farhi, E., Gutmann, S., and Spielman, D. A. (2003).<br />

Exponential algorithmic speedup by quantum walk, Proc. of the 35th Annual ACM<br />

Symposium on <strong>Theory</strong> of Comput<strong>in</strong>g, 59–68, quant-ph/0209131.<br />

Childs, A. M., and Eisenberg, J. M. (2005). Quantum algorithms for subset f<strong>in</strong>d<strong>in</strong>g, Quantum<br />

Information and Computation, 5, 593-604, quant-ph/0311038.<br />

Childs, A. M., Farhi, E., and Gutmann, S. (2002). An example of the difference between<br />

quantum and classical random walks, Quantum Information Process<strong>in</strong>g, 1, 35–43,<br />

quant-ph/0103020.<br />

Childs, A. M., and Goldstone, J. (2004a). Spatial search by quantum walk, Phys. Rev. A,<br />

70, 022314, quant-ph/0306054.<br />

Childs, A. M., and Goldstone, J. (2004b). Spatial search and the Dirac equation, Phys.<br />

Rev. A, 70, 042312, quant-ph/0405120.


444 N. Konno<br />

Childs, A. M., Reichardt, B. W., Spalek, R., and Zhang, S. (2007). Every NAND formula of<br />

size N can be evaluated <strong>in</strong> time N 1/2+o(1) on a quantum computer, quant-ph/0703015.<br />

Childs, A. M., and Lee, T. (2007). Optimal quantum adversary lower bounds for ordered<br />

search, arXiv:0708.3396.<br />

D’Alessandro, D., Parlangeli, G., and Albert<strong>in</strong>i, F. (2007). Non-stationary quantum walks<br />

on the cycle, arXiv:0708.0184.<br />

de Falco, D., and Tamascelli, D. (2006). Speed and entropy of an <strong>in</strong>teract<strong>in</strong>g cont<strong>in</strong>uous<br />

time quantum walk, J. Phys. A: Math. Gen., 39, 5873–5895, quant-ph/0604067.<br />

Di, T., Hillery, M., and Zubairy, M. S. (2004). Cavity QED-based quantum walk, Phys.<br />

Rev. A, 70, 032304.<br />

Dodds, P. S., Watts, D. J., and Sabel, C. F. (2003). Information exchange and the robustness<br />

of organizational networks, Proc. Natl. Acad. Sci. USA, 100, 12516–12521.<br />

Doern, S. (2005). Quantum complexity bounds for <strong>in</strong>dependent set problems, quantph/0510084.<br />

Doern, S., and Thierauf, T. (2007). The quantum query complexity of algebraic properties,<br />

arXiv:0705.1446.<br />

Douglas, B. L., and Wang, J. B. (2007a). <strong>Classical</strong>ly efficient graph isomorphism algorithm<br />

us<strong>in</strong>g quantum walks, arXiv:0705.2531.<br />

Douglas, B. L., and Wang, J. B. (2007b). Can quantum walks provide exponential<br />

speedups?, arXiv:0706.0304.<br />

Du, J., Li, H., Xu, X., Shi, M., Wu, J., Zhou, X., and Han, R. (2003). Experimental<br />

implementation of the quantum random-walk algorithm, Phys. Rev. A, 67, 042316,<br />

quant-ph/0203120.<br />

Dür, W., Raussendorf, R., Kendon, V. M., and Briegel, H.-J. (2002). Quantum random<br />

walks <strong>in</strong> optical lattices, Phys. Rev. A, 66, 052319, quant-ph/0207137.<br />

Durrett, R. (1999). Essentials of Stochastic Processes, Spr<strong>in</strong>ger-Verlag, New York.<br />

Durrett, R. (2004). <strong>Probability</strong>: <strong>Theory</strong> and Examples, 3rd ed. Brooks-Cole, Belmont, CA.<br />

Eckert, K., Mompart, J., Birkl, G., and Lewenste<strong>in</strong>, M. (2005). One- and two-dimensional<br />

quantum walks <strong>in</strong> arrays of optical traps, Phys. Rev. A, 72, 012327, quant-ph/0503084.<br />

Ell<strong>in</strong>as, D. (2005). On algebraic and quantum random walks, Quantum <strong>Probability</strong> and<br />

Inf<strong>in</strong>ite Dimensional Analysis: From Foundations to Applications, QP-PQ Vol. 18, eds.<br />

M. Schurmann and U. Franz, (World Scientific, 2005), 174–200, quant-ph/0510128.<br />

Ell<strong>in</strong>as, D., and Smyrnakis, I. (2005). Asymptotics of quantum random walk driven by optical<br />

cavity, Journal of Optics B: Quantum Semiclass. Opt., 7, S152–S157, quant-ph/<br />

0510112.<br />

Ell<strong>in</strong>as, D., and Smyrnakis, I. (2006). Quantization and asymptotic behaviour of ɛ V k quantumrandomwalkon<strong>in</strong>tegers,Physica<br />

A, 365, 222–228, quant-ph/0510098.<br />

Ell<strong>in</strong>as, D., and Smyrnakis, I. (2007). Quantum optical random walk: quantization rules<br />

and quantum simulation of asymptotics, Phys. Rev. A, 76, 022333, quant-ph/0611265.<br />

Endrejat, J., and Buettner, H. (2005). Entanglement measurement with discrete multiple<br />

co<strong>in</strong> quantum walks, J. Phys. A: Math. Gen., 38, 9289–9296, quant-ph/0507184.<br />

Ermann, L., Paz, J. P., and Saraceno, M. (2006). Decoherence <strong>in</strong>duced by a chaotic environment:<br />

A quantum walker with a complex co<strong>in</strong>, Phys. Rev. A, 73, 012302, quant-ph/<br />

0510037.<br />

Farhi, E., Goldstone, J., and Gutmann, S. (2007). A quantum algorithm for the Hamiltonian<br />

NAND tree, quant-ph/0702144.<br />

Farhi, E., and Gutmann, S. (1998). Quantum computation and decision trees, Phys. Rev.<br />

A, 58, 915–928.<br />

Fedichk<strong>in</strong>, L., Solenov, D., and Tamon, C. (2006). Mix<strong>in</strong>g and decoherence <strong>in</strong> cont<strong>in</strong>uoustime<br />

quantum walks on cycles, Quantum Information and Computation, 6, 263–276,<br />

quant-ph/0509163.<br />

Feldman, E., and Hillery, M. (2004a). Quantum walks on graphs and quantum scatter<strong>in</strong>g<br />

theory, Proceed<strong>in</strong>gs of Conference on Cod<strong>in</strong>g <strong>Theory</strong> and Quantum Comput<strong>in</strong>g, quantph/0403066.


Quantum Walks 445<br />

Feldman, E., and Hillery, M. (2004b). Scatter<strong>in</strong>g theory and discrete-time quantum walks,<br />

Phys. Lett. A, 324, 277–281, quant-ph/0312062.<br />

Feldman, E., and Hillery, M. (2007). Modify<strong>in</strong>g quantum walks: A scatter<strong>in</strong>g theory approach,<br />

arXiv:0705.4612.<br />

Fenner, S. A., and Zhang, Y. (2003). A note on the classical lower bound for a quantum<br />

walk algorithm, quant-ph/0312230.<br />

Flitney, A. P., Abbott, D., and Johnson, N. F. (2004). Quantum random walks with history<br />

dependence, J. Phys. A: Math. Gen., 37, 7581–7591, quant-ph/0311009.<br />

Francisco, D., Iemmi, C., Paz, J. P., and Ledesma, S. (2006). Simulat<strong>in</strong>g a quantum walk<br />

with classical optics, Phys. Rev. A, 74, 052327.<br />

Fuss, I., White, L. B., Sherman, P. J., and Naguleswaran, S. (2006). Momentum dynamics<br />

of one dimensional quantum walks, quant-ph/0604197.<br />

Fuss, I., White, L. B., Sherman, P. J., and Naguleswaran, S. (2007). An analytic solution<br />

for one-dimensional quantum walks, arXiv:0705.0077.<br />

Gabris, A., Kiss, T., and Jex, I. (2007). Scatter<strong>in</strong>g quantum random-walk search with<br />

errors, quant-ph/0701150.<br />

Gerhardt, H., and Watrous, J. (2003). Cont<strong>in</strong>uous-time quantum walks on the symmetric<br />

group, <strong>in</strong> Proceed<strong>in</strong>gs of the 7th International Workshop on Randomization and<br />

Approximation Techniques <strong>in</strong> Computer Science, quant-ph/0305182.<br />

Gottlieb, A. D. (2003). Two examples of discrete-time quantum walks tak<strong>in</strong>g cont<strong>in</strong>uous<br />

steps, quant-ph/0310026.<br />

Gottlieb, A. D. (2005). Convergence of cont<strong>in</strong>uous-time quantum walks on the l<strong>in</strong>e, Phys.<br />

Rev. E, 72, 047102, quant-ph/0409042.<br />

Gottlieb, A. D., Janson, S., and Scudo, P. F. (2005). Convergence of co<strong>in</strong>ed quantum<br />

walks on d-dimensional Euclidean space, Inf<strong>in</strong>ite Dimensional Analysis, Quantum<br />

<strong>Probability</strong> and Related Topics, 8, 129–140, quant-ph/0406072.<br />

Grimmett, G., Janson, S., and Scudo, P. F. (2004). Weak limits for quantum random walks,<br />

Phys. Rev. E, 69, 026119, quant-ph/0309135.<br />

Grimmett, G. R., and Stirzaker, D. R. (1992). <strong>Probability</strong> and Random Processes. Oxford<br />

University Press Inc., New York.<br />

Gröss<strong>in</strong>g, G., and Zeil<strong>in</strong>ger, A. (1988a). Quantum cellular automata, Complex Systems, 2,<br />

197–208.<br />

Gröss<strong>in</strong>g, G., and Zeil<strong>in</strong>ger, A. (1988b). A conservation law <strong>in</strong> quantum cellular automata,<br />

Physica D, 31, 70–77.<br />

Grover, L. (1996). A fast quantum mechanical algorithm for database search, Proc. of the<br />

28th Annual ACM Symposium on <strong>Theory</strong> of Comput<strong>in</strong>g, 212–219, quant-ph/9605043.<br />

Gudder, S. P. (1988). Qunatum <strong>Probability</strong>. Academic Press Inc., CA.<br />

Hamada, M., Konno, N., and Segawa, E. (2005). Relation between co<strong>in</strong>ed quantum walks<br />

and quantum cellular automata, RIMS Kokyuroku, No. 1422, 1–11, quant-ph/0408100.<br />

Haselgrove, H. L. (2005). Optimal state encod<strong>in</strong>g for quantum walks and quantum communication<br />

over sp<strong>in</strong> systems, Phys. Rev. A, 72, 062326, quant-ph/0404152.<br />

Hashimoto, Y. (2001). Quantum decomposition <strong>in</strong> discrete groups and <strong>in</strong>teract<strong>in</strong>g Fock<br />

spaces, Inf<strong>in</strong>ite Dimensional Analysis, Quantum <strong>Probability</strong> and Related Topics, 4,<br />

277–287.<br />

Hillery, M., Bergou, J., and Feldman, E. (2003). Quantum walks based on an <strong>in</strong>terferometric<br />

analogy, Phys. Rev. A, 68, 032314, quant-ph/0302161.<br />

H<strong>in</strong>es, A. P., and Stamp, P. C. E. (2007a). Quantum walks, quantum gates, and quantum<br />

computers, Phys. Rev. A, 75, 062321, quant-ph/0701088.<br />

H<strong>in</strong>es, A. P., and Stamp, P. C. E. (2007b). Decoherence <strong>in</strong> quantum walks and quantum<br />

computers, arXiv:0711.1555.<br />

Inokuchi, S., and Mizoguchi, Y. (2005). Generalized partitioned quantum cellular automata<br />

and quantization of classical CA, International Journal of Unconventional Comput<strong>in</strong>g,<br />

1, 149–160, quant-ph/0312102.<br />

Inui, N., Inokuchi, S., Mizoguchi, Y., and Konno, N. (2005). Statistical properties for a<br />

quantum cellular automaton, Phys. Rev. A, 72, 032323, quant-ph/0504104.


