1: \documentclass[11pt]{amsart}
2: \usepackage{amsbsy,amssymb,amscd,amsfonts,latexsym,amstext,delarray,
3: amsmath,graphicx}
4: \input{amssymb.sty}
5: \usepackage[dvips]{epsfig}
6:
7: %%change height page
8: \setlength{\textwidth}{15.5cm}
9: %%change width page
10: \setlength{\oddsidemargin}{0.5cm} \setlength{\topmargin}{.5cm}
11: \setlength{\evensidemargin}{\oddsidemargin}
12:
13: \input xypic
14:
15:
16:
17:
18: \newtheorem{thm}{Theorem}[section]
19: \newtheorem{prop}[thm]{Proposition}
20: \newtheorem{cor}[thm]{Corollary}
21: \newtheorem{lem}[thm]{Lemma}
22: \newtheorem{defn}[thm]{Definition}
23: \newtheorem{rem}[thm]{Remark}
24: \newtheorem{ex}[thm]{Example}
25:
26: \numberwithin{equation}{section}
27:
28: \def\bA{{\mathbb A}}
29: \def\bB{{\mathbb B}}
30: \def\bC{{\mathbb C}}
31: \def\bD{{\mathbb D}}
32: \def\bE{{\mathbb E}}
33: \def\bF{{\mathbb F}}
34: \def\bG{{\mathbb G}}
35: \def\bH{{\mathbb H}}
36: \def\bI{{\mathbb I}}
37: \def\bJ{{\mathbb J}}
38: \def\bK{{\mathbb K}}
39: \def\bL{{\mathbb L}}
40: \def\bM{{\mathbb M}}
41: \def\bN{{\mathbb N}}
42: \def\bO{{\mathbb O}}
43: \def\bP{{\mathbb P}}
44: \def\bQ{{\mathbb Q}}
45: \def\bR{{\mathbb R}}
46: \def\bS{{\mathbb S}}
47: \def\bT{{\mathbb T}}
48: \def\bU{{\mathbb U}}
49: \def\bV{{\mathbb V}}
50: \def\bW{{\mathbb W}}
51: \def\bX{{\mathbb X}}
52: \def\bY{{\mathbb Y}}
53: \def\bZ{{\mathbb Z}}
54:
55: \def\A{{\mathbb A}}
56: \def\C{{\mathbb C}}
57: \def\R{{\mathbb R}}
58: \def\F{{\mathbb F}}
59: \renewcommand{\H}{{\mathbb H}}
60: \def\K{{\mathbb K}}
61: \def\N{{\mathbb N}}
62: \renewcommand{\P}{{\mathbb P}}
63: \def\Q{{\mathbb Q}}
64: \def\Z{{\mathbb Z}}
65: \def\Zb{{\mathbb Z}}
66: \def\s{\sigma}
67:
68: \def\cal{\mathcal}
69: \def\gr{\mathrm{gr}}
70: \def\oplusinf{\mathop{\oplus}}
71: \def\otimesinf{\mathop{\otimes}}
72: \def\id{{\mbox{Id}}}
73: \def\ch{{\mbox{ch}}}
74: \def\im{{\mbox{Im}}}
75: \def\Aut{{\mbox{Aut}}}
76: \def\dim{{\mbox{dim}}}
77: \def\ker{{\mbox{Ker}}}
78: \def\Der{{\mbox{Der}}}
79: \def\Hom {{\mbox{Hom}}}
80: \def\hom {{\mbox{hom}}}
81: \def\End{{\mbox{End}}}
82: \def\Trace{{\mbox{Trace}}}
83: \def\Int{{\mbox{Int}}}
84: \def\alg{{\mbox{alg}}}
85: \def\Lie{{\mbox{Lie}}}
86: \def\Ad{{\mbox{Ad}}}
87: \def\Spin{{\mbox{Spin}}}
88: \def\wh{\widehat}
89:
90: \def\cali{{\cal I}}
91: \def\calp{{\cal P}}
92: \def\cala{{\cal A}}
93: \def\calb{{\cal B}}
94: \def\calk{{\cal K}}
95: \def\call{{\cal L}}
96: \def\calm{{\cal M}}
97: \def\caln{{\cal N}}
98: \def\calo{{\cal O}}
99: \def\calh{{\cal H}}
100: \def\cals{{\cal S}}
101: \def\calc{{\cal C}}
102: \def\cale{{\cal E}}
103: \def\calu{{\cal U}}