446 N. Konno<br />

Inui, N., Kasahara, K., Konishi, Y., and Konno, N. (2005). Evolution of cont<strong>in</strong>uous-time<br />

quantum random walks on circles, Fluctuation and Noise Letters, 5, L73–L83, quantph/0402062.<br />

Inui, N., Konishi, Y., and Konno, N. (2004). Localization of two-dimensional quantum<br />

walks, Phys. Rev. A, 69, 052323, quant-ph/0311118.<br />

Inui, N., Konishi, Y., Konno, N., and Soshi, T. (2005). Fluctuations of quantum random<br />

walks on circles, International Journal of Quantum Information, 3, 535–550, quantph/0309204.<br />

Inui, N., and Konno, N. (2005). Localization of multi-state quantum walk <strong>in</strong> one dimension,<br />

Physica A, 353, 133–144, quant-ph/0403153.<br />

Inui, N., Konno, N., and Segawa, E. (2005). One-dimensional three-state quantum walk,<br />

Phys. Rev. E, 72, 056112, quant-ph/0507207.<br />

Inui, N., Nakamura, K., Ide, Y., and Konno, N. (2007). Effect of successive observation on<br />

quantum cellular automaton, Journal of the Physical Society of Japan, 76, 084001.<br />

Jafarizadeh, M. A., and Salimi, S. (2006). Investigation of cont<strong>in</strong>uous-time quantum walk<br />

via modules of Bose-Mesner and Terwilliger algebras, J.Phys.A:Math.Gen., 39,<br />

13295–13323, quant-ph/0603139.<br />

Jafarizadeh, M. A., and Salimi, S. (2007). Investigation of cont<strong>in</strong>uous-time quantum walk<br />

via spectral distribution associated with adjacency matrix, Annals of Physics, 322,<br />

1005–1033, quant-ph/0510174.<br />

Jafarizadeh, M. A., and Sufiani, R. (2006). Investigation of cont<strong>in</strong>uous-time quantum walk<br />

on root lattice An and honeycomb lattice, math-ph/0608067.<br />

Jafarizadeh, M. A., and Sufiani, R. (2007). Investigation of cont<strong>in</strong>uous-time quantum walks<br />

via spectral analysis and Laplace transform, arXiv:0704.2602.<br />

Jafarizadeh, M. A., Sufiani, R., Salimi, S., and Jafarizadeh, S. (2007). Investigation of<br />

cont<strong>in</strong>uous-time quantum walk by us<strong>in</strong>g Krylov subspace-Lanczos algorithm, Eur.<br />

Phys.J.B, 59, 199–216, quant-ph/0606241.<br />

Jeong, H., Paternostro, M., and Kim, M. S. (2004). Simulation of quantum random walks<br />

us<strong>in</strong>g <strong>in</strong>terference of classical field, Phys. Rev. A, 69, 012310, quant-ph/0305008.<br />

Joo, J., Knight, P. L., and Pachos, J. K. (2006). S<strong>in</strong>gle atom quantum walk with 1D optical<br />

superlattices, quant-ph/0606087.<br />

Katori, M., Fuj<strong>in</strong>o, S., and Konno, N. (2005). Quantum walks and orbital states of a Weyl<br />

particle, Phys. Rev. A, 72, 012316, quant-ph/0503142.<br />

Keat<strong>in</strong>g, J. P., L<strong>in</strong>den, N., Matthews, J. C. F., and W<strong>in</strong>ter, A. (2007). Localization and its<br />

consequences for quantum walk algorithms and quantum communication, Phys. Rev.<br />

A, 76, 012315, quant-ph/0606205.<br />

Kempe, J. (2002). Quantum random walks hit exponentially faster, Proc. of 7th Intern.<br />

Workshop on Randomization and Approximation Techniques <strong>in</strong> Comp. Sc. (RAN-<br />

DOM’03), 354-369, quant-ph/0205083.<br />

Kempe, J. (2003). Quantum random walks - an <strong>in</strong>troductory overview, Contemporary<br />

Physics, 44, 307–327, quant-ph/0303081.<br />

Kempe, J. (2005). Discrete quantum walks hit exponentially faster, Probab. <strong>Theory</strong> Related<br />

Fields, 133, 215–235.<br />

Kendon, V. (2006a). Quantum walks on general graphs, International Journal of Quantum<br />

Information, 4, 791–805, quant-ph/0306140.<br />

Kendon, V. (2006b). A random walk approach to quantum algorithms, 2006 Triennial<br />

Issue of Phil. Trans. R. Soc. A, 364, 3407–3422, quant-ph/0609035.<br />

Kendon, V. (2007). Decoherence <strong>in</strong> quantum walks - a review, Math. Struct. <strong>in</strong> Comp.<br />

Sci., 17, 1169-1220, quant-ph/0606016.<br />

Kendon, V., and Maloyer, O. (2006). Optimal computation with non-unitary quantum<br />

walks, quant-ph/0610240.<br />

Kendon, V., and Sanders, B. C. (2005). Complementarity and quantum walks, Phys. Rev.<br />

A, 71, 022307, quant-ph/0404043.<br />

Kendon, V., and Tregenna, B. (2003a). Decoherence can be useful <strong>in</strong> quantum walks, Phys.<br />

Rev. A, 67, 042315, quant-ph/0209005.


Quantum Walks 447<br />

Kendon, V., and Tregenna, B. (2003b). Decoherence <strong>in</strong> a quantum walk on the l<strong>in</strong>e, <strong>in</strong><br />

Proceed<strong>in</strong>gs of the 6th International Conference on Quantum Communication, Measurement<br />

and Comput<strong>in</strong>g, eds. J. H. Shapiro and O. Hirota (R<strong>in</strong>ton Press, Pr<strong>in</strong>ceton,<br />

NJ, 2003), quant-ph/0210047.<br />

Kendon, V., and Tregenna, B. (2003c). Decoherence <strong>in</strong> discrete quantum walks, <strong>in</strong> Quantum<br />

Decoherence and Entropy <strong>in</strong> Complex Systems, ed. H.-T. Elze (Spr<strong>in</strong>ger, Berl<strong>in</strong>,<br />

2003), quant-ph/0301182.<br />

Kesten, H. (1959). Symmetric random walks on groups, Transactions of the American<br />

Mathematical Society, 92, 336–354.<br />

Khrennikov, A. Yu., and Nilsson, M. (2004). P -adic Determ<strong>in</strong>istic and Random Dynamics,<br />

Kluwer Academic Publishers.<br />

Knight, P. L., Roldán, E., and Sipe, J. E. (2003a). Quantum walk on the l<strong>in</strong>e as an<br />

<strong>in</strong>terference phenomenon, Phys. Rev. A, 68, 020301, quant-ph/0304201.<br />

Knight, P. L., Roldán, E., and Sipe, J. E. (2003b). Optical cavity implementations of the<br />

quantum walk, Optics Communications, 227, 147–157, quant-ph/0305165.<br />

Knight, P. L., Roldán, E., and Sipe, J. E. (2004). Propagat<strong>in</strong>g quantum walks: the orig<strong>in</strong><br />

of <strong>in</strong>terference structures, J. Mod. Opt., 51, 1761–1777, quant-ph/0312133.<br />

Konno, N. (2002a). Quantum random walks <strong>in</strong> one dimension, Quantum Information<br />

Process<strong>in</strong>g, 1, 345–354, quant-ph/0206053.<br />

Konno, N. (2002b). Limit theorems and absorption problems for quantum random walks<br />

<strong>in</strong> one dimension, Quantum Information and Computation, 2, 578–595, quant-ph/<br />

0210011.<br />

Konno, N. (2003). Limit theorems and absorption problems for one-dimensional correlated<br />

random walks, quant-ph/0310191.<br />

Konno, N. (2005a). A new type of limit theorems for the one-dimensional quantum random<br />

walk, Journal of the Mathematical Society of Japan, 57, 1179–1195, quant-ph/<br />

0206103.<br />

Konno, N. (2005b). A path <strong>in</strong>tegral approach for disordered quantum walks <strong>in</strong> one dimension,<br />

Fluctuation and Noise Letters, 5, L529–L537, quant-ph/0406233.<br />

Konno, N. (2005c). Limit theorem for cont<strong>in</strong>uous-time quantum walk on the l<strong>in</strong>e, Phys.<br />

Rev. E, 72, 026113, quant-ph/0408140.<br />

Konno, N. (2006a). Cont<strong>in</strong>uous-time quantum walks on trees <strong>in</strong> quantum probability theory,<br />

Inf<strong>in</strong>ite Dimensional Analysis, Quantum <strong>Probability</strong> and Related Topics, 9, 287–<br />

297, quant-ph/0602213.<br />

Konno, N. (2006b). Cont<strong>in</strong>uous-time quantum walks on ultrametric spaces, International<br />

Journal of Quantum Information, 4, 1023–1035, quant-ph/0602070.<br />

Konno, N. (2008). Quantum Walks, Sangyo Tosho, Tokyo (<strong>in</strong> Japanese).<br />

Konno, N., Mitsuda, K., Soshi, T., and Yoo, H. J. (2004). Quantum walks and reversible<br />

cellular automata, Phys. Lett. A, 330, 408–417, quant-ph/0403107.<br />

Konno, N., Namiki, T., and Soshi, T. (2004). Symmetry of distribution for the onedimensional<br />

Hadamard walk, Interdiscipl<strong>in</strong>ary Information Sciences, 10, 11–22,<br />

quant-ph/0205065.<br />

Konno, N., Namiki, T., Soshi, T., and Sudbury, A. (2003). Absorption problems for quantum<br />

walks <strong>in</strong> one dimension, J. Phys. A: Math. Gen., 36, 241–253, quant-ph/0208122.<br />

Kosik, J., and Buzek, V. (2005). Scatter<strong>in</strong>g model for quantum random walks on a hypercube,<br />

Phys. Rev. A., 71, 012306, quant-ph/0410154.<br />

Kosik, J., Buzek, V., and Hillery, M. (2006). Quantum walks with random phase shifts,<br />

Phys. Rev. A., 74, 022310, quant-ph/0607092.<br />

Kosik, J., Miszczak, J. A., and Buzek, V. (2007). Quantum Parrondo’s game with random<br />

strategies, arXiv:0704.2937.<br />

Krovi, H. (2007). Symmetry <strong>in</strong> quantum walks, Ph.D. thesis, University of Southern California,<br />

2007, arXiv:0711.1694.<br />

Krovi, H., and Brun, T. A. (2006a). Hitt<strong>in</strong>g time for quantum walks on the hypercube,<br />

Phys. Rev. A, 73, 032341, quant-ph/0510136.


448 N. Konno<br />

Krovi, H., and Brun, T. A. (2006b). Quantum walks with <strong>in</strong>f<strong>in</strong>ite hitt<strong>in</strong>g times, Phys. Rev.<br />

A, 74, 042334, quant-ph/0606094.<br />

Krovi, H., and Brun, T. A. (2007). Quantum walks on quotient graphs, Phys. Rev. A, 75,<br />

062332, quant-ph/0701173.<br />

Kurzynski, P. (2006). Relativistic effects <strong>in</strong> quantum walks: Kle<strong>in</strong>’s paradox and Zitterbewegung,<br />

quant-ph/0606171.<br />

Lakshm<strong>in</strong>arayan, A. (2003). What is random about a quantum random walk?, quant-ph/<br />

0305026.<br />

Leroux, P. (2005). Coassociative grammar, periodic orbits and quantum random walk over<br />

Z 1 , International Journal of Mathematics and Mathematical Sciences, 2005, 3979–<br />

3996, quant-ph/0209100.<br />

Lo, P., Rajaram, S., Schepens, D., Sullivan, D., Tamon, C., and Ward, J. (2006). Mix<strong>in</strong>g<br />

of quantum walk on circulant bunkbeds, Quantum Information and Computation, 6,<br />

370–381, quant-ph/0509059.<br />

Lopez, C. C., and Paz, J. P. (2003). Phase-space approach to the study of decoherence <strong>in</strong><br />

quantum walks, Phys. Rev. A, 68, 052305, quant-ph/0308104.<br />

Love, P. J., and Boghosian, B. M. (2005). From Dirac to diffusion: decoherence <strong>in</strong> quantum<br />

lattice gases, Quantum Information Process<strong>in</strong>g, 4, 335–354, quant-ph/0507022.<br />

Ma, Z.-Y., Burnett, K., d’Arcy, M. B., and Gard<strong>in</strong>er, S. A. (2006). Quantum random walks<br />

us<strong>in</strong>g quantum accelerator modes, Phys. Rev. A, 73, 013401, physics/0508182.<br />

Mackay, T. D., Bartlett, S. D., Stephanson, L. T., and Sanders, B. C. (2002). Quantum<br />

walks <strong>in</strong> higher dimensions, J. Phys. A: Math. Gen., 35, 2745–2753, quant-ph/<br />

0108004.<br />

Macucci., M. (ed.) (2006). Quantum Cellular Automata, Imperial College Press.<br />

Magniez, F., Nayak, A., Roland, J., and Santha, M. (2006). Search via quantum walk,<br />

quant-ph/0608026.<br />

Magniez, F., Santha, M., and Szegedy, M. (2005). Quantum algorithm for the triangle<br />

problem, Proceed<strong>in</strong>gs of the sixteenth annual ACM-SIAM symposium on Discrete<br />

algorithms, January 23-25, 2005, Vancouver, British Columbia, quant-ph/0310134.<br />

Maloyer, O., and Kendon, V. (2007). Decoherence vs entanglement <strong>in</strong> co<strong>in</strong>ed quantum<br />

walks, New J. Phys., 9, 87, quant-ph/0612229.<br />

Manouchehri, K., and Wang, J. B. (2006a). Physical implementation of quantum random<br />

walks, quant-ph/0609088.<br />

Manouchehri, K., and Wang, J. B. (2006b). Cont<strong>in</strong>uous-time quantum random walks require<br />

discrete space, quant-ph/0611129.<br />

Mart<strong>in</strong>, X., O’Connor, D., and Sork<strong>in</strong>, R. D. (2005). Random walk <strong>in</strong> generalized quantum<br />

theory, Phys. Rev. D, 71, 024029, gr-qc/0403085.<br />

McGuigan, M. (2003). Quantum cellular automata from lattice field theories, quant-ph/<br />

0307176.<br />

Meyer, D. A. (1996). From quantum cellular automata to quantum lattice gases, J. Statist.<br />

Phys., 85, 551–574, quant-ph/9604003.<br />

Meyer, D. A. (1997). Qunatum mechanics of lattice gas automata: one particle plane waves<br />

and potentials, Phys. Rev. E, 55, 5261–5269, quant-ph/9611005.<br />

Meyer, D. A. (1998). Qunatum mechanics of lattice gas automata: boundary conditions<br />

and other <strong>in</strong>homogeneties, J.Phys.A:Math.Gen., 31, 2321–2340, quant-ph/9712052.<br />

Meyer, D. A., and Blumer, H. (2002). Parrondo games as lattice gas automata, J. Statist.<br />

Phys., 107, 225–239, quant-ph/0110028.<br />

Miyazaki, T., Katori, M., and Konno, N. (2007). Wigner formula of rotation matrices and<br />

quantum walks, Phys. Rev. A, 76, 012332, quant-ph/0611022.<br />

Montanaro, A. (2007). Quantum walks on directed graphs, Quantum Information and<br />

Computation, 7, 93–102, quant-ph/0504116.<br />

Moore, C., and Russell, A. (2001). Quantum walks on the hypercubes, quant-ph/0104137.<br />

Mülken, O. (2007). Inefficient quantum walks on networks: the role of the density of states,<br />

arXiv:0710.3453.


Quantum Walks 449<br />

Mülken, O., Bierbaum, V., and Blumen, A. (2006). Coherent exciton transport <strong>in</strong> dendrimers<br />

and cont<strong>in</strong>uous-time quantum walks, J. Chem. Phys., 124, 124905, cond-mat/<br />

0602040.<br />

Mülken, O., Bierbaum, V., and Blumen, A. (2007). Localization of coherent exciton transport<br />

<strong>in</strong> phase space, Phys. Rev. E, 75, 031121, quant-ph/0701034.<br />

Mülken, O., and Blumen, A. (2005a). Slow transport by cont<strong>in</strong>uous time quantum walks,<br />

Phys. Rev. E, 71, 016101, quant-ph/0410243.<br />

Mülken, O., and Blumen, A. (2005b). Spacetime structures of cont<strong>in</strong>uous time quantum<br />

walks, Phys. Rev. E, 71, 036128, quant-ph/0502004.<br />

Mülken, O., and Blumen, A. (2006a). Cont<strong>in</strong>uous time quantum walks <strong>in</strong> phase space,<br />

Phys. Rev. A, 73, 012105, quant-ph/0509141.<br />

Mülken, O., and Blumen, A. (2006b). Efficiency of quantum and classical transport on<br />

graphs, Phys. Rev. E, 73, 066117, quant-ph/0602120.<br />

Mülken, O., Blumen, A., Amthor, T., Giese, C., Reetz-Lamour, M., and Weidemueller, M.<br />

(2007). Survival probabilities <strong>in</strong> coherent exciton transfer with trapp<strong>in</strong>g, arXiv:0705.<br />

3700.<br />

Mülken, O., Pernice, V., and Blumen, A. (2007). Quantum transport on small-world<br />

networks: A cont<strong>in</strong>uous-time quantum walk approach, Phys. Rev. E, 76, 051125,<br />

arXiv:0705.1608.<br />

Mülken, O., Volta, A., and Blumen, A. (2005). Asymmetries <strong>in</strong> symmetric quantum walks<br />

on two-dimensional networks, Phys. Rev. A, 72, 042334, quant-ph/0507198.<br />

Navarrete, C., Perez, A., and Roldan, E. (2007). Nonl<strong>in</strong>ear optical Galton board, Phys.<br />

Rev. A, 75, 062333, quant-ph/0604084.<br />

Nayak, A., and Vishwanath, A. (2000). Quantum walk on the l<strong>in</strong>e, quant-ph/0010117.<br />

Nielsen, M. A., and Chuang, I. L. (2000). Qunatum Computation and Quantum Information,<br />

Cambridge University Press.<br />

Obata, N. (2004). Quantum probabilistic approach to spectral analysis of star graphs,<br />

Interdiscipl<strong>in</strong>ary Information Sciences, 10 41–52.<br />

Obata, N. (2006). A note on Konno’s paper on quantum walk, Inf<strong>in</strong>ite Dimensional Analysis,<br />

Quantum <strong>Probability</strong> and Related Topics, 9, 299–304.<br />

Ogielski, A. T., and Ste<strong>in</strong>, D. L. (1985). Dynamics on ultrametric spaces, Phys. Rev. Lett.,<br />

55, 1634–1637.<br />

Oka, T., Konno, N., Arita, R., and Aoki, H. (2005). Breakdown of an electric-field driven<br />

system: a mapp<strong>in</strong>g to a quantum walk, Phys. Rev. Lett., 94, 100602, quant-ph/<br />

0407013.<br />

Oliveira, A. C., Portugal, R., and Donangelo, R. (2006). Decoherence <strong>in</strong> two-dimensional<br />

quantum walks, Phys. Rev. A, 74, 012312.<br />

Oliveira, A. C., Portugal, R., and Donangelo, R. (2007). Simulation of the s<strong>in</strong>gle- and<br />

double-slit experiments with quantum walkers, arXiv:0706.3181.<br />

Omar, Y., Paunković, N., Sheridan, L., and Bose, S. (2006). Quantum walk on a l<strong>in</strong>e with<br />

two entangled particles, Phys. Rev. A, 74, 042304, quant-ph/0411065.<br />

Oliveira, A. C., Portugal, R., and Donangelo, R. (2006). Decoherence <strong>in</strong> two-dimensional<br />

quantum walks, Phys. Rev. A, 74, 012312.<br />

Oliveira, A. C., Portugal, R., and Donangelo, R. (2007). Simulation of the s<strong>in</strong>gle- and<br />

double-slit experiments with quantum walkers, arXiv:0706.3181.<br />

Osborne, T. J., and Sever<strong>in</strong>i, S. (2004). Quantum algorithms and cover<strong>in</strong>g spaces, quantph/0403127.<br />

Parashar, P. (2007). Equal superposition transformations and quantum random walks,<br />

arXiv:0709.3406.<br />

Patel, A., Raghunathan, K. S., and Rungta, P. (2005a). Quantum random walks do not<br />

need a co<strong>in</strong> toss, Phys. Rev. A, 71, 032347, quant-ph/0405128.<br />

Patel, A., Raghunathan, K. S., and Rungta, P. (2005b). Quantum random walks without<br />

a co<strong>in</strong> toss, Invited lecture at the Workshop on Quantum Information, Computation<br />

and Communication (QICC-2005), IIT Kharagpur, India, February 2005, quant-ph/<br />

0506221.


450 N. Konno<br />

Pathak, P. K., and Agarwal, G. S. (2007). Quantum random walk of two photons <strong>in</strong><br />

separable and entangled states, Phys. Rev. A, 75, 032351, quant-ph/0604138.<br />

Perets, H. B., Lah<strong>in</strong>i, Y., Pozzi, F., Sorel, M., Morandotti, R., and Silberberg, Y. (2007).<br />

Realization of quantum walks with negligible decoherence <strong>in</strong> waveguide lattices,<br />

arXiv:0707.0741.<br />

Prokofev, N. V., and Stamp, P. C. E. (2006). Decoherence and quantum walks: Anomalous<br />

diffusion and ballistic tails, Phys. Rev. A, 74, 020102, cond-mat/0605097.<br />

Ribeiro, P., Milman, P., and Mosseri, R. (2004). Aperiodic quantum random walks, Phys.<br />

Rev. Lett., 93, 190503, quant-ph/0406071.<br />

Richter, P. C. (2007a). Almost uniform sampl<strong>in</strong>g via quantum walks, New Journal of<br />

Physics, 9, 72, quant-ph/0606202.<br />

Richter, P. C. (2007b). Quantum speedup of classical mix<strong>in</strong>g processes, Phys. Rev. A, 76,<br />

042306, quant-ph/0609204.<br />

Roland, J., and Cerf, N. J. (2005). Noise resistance of adiabatic quantum computation<br />

us<strong>in</strong>g random matrix theory, Phys. Rev. A, 71, 032330.<br />

Roldan, E., and Soriano, J. C. (2005). Optical implementability of the two-dimensional<br />

quantum walk, Journal of Modern Optics, 52, 2649–2657, quant-ph/0503069.<br />

Romanelli, A. (2007). Measurements <strong>in</strong> the Lévy quantum walk, Phys. Rev. A, 76, 054306,<br />

arXiv:0710.1320.<br />

Romanelli, A., Auyuanet, A., Siri, R., Abal, G., and Donangelo, R. (2005). Generalized<br />

quantum walk <strong>in</strong> momentum space, Physica A, 352, 409–418, quant-ph/0408183.<br />

Romanelli, A., Sicardi Schif<strong>in</strong>o, A. C., Abal, G., Siri, R., and Donangelo, R. (2003). Markovian<br />

behaviour and constra<strong>in</strong>ed maximization of the entropy <strong>in</strong> chaotic quantum systems,<br />

Phys. Lett. A, 313, 325–329, quant-ph/0204135.<br />

Romanelli, A., Sicardi Schif<strong>in</strong>o, A. C., Siri, R., Abal, G., Auyuanet, A., and Donangelo, R.<br />

(2004). Quantum random walk on the l<strong>in</strong>e as a Markovian process, Physica A, 338,<br />

395–405, quant-ph/0310171.<br />

Romanelli, A., Siri, R., Abal, G., Auyuanet, A., and Donangelo, R. (2005). Decoherence<br />

<strong>in</strong> the quantum walk on the l<strong>in</strong>e, Physica A, 347, 137–152, quant-ph/0403192.<br />

Romanelli, A., Siri, R., and Micenmacher, V. (2007). Sub-ballistic behavior <strong>in</strong> quantum<br />

systems with Levy noise, Phys. Rev. E, 76, 037202, arXiv:0705.0370.<br />

Ryan, C. A., Laforest, M., Boileau, J. C., and Laflamme, R. (2005). Experimental implementation<br />

of discrete time quantum random walk on an NMR quantum <strong>in</strong>formation<br />

processor, Phys. Rev. A, 72, 062317, quant-ph/0507267.<br />

Salimi, S. (2007). Quantum central limit theorem for cont<strong>in</strong>uous-time quantum walks on<br />

odd graphs, arXiv:0710.3043.<br />

Salimi, S. (2007). Study of cont<strong>in</strong>uous-time quantum walks on quotient graphs via quantum<br />

probability theory, arXiv:0710.5813.<br />

Sch<strong>in</strong>azi, R. B. (1999). <strong>Classical</strong> and Spatial Stochastic Processes, Birkhäuser.<br />

Schumacher, B., and Werner, R. F. (2004). Reversible quantum cellular automata, quantph/0405174.<br />

Sever<strong>in</strong>i, S. (2002). The underly<strong>in</strong>g digraph of a co<strong>in</strong>ed quantum random walk, Erato<br />

Conference <strong>in</strong> Quantum Information Science, 2003, quant-ph/0210055.<br />

Sever<strong>in</strong>i, S. (2003). On the digraph of a unitary matrix, SIAM Journal on Matrix Analysis<br />

and Applications, 25, 295–300, math.CO/0205187.<br />

Sever<strong>in</strong>i, S. (2006). On the structure of the adjacency matrix of the l<strong>in</strong>e digraph of a regular<br />

digraph, Discrete Appl. Math., 154, 1663–1665.<br />

Sever<strong>in</strong>i, S., and Tanner, G. (2004). Regular quantum graphs, J. Phys. A: Math. Gen.,<br />

37, 6675–6686, nl<strong>in</strong>.CD/0312031.<br />

Shafee, F. (2005). Quantum measurement as first passage random walks <strong>in</strong> Hilbert space,<br />

quant-ph/0502111.<br />

Shapira, D., Biham, O., Bracken, A. J., and Hackett, M. (2003). One dimensional quantum<br />

walk with unitary noise, Phys. Rev. A, 68, 062315, quant-ph/0309063.<br />

Shenvi, N., Kempe, J., and Whaley, K. B. (2003). Quantum random-walk search algorithm,<br />

Phys. Rev. A, 67, 052307, quant-ph/0210064.


Quantum Walks 451<br />

Sicardi Shif<strong>in</strong>o, A. C., Abal, G., Siri, R., Romanelli, A., and Donangelo, R. (2003). Intr<strong>in</strong>sic<br />

decoherence and irreversibility <strong>in</strong> the quasiperiodic kicked rotor, quant-ph/0308162.<br />

Solenov, D., and Fedichk<strong>in</strong>, L. (2006a). Non-unitary quantum walks on hyper-cycles, Phys.<br />

Rev. A, 73, 012308, quant-ph/0509078.<br />

Solenov, D., and Fedichk<strong>in</strong>, L. (2006b). Cont<strong>in</strong>uous-time quantum walks on a cycle graph,<br />

Phys. Rev. A, 73, 012313, quant-ph/0506096.<br />

Spitzer, F. (1964). Pr<strong>in</strong>ciples of Random Walk, Van Nostrand, Pr<strong>in</strong>ceton, NJ.<br />

Stefanak, M., Jex, I., and Kiss, T. (2007). Recurrence and Pólya number of quantum walks,<br />

arXiv:0705.1991.<br />

Stefanak, M., Kiss, T., Jex, I., and Mohr<strong>in</strong>g, B. (2006). The meet<strong>in</strong>g problem <strong>in</strong> the<br />

quantum random walk, J. Phys. A: Math. Gen., 39, 14965–14983, arXiv:0705.1985.<br />

Strauch, F. W. (2006a). Relativistic quantum walks, Phys. Rev. A, 73, 054302, quant-ph/<br />

0508096. Erratum, Phys. Rev. A, 73, 069908 (2006).<br />

Strauch, F. W. (2006b). Connect<strong>in</strong>g the discrete and cont<strong>in</strong>uous-time quantum walks,<br />

Phys. Rev. A, 74, 030301, quant-ph/0606050.<br />

Szegedy, M. (2004). Spectra of quantized walks and a √ δɛ rule, quant-ph/0401053.<br />

Tani, S. (2007). An improved claw f<strong>in</strong>d<strong>in</strong>g algorithm us<strong>in</strong>g quantum walk, arXiv:0708.<br />

2584.<br />

Tanner, G. (2005). From quantum graphs to quantum random walks, Non-L<strong>in</strong>ear Dynamics<br />

and Fundamental Interactions. Proceed<strong>in</strong>gs of the NATO Advanced Research Workshop<br />

held October 10-16, 2004, <strong>in</strong> Tashkent, Uzbekistan. Edited by F. Khanna and<br />

D. Matrasulov, Published by Spr<strong>in</strong>ger, Dordrecht, The Netherlands, 2006, p.69, quantph/0504224.<br />

Taylor, J. M. (2007). A quantum dot implementation of the quantum NAND algorithm,<br />

arXiv:0708.1484.<br />

Travaglione, B. C., and Milburn, G. J. (2002). Implement<strong>in</strong>g the quantum random walk,<br />

Phys. Rev. A, 65, 032310, quant-ph/0109076.<br />

Tregenna, B., Flanagan, W., Maile, R., and Kendon, V. (2003). Controll<strong>in</strong>g discrete quantum<br />

walks: co<strong>in</strong>s and <strong>in</strong>itial states, New Journal of Physics, 5, 83, quant-ph/0304204.<br />

Tucci, R. R. (2007). How to compile some NAND formula evaluators, arXiv: 0706.0479.<br />

Venegas-Andraca, S. E., Ball, J. L., Burnett, K., and Bose, S. (2005). Quantum walks with<br />

entangled co<strong>in</strong>s, New Journal of Physics, 7, 221, quant-ph/0411151.<br />

Vlasov, A. Y. (2004). On quantum cellular automata, quant-ph/0406119.<br />

Vlasov, A. Y. (2007). Programmable quantum state transfer, arXiv:0708.0145.<br />

Volta, A., Muelken, O., and Blumen, A. (2006). Quantum transport on two-dimensional<br />

regular graphs, J.Phys.A:Math.Gen., 39, 14997, quant-ph/0610212.<br />

Wang, J. B., and Douglas, B. L. (2007). Graph identification by quantum walks, quant-ph/<br />

0701033.<br />

Watrous, J. (2001). Quantum simulations of classical random walks and undirected graph<br />

connectivity, Journal of Computer and System Sciences, 62, 376–391, cs.CC/9812012.<br />

Watson, G. N. (1944). A Treatise on the <strong>Theory</strong> of Bessel Functions, 2nd edition, Cambridge<br />

University Press, Cambridge.<br />

Watts, D. J., Dodds, P. S., and M. E. J. Newman, M. E. J. (2002). Identity and search <strong>in</strong><br />

social networks, Science, 296, 1302–1305.<br />

Watts, D. J., Muhamad, R., Med<strong>in</strong>a, D. C., and Dodds, P. S. (2005). Multiscale, resurgent<br />

epidemics <strong>in</strong> a hierarchical metapopulation model,Proc. Natl. Acad. Sci. USA, 102,<br />

11157–11162.<br />

Wocjan, P. (2004). Estimat<strong>in</strong>g mix<strong>in</strong>g properties of local Hamiltonian dynamics and cont<strong>in</strong>uous<br />

quantum random walks is PSPACE-hard, quant-ph/0401184.<br />

Wojcik, A., Luczak, T., Kurzynski, P., Grudka, A., and Bednarska, M. (2004). Quasiperiodic<br />

dynamics of a quantum walk on the l<strong>in</strong>e, Phys. Rev. Lett., 93, 180601, quant-ph/<br />

0407128.<br />

Wójcik, D. K., and Dorfman, J. R. (2003). Diffusive-ballistic crossover <strong>in</strong> 1D quantum<br />

walks, Phys. Rev. Lett., 90, 230602, quant-ph/0209036.


452 N. Konno<br />

Wójcik, D. K., and Dorfman, J. R. (2004). Crossover from diffusive to ballistic transport<br />

<strong>in</strong> periodic quantum maps, Physica D, 187, 223-243, nl<strong>in</strong>.CD/0212036.<br />

Wolfram, S. (2002). A New K<strong>in</strong>d of Science. Wolfram Media Inc.<br />

Yamasaki, T., Kobayashi, H., and Imai, H. (2003). Analysis of absorb<strong>in</strong>g times of quantum<br />

walks, Phys. Rev. A, 68, 012302, quant-ph/0205045.<br />

Y<strong>in</strong>, Y., Katsanos, D. E., and Evangelou, S. N. (2007). Quantum walks on a random<br />

environment, arXiv:0708.1137.<br />

Zhang, P., Ren, X. F., Zou, X. B., Liu, B. H., Huang, Y. F., and Guo, G. C. (2007). Demonstration<br />

of one-dimensional quantum random walks us<strong>in</strong>g orbital angular momentum<br />

of photons, Phys. Rev. A, 75, 052310.<br />

Zhao, Z., Du, J., Li, H., Yang, T., Chen, Z., and Pan, J. (2002). Implement quantum<br />

random walks with l<strong>in</strong>ear optics elements, quant-ph/0212149.


Index<br />

absorption problem, 387<br />

adjo<strong>in</strong>t pair, 294<br />

affiliated, 282<br />

algebra<br />

C ∗ ,63<br />

Hopf, 73<br />

von Neumann, 64<br />

angle bracket, 41<br />

annihilation operator, 291<br />

arcs<strong>in</strong>e law, 431<br />

balayage, 19, 21<br />

Banach-Mazur distance, 125<br />

basic criterion for selfadjo<strong>in</strong>tness, 281<br />

Bayes formula, 289<br />

Betti numbers, 259<br />

Beurl<strong>in</strong>g and Deny, 162, 163, 166, 168, 203,<br />

206<br />

Beurl<strong>in</strong>g-Deny-Le Jan decomposition, 238,<br />

239, 256<br />

bialgebra, 73<br />

Bochner identity, 253<br />

Brownian motion, 33, 49, 51<br />

noncommutative, 99<br />

C ∗ -algebra, 168, 172, 182, 184, 186, 189,<br />

192, 195, 197, 224, 227, 228, 230, 238,<br />

243, 265, 269<br />

amenable, 256<br />

CAR, 184<br />

CCR, 185, 217<br />

Clifford-, 249, 250, 253, 259<br />

commutative, 256, 260, 265<br />

f<strong>in</strong>ite dimensional, 183, 258<br />

full matrix-, 265<br />

group-, 239<br />

<strong>in</strong>ductive limit, 173<br />

norm of a-, 265<br />

norm topology, 266<br />

of cont<strong>in</strong>uous functions, 246<br />

of all bounded operators, 265, 269<br />

of all compact operators, 183, 213<br />

quasi local, 219, 224<br />

weak topology, 266<br />

c.b., see completely bounded<br />

c.p., see completely positive<br />

canonical commutation relations, 101<br />

capacity, 31<br />

Cauchy<br />

completion, 57, 58<br />

problem, 54<br />

Chapman-Kolmogorov equation, 26, 28<br />

character, 82<br />

classical Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality, 137<br />

column space, 122<br />

column p-space, 136<br />

commutant, 280<br />

compatible couple, 128<br />

complete isometry, 121<br />

complete isomorphism, 121<br />

completely bounded map, 121<br />

completely isometric map, 121<br />

completely positive map, 119<br />

completely unconditional, 150<br />

cone<br />

non self-polar, 175<br />

positive, 178, 179<br />

self-polar, 170, 178<br />

standard, 174<br />

of f<strong>in</strong>ite von Neumann algebras, 174<br />

Connes, 162, 262<br />

conserved quantity, 362<br />

Constant<strong>in</strong>escu-Cornea theorem, 22<br />

creation operator, 291<br />

453


454 Index<br />

curvature, 246<br />

Gauss-, 248, 259<br />

operator, 247, 248, 253, 254, 257–259<br />

Ricci-, 246<br />

sectional, 254<br />

tensor, 254<br />

cyclic cohomology, 265<br />

de Moivre-Laplace theorem, 315<br />

derivation, 224, 227, 228, 230, 235, 237,<br />

248, 253, 260, 262<br />

and stochastic calculus, 228<br />

approximately bounded, 255, 256<br />

bounded, 219, 255, 257<br />

cha<strong>in</strong>-rule, 229<br />

closable, 219<br />

closed, 229<br />

completely unbounded, 255, 256<br />

covariant, 210, 246, 247<br />

decomposition, 239, 255, 256<br />

<strong>in</strong> free probability, 243, 244<br />

on Clifford C ∗ algebra, 242<br />

on group C ∗ -algebras, 241<br />

on the noncommutative torus, 243<br />

universal, 238<br />

detailed balance, 166, 181, 186, 187<br />

dilation, 67<br />

direct sum, 127<br />

Dirichlet<br />

form, 30<br />

problem, 3, 9–13, 17, 30, 48<br />

Dirichlet algebra, 231, 235, 238, 253,<br />

260–262<br />

Dirichlet form, 205, 207, 214, 224, 226, 228<br />

and Cartan’s symmetric spaces, 248<br />

and hypersurfaces, 248<br />

and Noncommutative Geometry, 259<br />

and topology of Riemann surfaces, 248<br />

bounded, 207<br />

C ∗ -, 226, 228, 231, 235–237, 248<br />

completely-, 221, 222, 226, 227, 231, 240,<br />

245, 256<br />

Dirac-, 257<br />

<strong>in</strong> free probability, 243<br />

<strong>in</strong> quantum statistical mechanics, 218<br />

nonsymmetric, 210<br />

on amenable C ∗ -algebras, 256<br />

on CAR algebra, 216<br />

on CCR algebra, 217<br />

on Clifford C ∗ -algebras, 242<br />

on commutative C ∗ -algebras, 238<br />

on group C ∗ -algebras, 239<br />

on type I factor, 211<br />

quasi-conformal equivalence, 261<br />

regular, 226, 230, 262<br />

strongly local, 261<br />

unbounded, 210<br />

Dirichlet <strong>in</strong>tegral, 227, 246, 261, 262<br />

disordered quantum walk, 351<br />

divergence, 5<br />

formula, 8<br />

theorem, 5<br />

Doob’s dondition<strong>in</strong>g, 86<br />

dual<br />

standard, 127<br />

Dynk<strong>in</strong> formula, 44, 47<br />

electrical<br />

field, 4<br />

permittivity, 4<br />

exponential, 88<br />

Feynman-Kac formula, 54<br />

filtration, 31<br />

Fock space, 97<br />

Fourier transform, 177<br />

fractal, 161, 167, 264<br />

Fredholm module, 262–264<br />

free entropy, 244, 245<br />

free Gaussian variable, 148<br />

free generators, 144<br />

free probability, 243<br />

free von Neumann algebra, 148<br />

free Weyl equation, 337<br />

function calculus, 279<br />

fundamental noises, 291<br />

gauge process, 291<br />

Gauss<br />

<strong>in</strong>tegral theorem, 12<br />

law, 4<br />

generalized circular system, 154<br />

generator, 27, 30<br />

gradient, 227<br />

Green<br />

identity, 12<br />

kernel, 12, 14, 46<br />

potential, 18, 46, 52<br />

group<br />

free, 107<br />

Heisenberg, 100<br />

quantum, 112<br />

group von Neumann algebra, 144<br />

Grover walk, 339<br />

Haagerup-Paulsen-Wittstock factorization,<br />

124<br />

Hadamard walk, 314


Index 455<br />

Hahn-Banach type extension, 124<br />

harmonic, 7, 8, 16, 21, 48<br />

function, 47<br />

fundamental harmonic function, 9<br />

superharmonic function, 7, 8, 16, 47, 48<br />

harmonic function, 86<br />

Hilbert transform, 244<br />

Hilbert transform <strong>in</strong> free probability, 245<br />

Hilbertian, 125<br />

homogeneous, 125<br />

<strong>in</strong>dependent <strong>in</strong>crements, 24, 32<br />

<strong>in</strong>formation state, 300<br />

<strong>in</strong>novat<strong>in</strong>g mart<strong>in</strong>gale, 303<br />

<strong>in</strong>novations, 303<br />

<strong>in</strong>stantaneous uniform mix<strong>in</strong>g property,<br />

434<br />

<strong>in</strong>teract<strong>in</strong>g Fock space, 412<br />

<strong>in</strong>terpolation, 130, 180<br />

<strong>in</strong>tersection, 128<br />

isomorphism<br />

complete, 121<br />

Itô<br />

formula, 43<br />

<strong>in</strong>tegral, 36<br />

isometry, 33, 34, 36<br />

Jordan decomposition<br />

<strong>in</strong> commutative von Neumann algebras,<br />

170<br />

<strong>in</strong> self-polar cones, 170<br />

<strong>in</strong> type I and type II von Neumann<br />

algebras, 171<br />

K-theory, 262, 264<br />

kernel<br />

λ-potential kernel, 28, 46, 52<br />

potential kernel, 53<br />

transition kernel, 25, 26<br />

Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality<br />

classical, 137<br />

for generalized circular system, 154<br />

for semicircular system, 149<br />

free generators, 145<br />

noncommutative, 141<br />

Kh<strong>in</strong>tch<strong>in</strong>e-Kahane <strong>in</strong>equality, 137<br />

Killed Brownian motion, 45, 51<br />

KMS<br />

condition, 184<br />

duality, 186<br />

symmetry, 186, 191<br />

Laplace<br />

equation, 5, 8, 9<br />

Laplacian, 5<br />

Laplacian operator, 185, 217, 218<br />

Bochner-, 165, 246, 251–254, 257<br />

Dirac-, 167, 247, 251–254, 259<br />

Hodge-de Rham-, 246, 247, 252<br />

transverse, 165<br />

Leibniz rule, 210, 227, 229, 239, 241, 242,<br />

253, 260, 263<br />

L<strong>in</strong>dblad generator, 183<br />

localization, 349<br />

map<br />

completely bounded, 121<br />

completely isometric, 121<br />

completely positive, 119<br />

positive, 119<br />

Markov<br />

cha<strong>in</strong>, 25, 26<br />

process, 24<br />

strong Markov property, 29, 45<br />

sub-Markovian semigroup, 30<br />

Markov cha<strong>in</strong>, 84, 112<br />

Mart<strong>in</strong><br />

boundary, 21, 22, 24, 57, 90<br />

compactification, 90<br />

space, 22, 58<br />

mart<strong>in</strong>gale, 31, 39<br />

Azéma, 42<br />

normal, 32, 36, 39<br />

Poisson, 36, 39<br />

supermart<strong>in</strong>gale, 48<br />

maximal structure, 122<br />

maximal torus, 85<br />

Maxwell equation, 5<br />

mean value property, 7, 13, 47<br />

m<strong>in</strong>imal structure, 121<br />

m<strong>in</strong>imal tensor product, 134<br />

modular<br />

automorphism group, 172<br />

on Powers factors, 173<br />

on type I factors, 173<br />

condition, 172<br />

conjugation, 171<br />

operator, 171<br />

on type I factors, 177<br />

Mokobodzki, 163<br />

No-Go lemma, 404<br />

noncommutative Lp-space, 132<br />

operator space structure, 133<br />

Noncommutative Geometry, 162, 165, 167,<br />

243, 259, 262<br />

noncommutative Kh<strong>in</strong>tch<strong>in</strong>e <strong>in</strong>equality,<br />

141<br />

Rademacher sequence, 141


456 Index<br />

nondemolition pr<strong>in</strong>ciple, 287<br />

normal<br />

derivative, 11<br />

vector, 5<br />

OH, 131<br />

operator<br />

annihilation, 96<br />

creation, 96<br />

operator Hilbert space, 131<br />

operator space<br />

abstract, 126<br />

concrete, 120<br />

Hilbertian, 125<br />

homogeneous, 125<br />

path <strong>in</strong>tegral, 315<br />

phase operator, 168, 262, 263<br />

Poisson<br />

compound process, 35<br />

distribution, 35<br />

equation, 5, 6, 50, 51, 53<br />

formula, 11, 15, 16<br />

<strong>in</strong>tegral, 15<br />

kernel, 16, 24<br />

mart<strong>in</strong>gale, 41, 42<br />

process, 35<br />

polar set, 29<br />

positive map, 119<br />

potential, 20, 57<br />

λ-potential, 28<br />

Green potential, 18, 46, 52<br />

logarithmic potential, 9<br />

Newton potential, 6, 9, 18, 28<br />

predictable representation property, 41<br />

process<br />

adapted process, 36<br />

predictable process, 36<br />

quantum Bessel, 103<br />

sp<strong>in</strong>, 70<br />

quadratic variation, 39–41<br />

quantum cellular automaton, 368<br />

quantum central limit theorem, 416<br />

quantum conditional expectation, 288<br />

quantum filter<strong>in</strong>g equation, 303<br />

quantum Itô rule, 294<br />

quantum lattice gas automaton, 374<br />

quantum probability space, 280<br />

quantum sp<strong>in</strong> system, 219<br />

Is<strong>in</strong>g and Heisenberg models, 220<br />

quantum stochastic differential equation,<br />

293<br />

quantum stochastic <strong>in</strong>tegral, 291<br />

quantum walk, 312<br />

qubit, 312<br />

quotient, 127<br />

random walk, 311<br />

Bernoulli, 68<br />

quantum, 68<br />

recurrent, 29<br />

reduced function, 18, 19, 21<br />

reference state, 301<br />

resolvent, 282<br />

equation, 27<br />

operator, 27<br />

reversible cellular automaton, 358<br />

row p-space, 136<br />

row space, 122<br />

Ruan’s axioms, 126<br />

Ruan’s characterization, 126<br />

second quantization, 284<br />

self-nondemolition, 299<br />

semicircular element, 148<br />

semigroup<br />

birth-and-death process-, 211<br />

completely Markovian, 182, 201, 228,<br />

240, 247, 252, 253<br />

<strong>in</strong> free probability, 245<br />

KMS-symmetric, 188<br />

completely positive, 182, 201<br />

norm cont<strong>in</strong>uous-, 182<br />

conservative, 237<br />

convolution, 198<br />

dynamical, 182<br />

ergodic, 201–203, 217, 218, 223<br />

Feller semigroup, 27<br />

Gaussian semigroup, 28<br />

generator, 268<br />

heat-, 246<br />

on the noncommutative torus, 243<br />

KMS-symmetric, 168, 181, 186, 195<br />

Markovian, 181, 182, 198, 201, 205, 211,<br />

223, 224, 247, 257, 259<br />

KMS-symmetric, 203<br />

nonsymmetric, 210<br />

on Hilbert spaces of standard forms,<br />

197<br />

on standard forms, 197<br />

traces, 226<br />

nonergodic, 259<br />

norm cont<strong>in</strong>uous, 267<br />

positive, 182, 201, 206, 225<br />

on Hilbert spaces of standard forms,<br />

197<br />

quantum Brownian motion-, 211


Index 457<br />

quantum Ornste<strong>in</strong>-Uhlenbeck-, 211<br />

Schröd<strong>in</strong>ger-, 198<br />

semigroup property, 26, 54<br />

strongly cont<strong>in</strong>uous, 267<br />

strongly cont<strong>in</strong>uous semigroup, 30<br />

subord<strong>in</strong>ation, 198<br />

transition semigroup, 26<br />

weakly cont<strong>in</strong>uous, 267<br />

weakly ∗ cont<strong>in</strong>uous, 267<br />

space of c.b. maps, 126<br />

spectral theorem, 278, 280<br />

spectrum of an operator, 278<br />

standard form of a von Neumann algebra,<br />

175<br />

generalized Hilbert-Schmidt, 176<br />

of group von Neumann algebras, 177<br />

of Hilbert-Schmidt on type I factors, 176<br />

uniqueness, 175<br />

standard semicircular system, 148<br />

state, 64, 169<br />

factor, 223<br />

ground-, 198, 202<br />

KMS, 183<br />

extremal, 223<br />

on CAR algebras, 184<br />

on CCR algebras, 185<br />

normal, 201<br />

pure, 95<br />

quasi-free, 216, 217<br />

regular, 217<br />

tracial, 243<br />

St<strong>in</strong>espr<strong>in</strong>g factorization, 120<br />

stochastic differential equation, 45<br />

stochastic <strong>in</strong>tegral, 36<br />

Stone generator, 282<br />

Stone’s theorem, 282<br />

stopp<strong>in</strong>g time, 29<br />

structure equation, 42, 43<br />

sum, 128<br />

surface measure, 4<br />

normalized surface measure, 7<br />

symmetric embedd<strong>in</strong>g, 179<br />

on type I factors, 180<br />

symmetric Fock space, 283<br />

Szegö-Jacobi parameter, 412<br />

theorem<br />

Choquet-Deny, 87<br />

Pitman, 110<br />

time<br />

discrete time, 48, 53, 57<br />

exit time, 29, 47<br />

hitt<strong>in</strong>g time, 29, 45<br />

time homogeneous, 26<br />

Tomita-Takesaki<br />

modular theory, 168<br />

theorem, 172<br />

transient, 29, 57<br />

transition<br />

transition density, 28, 45<br />

transition operator, 25, 45, 57<br />

ultrametric space, 420<br />

vector<br />

cyclic and separat<strong>in</strong>g, 171<br />

excessive, 198<br />

vector-valued Schatten classe, 134<br />

von Neumann algebra, 168, 170, 172, 180,<br />

181, 184, 187, 188, 192, 197, 199, 201,<br />

203, 205, 207, 214, 218, 221, 225, 227,<br />

265, 280<br />

commutant-, 171, 266<br />

commutative, 164, 170, 202, 266<br />

factor, 266<br />

group-, 177, 240<br />

topology<br />

σ-strong, 266<br />

σ-strong ∗ , 267<br />

σ-weak, 267<br />

strong, 266<br />

strong ∗ , 266<br />

weak ∗ , 267<br />

type I∞ factor, 211, 266<br />

type I and II, 171<br />

type II hyperf<strong>in</strong>ite factor, 242<br />

weak limit theorem, 328<br />

weight, 65<br />

Weyl operators, 283<br />

Weyl relations, 283


Lecture Notes <strong>in</strong> Mathematics<br />

For <strong>in</strong>formation about earlier volumes<br />

please contact your bookseller or Spr<strong>in</strong>ger<br />

LNM Onl<strong>in</strong>e archive: spr<strong>in</strong>gerl<strong>in</strong>k.com<br />

Vol. 1774: V. Runde, Lectures on Amenability (2002)<br />

Vol. 1775: W. H. Meeks, A. Ros, H. Rosenberg, The<br />

Global <strong>Theory</strong> of M<strong>in</strong>imal Surfaces <strong>in</strong> Flat Spaces.<br />

Mart<strong>in</strong>a Franca 1999. Editor: G. P. Pirola (2002)<br />

Vol. 1776: K. Behrend, C. Gomez, V. Tarasov, G. Tian,<br />

Quantum Comohology. Cetraro 1997. Editors: P. de Bartolomeis,<br />

B. Dubrov<strong>in</strong>, C. Re<strong>in</strong>a (2002)<br />

Vol. 1777: E. García-Río, D. N. Kupeli, R. Vázquez-<br />

Lorenzo, Osserman Manifolds <strong>in</strong> Semi-Riemannian<br />

Geometry (2002)<br />

Vol. 1778: H. Kiechle, <strong>Theory</strong> of K-Loops (2002)<br />

Vol. 1779: I. Chueshov, Monotone Random Systems<br />

(2002)<br />

Vol. 1780: J. H. Bru<strong>in</strong>ier, Borcherds Products on O(2,1)<br />

and Chern Classes of Heegner Divisors (2002)<br />

Vol. 1781: E. Bolthausen, E. Perk<strong>in</strong>s, A. van der Vaart,<br />

Lectures on <strong>Probability</strong> <strong>Theory</strong> and Statistics. Ecole d’<br />

Eté de Probabilités de Sa<strong>in</strong>t-Flour XXIX-1999. Editor:<br />

P. Bernard (2002)<br />

Vol. 1782: C.-H. Chu, A. T.-M. Lau, Harmonic Functions<br />

on Groups and Fourier Algebras (2002)<br />

Vol. 1783: L. Grüne, Asymptotic Behavior of Dynamical<br />

and Control Systems under Perturbation and Discretization<br />

(2002)<br />

Vol. 1784: L. H. Eliasson, S. B. Kuks<strong>in</strong>, S. Marmi, J.-C.<br />

Yoccoz, Dynamical Systems and Small Divisors. Cetraro,<br />

Italy 1998. Editors: S. Marmi, J.-C. Yoccoz (2002)<br />

Vol. 1785: J. Arias de Reyna, Po<strong>in</strong>twise Convergence of<br />

Fourier Series (2002)<br />

Vol. 1786: S. D. Cutkosky, Monomialization of Morphisms<br />

from 3-Folds to Surfaces (2002)<br />

Vol. 1787: S. Caenepeel, G. Militaru, S. Zhu, Frobenius<br />

and Separable Functors for Generalized Module Categories<br />

and Nonl<strong>in</strong>ear Equations (2002)<br />

Vol. 1788: A. Vasil’ev, Moduli of Families of Curves for<br />

Conformal and Quasiconformal Mapp<strong>in</strong>gs (2002)<br />

Vol. 1789: Y. Sommerhäuser, Yetter-Dr<strong>in</strong>fel’d Hopf algebras<br />

over groups of prime order (2002)<br />

Vol. 1790: X. Zhan, Matrix Inequalities (2002)<br />

Vol. 1791: M. Knebusch, D. Zhang, Manis Valuations<br />

and Prüfer Extensions I: A new Chapter <strong>in</strong> Commutative<br />

Algebra (2002)<br />

Vol. 1792: D. D. Ang, R. Gorenflo, V. K. Le, D. D. Trong,<br />

Moment <strong>Theory</strong> and Some Inverse Problems <strong>in</strong> <strong>Potential</strong><br />

<strong>Theory</strong> and Heat Conduction (2002)<br />

Vol. 1793: J. Cortés Monforte, Geometric, Control and<br />

Numerical Aspects of Nonholonomic Systems (2002)<br />

Vol. 1794: N. Pytheas Fogg, Substitution <strong>in</strong> Dynamics,<br />

Arithmetics and Comb<strong>in</strong>atorics. Editors: V. Berthé, S.<br />

Ferenczi, C. Mauduit, A. Siegel (2002)<br />

Vol. 1795: H. Li, Filtered-Graded Transfer <strong>in</strong> Us<strong>in</strong>g Noncommutative<br />

Gröbner Bases (2002)<br />

Vol. 1796: J.M. Melenk, hp-F<strong>in</strong>ite Element Methods for<br />

S<strong>in</strong>gular Perturbations (2002)<br />

Vol. 1797: B. Schmidt, Characters and Cyclotomic Fields<br />

<strong>in</strong> F<strong>in</strong>ite Geometry (2002)<br />

Vol. 1798: W.M. Oliva, Geometric Mechanics (2002)<br />

Vol. 1799: H. Pajot, Analytic Capacity, Rectifiability,<br />

Menger Curvature and the Cauchy Integral (2002)<br />

Vol. 1800: O. Gabber, L. Ramero, Almost R<strong>in</strong>g <strong>Theory</strong><br />

(2003)<br />

Vol. 1801: J. Azéma, M. Émery, M. Ledoux, M. Yor<br />

(Eds.), Sém<strong>in</strong>aire de Probabilités XXXVI (2003)<br />

Vol. 1802: V. Capasso, E. Merzbach, B. G. Ivanoff,<br />

M. Dozzi, R. Dalang, T. Mountford, Topics <strong>in</strong> Spatial<br />

Stochastic Processes. Mart<strong>in</strong>a Franca, Italy 2001. Editor:<br />

E. Merzbach (2003)<br />

Vol. 1803: G. Dolzmann, Variational Methods for Crystall<strong>in</strong>e<br />

Microstructure – Analysis and Computation<br />

(2003)<br />

Vol. 1804: I. Cherednik, Ya. Markov, R. Howe, G.<br />

Lusztig, Iwahori-Hecke Algebras and their Representation<br />

<strong>Theory</strong>. Mart<strong>in</strong>a Franca, Italy 1999. Editors: V. Baldoni,<br />

D. Barbasch (2003)<br />

Vol. 1805: F. Cao, Geometric Curve Evolution and Image<br />

Process<strong>in</strong>g (2003)<br />

Vol. 1806: H. Broer, I. Hoveijn. G. Lunther, G. Vegter,<br />

Bifurcations <strong>in</strong> Hamiltonian Systems. Comput<strong>in</strong>g S<strong>in</strong>gularities<br />

by Gröbner Bases (2003)<br />

Vol. 1807: V. D. Milman, G. Schechtman (Eds.), Geometric<br />

Aspects of Functional Analysis. Israel Sem<strong>in</strong>ar 2000-<br />

2002 (2003)<br />

Vol. 1808: W. Sch<strong>in</strong>dler, Measures with Symmetry Properties<br />

(2003)<br />

Vol. 1809: O. Ste<strong>in</strong>bach, Stability Estimates for Hybrid<br />

Coupled Doma<strong>in</strong> Decomposition Methods (2003)<br />

Vol. 1810: J. Wengenroth, Derived Functors <strong>in</strong> Functional<br />

Analysis (2003)<br />

Vol. 1811: J. Stevens, Deformations of S<strong>in</strong>gularities<br />

(2003)<br />

Vol. 1812: L. Ambrosio, K. Deckelnick, G. Dziuk,<br />

M. Mimura, V. A. Solonnikov, H. M. Soner, Mathematical<br />

Aspects of Evolv<strong>in</strong>g Interfaces. Madeira, Funchal,<br />

Portugal 2000. Editors: P. Colli, J. F. Rodrigues (2003)<br />

Vol. 1813: L. Ambrosio, L. A. Caffarelli, Y. Brenier,<br />

G. Buttazzo, C. Villani, Optimal Transportation and its<br />

Applications. Mart<strong>in</strong>a Franca, Italy 2001. Editors: L. A.<br />

Caffarelli, S. Salsa (2003)<br />

Vol. 1814: P. Bank, F. Baudo<strong>in</strong>, H. Föllmer, L.C.G.<br />

Rogers, M. Soner, N. Touzi, Paris-Pr<strong>in</strong>ceton Lectures on<br />

Mathematical F<strong>in</strong>ance 2002 (2003)<br />

Vol. 1815: A. M. Vershik (Ed.), Asymptotic Comb<strong>in</strong>atorics<br />

with Applications to Mathematical Physics.<br />

St. Petersburg, Russia 2001 (2003)<br />

Vol. 1816: S. Albeverio, W. Schachermayer, M. Talagrand,<br />

Lectures on <strong>Probability</strong> <strong>Theory</strong> and Statistics.<br />

Ecole d’Eté de Probabilités de Sa<strong>in</strong>t-Flour XXX-2000.<br />

Editor: P. Bernard (2003)


Vol. 1817: E. Koel<strong>in</strong>k, W. Van Assche (Eds.), Orthogonal<br />

Polynomials and Special Functions. Leuven 2002 (2003)<br />

Vol. 1818: M. Bildhauer, Convex Variational Problems<br />

with L<strong>in</strong>ear, nearly L<strong>in</strong>ear and/or Anisotropic Growth<br />

Conditions (2003)<br />

Vol. 1819: D. Masser, Yu. V. Nesterenko, H. P. Schlickewei,<br />

W. M. Schmidt, M. Waldschmidt, Diophant<strong>in</strong>e<br />

Approximation. Cetraro, Italy 2000. Editors: F. Amoroso,<br />

U. Zannier (2003)<br />

Vol. 1820: F. Hiai, H. Kosaki, Means of Hilbert Space<br />

Operators (2003)<br />

Vol. 1821: S. Teufel, Adiabatic Perturbation <strong>Theory</strong> <strong>in</strong><br />

Quantum Dynamics (2003)<br />

Vol. 1822: S.-N. Chow, R. Conti, R. Johnson, J. Mallet-<br />

Paret, R. Nussbaum, Dynamical Systems. Cetraro, Italy<br />

2000. Editors: J. W. Macki, P. Zecca (2003)<br />

Vol. 1823: A. M. Anile, W. Allegretto, C. R<strong>in</strong>ghofer,<br />

Mathematical Problems <strong>in</strong> Semiconductor Physics.<br />

Cetraro, Italy 1998. Editor: A. M. Anile (2003)<br />

Vol. 1824: J. A. Navarro González, J. B. Sancho de Salas,<br />

C ∞ – Differentiable Spaces (2003)<br />

Vol. 1825: J. H. Bramble, A. Cohen, W. Dahmen, Multiscale<br />

Problems and Methods <strong>in</strong> Numerical Simulations,<br />

Mart<strong>in</strong>a Franca, Italy 2001. Editor: C. Canuto (2003)<br />

Vol. 1826: K. Dohmen, Improved Bonferroni Inequalities<br />

via Abstract Tubes. Inequalities and Identities of<br />

Inclusion-Exclusion Type. VIII, 113 p, 2003.<br />

Vol. 1827: K. M. Pilgrim, Comb<strong>in</strong>ations of Complex<br />

Dynamical Systems. IX, 118 p, 2003.<br />

Vol. 1828: D. J. Green, Gröbner Bases and the Computation<br />

of Group Cohomology. XII, 138 p, 2003.<br />

Vol. 1829: E. Altman, B. Gaujal, A. Hordijk, Discrete-<br />

Event Control of Stochastic Networks: Multimodularity<br />

and Regularity. XIV, 313 p, 2003.<br />

Vol. 1830: M. I. Gil’, Operator Functions and Localization<br />

of Spectra. XIV, 256 p, 2003.<br />

Vol. 1831: A. Connes, J. Cuntz, E. Guentner, N. Higson,<br />

J. E. Kam<strong>in</strong>ker, Noncommutative Geometry, Mart<strong>in</strong>a<br />

Franca, Italy 2002. Editors: S. Doplicher, L. Longo<br />

(2004)<br />

Vol. 1832: J. Azéma, M. Émery, M. Ledoux, M. Yor<br />

(Eds.), Sém<strong>in</strong>aire de Probabilités XXXVII (2003)<br />

Vol. 1833: D.-Q. Jiang, M. Qian, M.-P. Qian, Mathematical<br />

<strong>Theory</strong> of Nonequilibrium Steady States. On the Frontier<br />

of <strong>Probability</strong> and Dynamical Systems. IX, 280 p,<br />

2004.<br />

Vol. 1834: Yo. Yomd<strong>in</strong>, G. Comte, Tame Geometry with<br />

Application <strong>in</strong> Smooth Analysis. VIII, 186 p, 2004.<br />

Vol. 1835: O.T. Izhbold<strong>in</strong>, B. Kahn, N.A. Karpenko,<br />

A. Vishik, Geometric Methods <strong>in</strong> the Algebraic <strong>Theory</strong><br />

of Quadratic Forms. Summer School, Lens, 2000. Editor:<br />

J.-P. Tignol (2004)<br />

Vol. 1836: C.Nǎstǎsescu, F. Van Oystaeyen, Methods of<br />

Graded R<strong>in</strong>gs. XIII, 304 p, 2004.<br />

Vol. 1837: S. Tavaré, O. Zeitouni, Lectures on <strong>Probability</strong><br />

<strong>Theory</strong> and Statistics. Ecole d’Eté de Probabilités de<br />

Sa<strong>in</strong>t-Flour XXXI-2001. Editor: J. Picard (2004)<br />

Vol. 1838: A.J. Ganesh, N.W. O’Connell, D.J. Wischik,<br />

Big Queues. XII, 254 p, 2004.<br />

Vol. 1839: R. Gohm, Noncommutative Stationary<br />

Processes. VIII, 170 p, 2004.<br />

Vol. 1840: B. Tsirelson, W. Werner, Lectures on <strong>Probability</strong><br />

<strong>Theory</strong> and Statistics. Ecole d’Eté de Probabilités de<br />

Sa<strong>in</strong>t-Flour XXXII-2002. Editor: J. Picard (2004)<br />

Vol. 1841: W. Reichel, Uniqueness Theorems for Variational<br />

Problems by the Method of Transformation<br />

Groups (2004)<br />

Vol. 1842: T. Johnsen, A. L. Knutsen, K3 Projective Models<br />

<strong>in</strong> Scrolls (2004)<br />

Vol. 1843: B. Jefferies, Spectral Properties of Noncommut<strong>in</strong>g<br />

Operators (2004)<br />

Vol. 1844: K.F. Siburg, The Pr<strong>in</strong>ciple of Least Action <strong>in</strong><br />

Geometry and Dynamics (2004)<br />

Vol. 1845: M<strong>in</strong> Ho Lee, Mixed Automorphic Forms,<br />

Torus Bundles, and Jacobi Forms (2004)<br />

Vol. 1846: H. Ammari, H. Kang, Reconstruction of Small<br />

Inhomogeneities from Boundary Measurements (2004)<br />

Vol. 1847: T.R. Bielecki, T. Björk, M. Jeanblanc, M.<br />

Rutkowski, J.A. Sche<strong>in</strong>kman, W. Xiong, Paris-Pr<strong>in</strong>ceton<br />

Lectures on Mathematical F<strong>in</strong>ance 2003 (2004)<br />

Vol. 1848: M. Abate, J. E. Fornaess, X. Huang, J. P.<br />

Rosay, A. Tumanov, Real Methods <strong>in</strong> Complex and CR<br />

Geometry, Mart<strong>in</strong>a Franca, Italy 2002. Editors: D. Zaitsev,<br />

G. Zampieri (2004)<br />

Vol. 1849: Mart<strong>in</strong> L. Brown, Heegner Modules and Elliptic<br />

Curves (2004)<br />

Vol. 1850: V. D. Milman, G. Schechtman (Eds.), Geometric<br />

Aspects of Functional Analysis. Israel Sem<strong>in</strong>ar 2002-<br />

2003 (2004)<br />

Vol. 1851: O. Catoni, Statistical Learn<strong>in</strong>g <strong>Theory</strong> and<br />

Stochastic Optimization (2004)<br />

Vol. 1852: A.S. Kechris, B.D. Miller, Topics <strong>in</strong> Orbit<br />

Equivalence (2004)<br />

Vol. 1853: Ch. Favre, M. Jonsson, The Valuative Tree<br />

(2004)<br />

Vol. 1854: O. Saeki, Topology of S<strong>in</strong>gular Fibers of Differential<br />

Maps (2004)<br />

Vol. 1855: G. Da Prato, P.C. Kunstmann, I. Lasiecka,<br />

A. Lunardi, R. Schnaubelt, L. Weis, Functional Analytic<br />

Methods for Evolution Equations. Editors: M. Iannelli,<br />

R. Nagel, S. Piazzera (2004)<br />

Vol. 1856: K. Back, T.R. Bielecki, C. Hipp, S. Peng,<br />

W. Schachermayer, Stochastic Methods <strong>in</strong> F<strong>in</strong>ance, Bressanone/Brixen,<br />

Italy, 2003. Editors: M. Fritelli, W. Runggaldier<br />

(2004)<br />

Vol. 1857: M. Émery, M. Ledoux, M. Yor (Eds.), Sém<strong>in</strong>aire<br />

de Probabilités XXXVIII (2005)<br />

Vol. 1858: A.S. Cherny, H.-J. Engelbert, S<strong>in</strong>gular Stochastic<br />

Differential Equations (2005)<br />

Vol. 1859: E. Letellier, Fourier Transforms of Invariant<br />

Functions on F<strong>in</strong>ite Reductive Lie Algebras (2005)<br />

Vol. 1860: A. Borisyuk, G.B. Ermentrout, A. Friedman,<br />

D. Terman, Tutorials <strong>in</strong> Mathematical Biosciences I.<br />

Mathematical Neurosciences (2005)<br />

Vol. 1861: G. Benett<strong>in</strong>, J. Henrard, S. Kuks<strong>in</strong>, Hamiltonian<br />

Dynamics – <strong>Theory</strong> and Applications, Cetraro,<br />

Italy, 1999. Editor: A. Giorgilli (2005)<br />

Vol. 1862: B. Helffer, F. Nier, Hypoelliptic Estimates and<br />

Spectral <strong>Theory</strong> for Fokker-Planck Operators and Witten<br />

Laplacians (2005)<br />

Vol. 1863: H. Führ, Abstract Harmonic Analysis of Cont<strong>in</strong>uous<br />

Wavelet Transforms (2005)<br />

Vol. 1864: K. Efstathiou, Metamorphoses of Hamiltonian<br />

Systems with Symmetries (2005)<br />

Vol. 1865: D. Applebaum, B.V. R. Bhat, J. Kustermans,<br />

J. M. L<strong>in</strong>dsay, Quantum Independent Increment<br />

Processes I. From <strong>Classical</strong> <strong>Probability</strong> to Quantum<br />

Stochastic Calculus. Editors: M. Schürmann, U. Franz<br />

(2005)<br />

Vol. 1866: O.E. Barndorff-Nielsen, U. Franz, R. Gohm,<br />

B. Kümmerer, S. Thorbjønsen, Quantum Independent<br />

Increment Processes II. Structure of Quantum Lévy<br />

Processes, <strong>Classical</strong> <strong>Probability</strong>, and Physics. Editors: M.<br />

Schürmann, U. Franz, (2005)


Vol. 1867: J. Sneyd (Ed.), Tutorials <strong>in</strong> Mathematical Biosciences<br />

II. Mathematical Model<strong>in</strong>g of Calcium Dynamics<br />

and Signal Transduction. (2005)<br />

Vol. 1868: J. Jorgenson, S. Lang, Posn(R) and Eisenste<strong>in</strong><br />

Series. (2005)<br />

Vol. 1869: A. Dembo, T. Funaki, Lectures on <strong>Probability</strong><br />

<strong>Theory</strong> and Statistics. Ecole d’Eté de Probabilités de<br />

Sa<strong>in</strong>t-Flour XXXIII-2003. Editor: J. Picard (2005)<br />

Vol. 1870: V.I. Gurariy, W. Lusky, Geometry of Müntz<br />

Spaces and Related Questions. (2005)<br />

Vol. 1871: P. Constant<strong>in</strong>, G. Gallavotti, A.V. Kazhikhov,<br />

Y. Meyer, S. Ukai, Mathematical Foundation of Turbulent<br />

Viscous Flows, Mart<strong>in</strong>a Franca, Italy, 2003. Editors:<br />

M. Cannone, T. Miyakawa (2006)<br />

Vol. 1872: A. Friedman (Ed.), Tutorials <strong>in</strong> Mathematical<br />

Biosciences III. Cell Cycle, Proliferation, and Cancer<br />

(2006)<br />

Vol. 1873: R. Mansuy, M. Yor, Random Times and Enlargements<br />

of Filtrations <strong>in</strong> a Brownian Sett<strong>in</strong>g (2006)<br />

Vol. 1874: M. Yor, M. Émery (Eds.), In Memoriam Paul-<br />

André Meyer - Sém<strong>in</strong>aire de Probabilités XXXIX (2006)<br />

Vol. 1875: J. Pitman, Comb<strong>in</strong>atorial Stochastic Processes.<br />

Ecole d’Eté de Probabilités de Sa<strong>in</strong>t-Flour XXXII-2002.<br />

Editor: J. Picard (2006)<br />

Vol. 1876: H. Herrlich, Axiom of Choice (2006)<br />

Vol. 1877: J. Steud<strong>in</strong>g, Value Distributions of L-Functions<br />

(2007)<br />

Vol. 1878: R. Cerf, The Wulff Crystal <strong>in</strong> Is<strong>in</strong>g and Percolation<br />

Models, Ecole d’Eté de Probabilités de Sa<strong>in</strong>t-Flour<br />

XXXIV-2004. Editor: Jean Picard (2006)<br />

Vol. 1879: G. Slade, The Lace Expansion and its Applications,<br />

Ecole d’Eté de Probabilités de Sa<strong>in</strong>t-Flour XXXIV-<br />

2004. Editor: Jean Picard (2006)<br />

Vol. 1880: S. Attal, A. Joye, C.-A. Pillet, Open Quantum<br />

Systems I, The Hamiltonian Approach (2006)<br />

Vol. 1881: S. Attal, A. Joye, C.-A. Pillet, Open Quantum<br />

Systems II, The Markovian Approach (2006)<br />

Vol. 1882: S. Attal, A. Joye, C.-A. Pillet, Open Quantum<br />

Systems III, Recent Developments (2006)<br />

Vol. 1883: W. Van Assche, F. Marcellàn (Eds.), Orthogonal<br />

Polynomials and Special Functions, Computation and<br />

Application (2006)<br />

Vol. 1884: N. Hayashi, E.I. Kaik<strong>in</strong>a, P.I. Naumk<strong>in</strong>,<br />

I.A. Shishmarev, Asymptotics for Dissipative Nonl<strong>in</strong>ear<br />

Equations (2006)<br />

Vol. 1885: A. Telcs, The Art of Random Walks (2006)<br />

Vol. 1886: S. Takamura, Splitt<strong>in</strong>g Deformations of Degenerations<br />

of Complex Curves (2006)<br />

Vol. 1887: K. Habermann, L. Habermann, Introduction to<br />

Symplectic Dirac Operators (2006)<br />

Vol. 1888: J. van der Hoeven, Transseries and Real Differential<br />

Algebra (2006)<br />

Vol. 1889: G. Osipenko, Dynamical Systems, Graphs, and<br />

Algorithms (2006)<br />

Vol. 1890: M. Bunge, J. Funk, S<strong>in</strong>gular Cover<strong>in</strong>gs of<br />

Toposes (2006)<br />

Vol. 1891: J.B. Friedlander, D.R. Heath-Brown,<br />

H. Iwaniec, J. Kaczorowski, Analytic Number <strong>Theory</strong>,<br />

Cetraro, Italy, 2002. Editors: A. Perelli, C. Viola (2006)<br />

Vol. 1892: A. Baddeley, I. Bárány, R. Schneider, W. Weil,<br />

Stochastic Geometry, Mart<strong>in</strong>a Franca, Italy, 2004. Editor:<br />

W. Weil (2007)<br />

Vol. 1893: H. Hanßmann, Local and Semi-Local Bifurcations<br />

<strong>in</strong> Hamiltonian Dynamical Systems, Results and<br />

Examples (2007)<br />

Vol. 1894: C.W. Groetsch, Stable Approximate Evaluation<br />

of Unbounded Operators (2007)<br />

Vol. 1895: L. Molnár, Selected Preserver Problems on<br />

Algebraic Structures of L<strong>in</strong>ear Operators and on Function<br />

Spaces (2007)<br />

Vol. 1896: P. Massart, Concentration Inequalities and<br />

Model Selection, Ecole d’Été de Probabilités de Sa<strong>in</strong>t-<br />

Flour XXXIII-2003. Editor: J. Picard (2007)<br />

Vol. 1897: R. Doney, Fluctuation <strong>Theory</strong> for Lévy<br />

Processes, Ecole d’Été de Probabilités de Sa<strong>in</strong>t-Flour<br />

XXXV-2005. Editor: J. Picard (2007)<br />

Vol. 1898: H.R. Beyer, Beyond Partial Differential Equations,<br />

On l<strong>in</strong>ear and Quasi-L<strong>in</strong>ear Abstract Hyperbolic<br />

Evolution Equations (2007)<br />

Vol. 1899: Sém<strong>in</strong>aire de Probabilités XL. Editors:<br />

C. Donati-Mart<strong>in</strong>, M. Émery, A. Rouault, C. Stricker<br />

(2007)<br />

Vol. 1900: E. Bolthausen, A. Bovier (Eds.), Sp<strong>in</strong> Glasses<br />

(2007)<br />

Vol. 1901: O. Wittenberg, Intersections de deux<br />

quadriques et p<strong>in</strong>ceaux de courbes de genre 1, Intersections<br />

of Two Quadrics and Pencils of Curves of Genus 1<br />

(2007)<br />

Vol. 1902: A. Isaev, Lectures on the Automorphism<br />

Groups of Kobayashi-Hyperbolic Manifolds (2007)<br />

Vol. 1903: G. Kres<strong>in</strong>, V. Maz’ya, Sharp Real-Part Theorems<br />

(2007)<br />

Vol. 1904: P. Giesl, Construction of Global Lyapunov<br />

Functions Us<strong>in</strong>g Radial Basis Functions (2007)<br />

Vol. 1905: C. Prévôt, M. Röckner, A Concise Course on<br />

Stochastic Partial Differential Equations (2007)<br />

Vol. 1906: T. Schuster, The Method of Approximate<br />

Inverse: <strong>Theory</strong> and Applications (2007)<br />

Vol. 1907: M. Rasmussen, Attractivity and Bifurcation<br />

for Nonautonomous Dynamical Systems (2007)<br />

Vol. 1908: T.J. Lyons, M. Caruana, T. Lévy, Differential<br />

Equations Driven by Rough Paths, Ecole d’Été de Probabilités<br />

de Sa<strong>in</strong>t-Flour XXXIV-2004 (2007)<br />

Vol. 1909: H. Akiyoshi, M. Sakuma, M. Wada,<br />

Y. Yamashita, Punctured Torus Groups and 2-Bridge<br />

Knot Groups (I) (2007)<br />

Vol. 1910: V.D. Milman, G. Schechtman (Eds.), Geometric<br />

Aspects of Functional Analysis. Israel Sem<strong>in</strong>ar<br />

2004-2005 (2007)<br />

Vol. 1911: A. Bressan, D. Serre, M. Williams,<br />

K. Zumbrun, Hyperbolic Systems of Balance Laws.<br />

Cetraro, Italy 2003. Editor: P. Marcati (2007)<br />

Vol. 1912: V. Ber<strong>in</strong>de, Iterative Approximation of Fixed<br />

Po<strong>in</strong>ts (2007)<br />

Vol. 1913: J.E. Marsden, G. Misiołek, J.-P. Ortega,<br />

M. Perlmutter, T.S. Ratiu, Hamiltonian Reduction by<br />

Stages (2007)<br />

Vol. 1914: G. Kutyniok, Aff<strong>in</strong>e Density <strong>in</strong> Wavelet<br />

Analysis (2007)<br />

Vol. 1915: T. Bıyıkoˇglu, J. Leydold, P.F. Stadler,<br />

Laplacian Eigenvectors of Graphs. Perron-Frobenius and<br />

Faber-Krahn Type Theorems (2007)<br />

Vol. 1916: C. Villani, F. Rezakhanlou, Entropy Methods<br />

for the Boltzmann Equation. Editors: F. Golse, S. Olla<br />

(2008)<br />

Vol. 1917: I. Veselić, Existence and Regularity Properties<br />

of the Integrated Density of States of Random<br />

Schröd<strong>in</strong>ger (2008)<br />

Vol. 1918: B. Roberts, R. Schmidt, Local Newforms for<br />

GSp(4) (2007)<br />

Vol. 1919: R.A. Carmona, I. Ekeland, A. Kohatsu-<br />

Higa, J.-M. Lasry, P.-L. Lions, H. Pham, E. Tafl<strong>in</strong>,<br />

Paris-Pr<strong>in</strong>ceton Lectures on Mathematical F<strong>in</strong>ance 2004.


Editors: R.A. Carmona, E. Ç<strong>in</strong>lar, I. Ekeland, E. Jou<strong>in</strong>i,<br />

J.A. Sche<strong>in</strong>kman, N. Touzi (2007)<br />

Vol. 1920: S.N. Evans, <strong>Probability</strong> and Real Trees. Ecole<br />

d’Été de Probabilités de Sa<strong>in</strong>t-Flour XXXV-2005 (2008)<br />

Vol. 1921: J.P. Tian, Evolution Algebras and their Applications<br />

(2008)<br />

Vol. 1922: A. Friedman (Ed.), Tutorials <strong>in</strong> Mathematical<br />

BioSciences IV. Evolution and Ecology (2008)<br />

Vol. 1923: J.P.N. Bishwal, Parameter Estimation <strong>in</strong><br />

Stochastic Differential Equations (2008)<br />

Vol. 1924: M. Wilson, Littlewood-Paley <strong>Theory</strong> and<br />

Exponential-Square Integrability (2008)<br />

Vol. 1925: M. du Sautoy, L. Woodward, Zeta Functions<br />

of Groups and R<strong>in</strong>gs (2008)<br />

Vol. 1926: L. Barreira, V. Claudia, Stability of Nonautonomous<br />

Differential Equations (2008)<br />

Vol. 1927: L. Ambrosio, L. Caffarelli, M.G. Crandall,<br />

L.C. Evans, N. Fusco, Calculus of Variations and Non-<br />

L<strong>in</strong>ear Partial Differential Equations. Cetraro, Italy 2005.<br />

Editors: B. Dacorogna, P. Marcell<strong>in</strong>i (2008)<br />

Vol. 1928: J. Jonsson, Simplicial Complexes of Graphs<br />

(2008)<br />

Vol. 1929: Y. Mishura, Stochastic Calculus for Fractional<br />

Brownian Motion and Related Processes (2008)<br />

Vol. 1930: J.M. Urbano, The Method of Intr<strong>in</strong>sic Scal<strong>in</strong>g.<br />

A Systematic Approach to Regularity for Degenerate and<br />

S<strong>in</strong>gular PDEs (2008)<br />

Vol. 1931: M. Cowl<strong>in</strong>g, E. Frenkel, M. Kashiwara,<br />

A. Valette, D.A. Vogan, Jr., N.R. Wallach, Representation<br />

<strong>Theory</strong> and Complex Analysis. Venice, Italy 2004.<br />

Editors: E.C. Tarabusi, A. D’Agnolo, M. Picardello<br />

(2008)<br />

Vol. 1932: A.A. Agrachev, A.S. Morse, E.D. Sontag,<br />

H.J. Sussmann, V.I. Utk<strong>in</strong>, Nonl<strong>in</strong>ear and Optimal Control<br />

<strong>Theory</strong>. Cetraro, Italy 2004. Editors: P. Nistri, G. Stefani<br />

(2008)<br />

Vol. 1933: M. Petkovic, Po<strong>in</strong>t Estimation of Root F<strong>in</strong>d<strong>in</strong>g<br />

Methods (2008)<br />

Vol. 1934: C. Donati-Mart<strong>in</strong>, M. Émery, A. Rouault,<br />

C. Stricker (Eds.), Sém<strong>in</strong>aire de Probabilités XLI (2008)<br />

Vol. 1935: A. Unterberger, Alternative Pseudodifferential<br />

Analysis (2008)<br />

Vol. 1936: P. Magal, S. Ruan (Eds.), Structured Population<br />

Models <strong>in</strong> Biology and Epidemiology (2008)<br />

Vol. 1937: G. Capriz, P. Giov<strong>in</strong>e, P.M. Mariano (Eds.),<br />

Mathematical Models of Granular Matter (2008)<br />

Vol. 1938: D. Auroux, F. Catanese, M. Manetti, P. Seidel,<br />

B. Siebert, I. Smith, G. Tian, Symplectic 4-Manifolds<br />

and Algebraic Surfaces. Cetraro, Italy 2003. Editors:<br />

F. Catanese, G. Tian (2008)<br />

Vol. 1939: D. Boffi, F. Brezzi, L. Demkowicz, R.G.<br />

Durán, R.S. Falk, M. Fort<strong>in</strong>, Mixed F<strong>in</strong>ite Elements,<br />

Compatibility Conditions, and Applications. Cetraro,<br />

Italy 2006. Editors: D. Boffi, L. Gastaldi (2008)<br />

Vol. 1940: J. Banasiak, V. Capasso, M.A.J. Chapla<strong>in</strong>,<br />

M. Lachowicz, J. Mie¸kisz, Multiscale Problems <strong>in</strong><br />

the Life Sciences. From Microscopic to Macroscopic.<br />

Be¸dlewo, Poland 2006. Editors: V. Capasso, M. Lachowicz<br />

(2008)<br />

Vol. 1941: S.M.J. Haran, Arithmetical Investigations.<br />

Representation <strong>Theory</strong>, Orthogonal Polynomials, and<br />

Quantum Interpolations (2008)<br />

Vol. 1942: S. Albeverio, F. Flandoli, Y.G. S<strong>in</strong>ai, SPDE <strong>in</strong><br />

Hydrodynamic. Recent Progress and Prospects. Cetraro,<br />

Italy 2005. Editors: G. Da Prato, M. Röckner (2008)<br />

Vol. 1943: L.L. Bonilla (Ed.), Inverse Problems and Imag<strong>in</strong>g.<br />

Mart<strong>in</strong>a Franca, Italy 2002 (2008)<br />

Vol. 1944: A. Di Bartolo, G. Falcone, P. Plaumann,<br />

K. Strambach, Algebraic Groups and Lie Groups with<br />

Few Factors (2008)<br />

Vol. 1945: F. Brauer, P. van den Driessche, J. Wu (Eds.),<br />

Mathematical Epidemiology (2008)<br />

Vol. 1946: G. Allaire, A. Arnold, P. Degond, T.Y. Hou,<br />

Quantum Transport. Modell<strong>in</strong>g, Analysis and Asymptotics.<br />

Cetraro, Italy 2006. Editors: N.B. Abdallah, G. Frosali<br />

(2008)<br />

Vol. 1947: D. Abramovich, M. Mariño, M. Thaddeus,<br />

R. Vakil, Enumerative Invariants <strong>in</strong> Algebraic Geometry<br />

and Str<strong>in</strong>g <strong>Theory</strong>. Cetraro, Italy 2005. Editors:<br />

K. Behrend, M. Manetti (2008)<br />

Vol. 1948: F. Cao, J-L. Lisani, J-M. Morel, P. Musé,<br />

F. Sur, A <strong>Theory</strong> of Shape Identification (2008)<br />

Vol. 1949: H.G. Feicht<strong>in</strong>ger, B. Helffer, M.P. Lamoureux,<br />

N. Lerner, J. Toft, Pseudo-Differential Operators. Quantization<br />

and Signals. Cetraro, Italy 2006. Editors: L.<br />

Rod<strong>in</strong>o, M.W. Wong (2008)<br />

Vol. 1950: M. Bramson, Stability of Queue<strong>in</strong>g Networks,<br />

Ecole d’Eté de Probabilités de Sa<strong>in</strong>t-Flour XXXVI-2006<br />

(2008)<br />

Vol. 1951: A. Moltó, J. Orihuela, S. Troyanski,<br />

M. Valdivia, A Non L<strong>in</strong>ear Transfer Technique for<br />

Renorm<strong>in</strong>g (2008)<br />

Vol. 1952: R. Mikhailov, I.B.S. Passi, Lower Central and<br />

Dimension Series of Groups (2008)<br />

Vol. 1953: K. Arw<strong>in</strong>i, C.T.J. Dodson, Information Geometry<br />

(2008)<br />

Vol. 1954: P. Biane, L. Bouten, F. Cipriani, N. Konno,<br />

N. Privault, Q. Xu, Quantum <strong>Potential</strong> <strong>Theory</strong>. Editors:<br />

U. Franz, M. Schürmann (2008)<br />

Vol. 1955: M. Bernot, V. Caselles, J.-M. Morel, Optimal<br />

transportation networks (2008)<br />

Vol. 1956: C.H. Chu, Matrix Convolution Operators on<br />

Groups (2008)<br />

Vol. 1957: A. Guionnet, On Random Matrices: Macroscopic<br />

Asymptotics, Ecole d’Eté de Probabilités de Sa<strong>in</strong>t-<br />

Flour XXXVI-2006 (2008)<br />

Vol. 1958: M.C. Olsson, Compactify<strong>in</strong>g Moduli Spaces<br />

for Abelian Varieties (2008)<br />

Recent Repr<strong>in</strong>ts and New Editions<br />

Vol. 1702: J. Ma, J. Yong, Forward-Backward Stochastic<br />

Differential Equations and their Applications. 1999 –<br />

Corr. 3rd pr<strong>in</strong>t<strong>in</strong>g (2007)<br />

Vol. 830: J.A. Green, Polynomial Representations of<br />

GLn, with an Appendix on Schensted Correspondence<br />

and Littelmann Paths by K. Erdmann, J.A. Green and<br />

M. Schoker 1980 – 2nd corr. and augmented edition<br />

(2007)<br />

Vol. 1693: S. Simons, From Hahn-Banach to Monotonicity<br />

(M<strong>in</strong>imax and Monotonicity 1998) – 2nd exp. edition<br />

(2008)<br />

Vol. 470: R.E. Bowen, Equilibrium States and the Ergodic<br />

<strong>Theory</strong> of Anosov Diffeomorphisms. With a preface by<br />

D. Ruelle. Edited by J.-R. Chazottes. 1975 – 2nd rev.<br />

edition (2008)<br />

Vol. 523: S.A. Albeverio, R.J. Høegh-Krohn, S. Mazzucchi,<br />

Mathematical <strong>Theory</strong> of Feynman Path Integral.<br />

1976 – 2nd corr. and enlarged edition (2008)<br />

Vol. 1764: A. Cannas da Silva, Lectures on Symplectic<br />

Geometry 2001 – Corr. 2nd pr<strong>in</strong>t<strong>in</strong>g (2008)


LECTURE NOTES IN MATHEMATICS<br />

Edited by J.-M. Morel, F. Takens, B. Teissier, P.K. Ma<strong>in</strong>i<br />

123<br />

Editorial Policy (for Multi-Author Publications: Summer Schools/Intensive Courses)<br />

1. Lecture Notes aim to report new developments <strong>in</strong> all areas of mathematics and their<br />

applications - quickly, <strong>in</strong>formally and at a high level. Mathematical texts analys<strong>in</strong>g new<br />

developments <strong>in</strong> modell<strong>in</strong>g and numerical simulation are welcome. Manuscripts should be<br />

reasonably self-conta<strong>in</strong>ed and rounded off. Thus they may, and often will, present not only<br />

results of the author but also related work by other people. They should provide sufficient<br />

motivation, examples and applications. There should also be an <strong>in</strong>troduction mak<strong>in</strong>g the<br />

text comprehensible to a wider audience. This clearly dist<strong>in</strong>guishes Lecture Notes from<br />

journal articles or technical reports which normally are very concise. Articles <strong>in</strong>tended for<br />

a journal but too long to be accepted by most journals, usually do not have this “lecture<br />

notes” character.<br />

2. In general SUMMER SCHOOLS and other similar INTENSIVE COURSES are held to<br />

present mathematical topics that are close to the frontiers of recent research to an audience<br />

at the beg<strong>in</strong>n<strong>in</strong>g or <strong>in</strong>termediate graduate level, who may want to cont<strong>in</strong>ue with this area of<br />

work, for a thesis or later. This makes demands on the didactic aspects of the presentation.<br />

Because the subjects of such schools are advanced, there often exists no textbook, and so<br />

ideally, the publication result<strong>in</strong>g from such a school could be a first approximation to such<br />

a textbook. Usually several authors are <strong>in</strong>volved <strong>in</strong> the writ<strong>in</strong>g, so it is not always simple<br />

to obta<strong>in</strong> a unified approach to the presentation.<br />

For prospective publication <strong>in</strong> LNM, the result<strong>in</strong>g manuscript should not be just a collection<br />

of course notes, each of which has been developed by an <strong>in</strong>dividual author with<br />

little or no co-ord<strong>in</strong>ation with the others, and with little or no common concept. The subject<br />

matter should dictate the structure of the book, and the authorship of each part or<br />

chapter should take secondary importance. Of course the choice of authors is crucial to the<br />

quality of the material at the school and <strong>in</strong> the book, and the <strong>in</strong>tention here is not to belittle<br />

their impact, but simply to say that the book should be planned to be written by these<br />

authors jo<strong>in</strong>tly, and not just assembled as a result of what these authors happen to submit.<br />

This represents considerable preparatory work (as it is imperative to ensure that the<br />

authors know these criteria before they <strong>in</strong>vest work on a manuscript), and also considerable<br />

edit<strong>in</strong>g work afterwards, to get the book <strong>in</strong>to f<strong>in</strong>al shape. Still it is the form that holds<br />

the most promise of a successful book that will be used by its <strong>in</strong>tended audience, rather<br />

than yet another volume of proceed<strong>in</strong>gs for the library shelf.<br />

3. Manuscripts should be submitted either to Spr<strong>in</strong>ger’s mathematics editorial <strong>in</strong> Heidelberg,<br />

or to one of the series editors. Volume editors are expected to arrange for the referee<strong>in</strong>g, to<br />

the usual scientific standards, of the <strong>in</strong>dividual contributions. If the result<strong>in</strong>g reports can be<br />

forwarded to us (series editors or Spr<strong>in</strong>ger) this is very helpful. If no reports are forwarded<br />

or if other questions rema<strong>in</strong> unclear <strong>in</strong> respect of homogeneity etc, the series editors may<br />

wish to consult external referees for an overall evaluation of the volume. A f<strong>in</strong>al decision<br />

to publish can be made only on the basis of the complete manuscript; however a prelim<strong>in</strong>ary<br />

decision can be based on a pre-f<strong>in</strong>al or <strong>in</strong>complete manuscript. The strict m<strong>in</strong>imum<br />

amount of material that will be considered should <strong>in</strong>clude a detailed outl<strong>in</strong>e describ<strong>in</strong>g the<br />

planned contents of each chapter.<br />

Volume editors and authors should be aware that <strong>in</strong>complete or <strong>in</strong>sufficiently close to<br />

f<strong>in</strong>al manuscripts almost always result <strong>in</strong> longer evaluation times. They should also be<br />

aware that parallel submission of their manuscript to another publisher while under consideration<br />

for LNM will <strong>in</strong> general lead to immediate rejection.


4. Manuscripts should <strong>in</strong> general be submitted <strong>in</strong> English. F<strong>in</strong>al manuscripts should conta<strong>in</strong><br />

at least 100 pages of mathematical text and should always <strong>in</strong>clude<br />

– a general table of contents;<br />

– an <strong>in</strong>formative <strong>in</strong>troduction, with adequate motivation and perhaps some historical<br />

remarks: it should be accessible to a reader not <strong>in</strong>timately familiar with the topic<br />

treated;<br />

– a global subject <strong>in</strong>dex: as a rule this is genu<strong>in</strong>ely helpful for the reader.<br />

Lecture Notes volumes are, as a rule, pr<strong>in</strong>ted digitally from the authors’ files. We strongly<br />

recommend that all contributions <strong>in</strong> a volume be written <strong>in</strong> LaTeX2e. To ensure best<br />

results, authors are asked to use the LaTeX2e style files available from Spr<strong>in</strong>ger’s webserver<br />

at<br />

ftp://ftp.spr<strong>in</strong>ger.de/pub/tex/latex/svmultt1/ (for summer schools/tutorials).<br />

Additional technical <strong>in</strong>structions are available on request from: lnm@spr<strong>in</strong>ger.com.<br />

5. Careful preparation of the manuscripts will help keep production time short besides<br />

ensur<strong>in</strong>g satisfactory appearance of the f<strong>in</strong>ished book <strong>in</strong> pr<strong>in</strong>t and onl<strong>in</strong>e. After acceptance<br />

of the manuscript authors will be asked to prepare the f<strong>in</strong>al LaTeX source<br />

files (and also the correspond<strong>in</strong>g dvi-, pdf- or zipped ps-file) together with the f<strong>in</strong>al<br />

pr<strong>in</strong>tout made from these files. The LaTeX source files are essential for produc<strong>in</strong>g<br />

the full-text onl<strong>in</strong>e version of the book. For the exist<strong>in</strong>g onl<strong>in</strong>e volumes of LNM see:<br />

www.spr<strong>in</strong>gerl<strong>in</strong>k.com/content/110312<br />

The actual production of a Lecture Notes volume takes approximately 12 weeks.<br />

6. Volume editors receive a total of 50 free copies of their volume to be shared with the authors,<br />

but no royalties. They and the authors are entitled to a discount of 33.3% on the price<br />

of Spr<strong>in</strong>ger books purchased for their personal use, if order<strong>in</strong>g directly from Spr<strong>in</strong>ger.<br />

7. Commitment to publish is made by letter of <strong>in</strong>tent rather than by sign<strong>in</strong>g a formal contract.<br />

Spr<strong>in</strong>ger-Verlag secures the copyright for each volume. Authors are free to reuse material<br />

conta<strong>in</strong>ed <strong>in</strong> their LNM volumes <strong>in</strong> later publications: a brief written (or e-mail) request<br />

for formal permission is sufficient.<br />

Addresses:<br />

Professor J.-M. Morel, CMLA,<br />

École Normale Supérieure de Cachan,<br />

61 Avenue du Président Wilson,<br />

94235 Cachan Cedex, France<br />

E-mail: Jean-Michel.Morel@cmla.ens-cachan.fr<br />

Professor F. Takens, Mathematisch Instituut,<br />

Rijksuniversiteit Gron<strong>in</strong>gen, Postbus 800,<br />

9700 AV Gron<strong>in</strong>gen, The Netherlands<br />

E-mail: F.Takens@math.rug.nl<br />

For the “Mathematical Biosciences Subseries” of LNM:<br />

Professor P.K. Ma<strong>in</strong>i, Center for Mathematical Biology<br />

Mathematical Institute, 24-29 St Giles,<br />

Oxford OX1 3LP, UK<br />

E-mail: ma<strong>in</strong>i@maths.ox.ac.uk<br />

Spr<strong>in</strong>ger, Mathematics Editorial I, Tiergartenstr. 17,<br />

69121 Heidelberg, Germany,<br />

Tel.: +49 (6221) 487-8259<br />

Fax: +49 (6221) 4876-8259<br />

E-mail: lnm@spr<strong>in</strong>ger.com<br />

Professor B. Teissier,<br />

Institut Mathématique de Jussieu,<br />

UMR 7586 du CNRS,<br />

Équipe “Géométrie et Dynamique”,<br />

175 rue du Chevaleret<br />

75013 Paris, France<br />

E-mail: teissier@math.jussieu.fr

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!