104: \def\calf{{\cal F}}
105: \def\calr{{\cal R}}
106: \def\calV{{\cal V}}
107: \def\caly{{\cal Y}}
108: \def\fraca{{\mathfrak A}}
109: \def\fraci{{\mathfrak I}}
110: \def\fracg{{\mathfrak g}}
111: \def\frach{{\mathfrak h}}
112: \def\fracd{{\mathfrak d}}
113: \def\fracS{{\mathfrak {S}}}
114: \def\fracd{{\mathfrak d}}
115: \def\fracsu{{\mathfrak {su}}}
116: \def\fracsl{{\mathfrak {sl}}}
117: \def\fracso{{\mathfrak {so}}}
118: \def\calg{{C_{\mathrm{alg}}}}
119: \def\bbbone{\mbox{\rm 1\hspace {-.6em} l}}
120: \def\ue{\underline{E}}
121: \def\uf{\underline{F}}
122: \def\uk{\underline{K}}
123: \def\ux{\underline{X}}
124: \def\gg{{\mathbf g}}
125: \def\ag{{\mathbf a}}
126: \def\ug{{\mathbf u}}
127: \def\mg{{\mathbf m}}
128: \def\vg{{\mathbf v}}
129: \def\ung{{\mathbf 1}}
130: \def\Yg{{\mathbf Y}}
131: \def\Tg{{\mathbf T}}
132: \def\Lg{{\mathbf L}}
133: \def\Rg{{\mathbf R}}
134: \def\vg{{\mathbf v}}
135: \def\hg{{\mathbf H}}
136: %\def\alg{{\mathbf{Alg}}}
137: \def\vect{{\mathbf{Vect}}}
138: \def\Homg {{\mathbf{Hom}}}
139: \def\homg {{\mathbf{hom}}}
140: \def\endg{{\mathbf{end}}}
141:
142:
143: \def\af{{\mathfrak a}}
144: \newcommand{\fg}{{\mathfrak{g}}}
145: \newcommand{\fh}{{\mathfrak{h}}}
146: \def\m{{\mathfrak m}}
147: \def\n{{\mathfrak n}}
148: \def\p{{\mathfrak p}}
149: \def\fU{{\mathfrak U}}
150:
151: \def\cutint{{\int \!\!\!\!\!\! -}}
152: \def\qqq{{\,,\quad \forall}}
153: \def\cA{{\mathcal A}}
154: \def\cB{{\mathcal B}}
155: \def\cC{{\mathcal C}}
156: \def\cD{{\mathcal D}}
157: \def\cE{{\mathcal E}}
158: \def\cF{{\mathcal F}}
159: \def\cG{{\mathcal G}}
160: \def\cH{{\mathcal H}}
161: \def\cI{{\mathcal I}}
162: \def\cJ{{\mathcal J}}
163: \def\cK{{\mathcal K}}
164: \def\cL{{\mathcal L}}
165: \def\cM{{\mathcal M}}
166: \def\cN{{\mathcal N}}
167: \def\cO{{\mathcal O}}
168: \def\cP{{\mathcal P}}
169: \def\cQ{{\mathcal Q}}
170: \def\cR{{\mathcal R}}
171: \def\cS{{\mathcal S}}
172: \def\cT{{\mathcal T}}
173: \def\cU{{\mathcal U}}
174: \def\cV{{\mathcal V}}
175: \def\cX{{\mathcal X}}
176: \def\cY{{\mathcal Y}}
177: \def\cZ{{\mathcal Z}}
178:
179: \newcommand{\ie}{{\it i.e.\/}\ }
180: \newcommand{\eg}{{\it e.g.\/}\ }
181: \newcommand{\cf}{{\it cf.\/}\ }
182: \newcommand{\opcit}{{\it op.cit.\/}\ }
183: \newcommand{\resp}{{\it resp.\/}\ }
184: \def\text{\hbox}
185:
186: \def\Aut{{\rm Aut}}
187: \def\Coker{{\rm Coker}}
188: \def\dist{{\rm dist}}
189: \def\Dom{{\rm Dom}}
190: \def\End{{\rm End}}
191: \def\Ext{{\rm Ext}}
192: \def\Gal{{\rm Gal}}
193: \def\GL{{\rm GL}}
194: \def\Gr{{\rm Gr}}
195: \def\Hom{{\rm Hom}}
196: \def\id{{\rm id}}
197: \def\Ind{{\rm Ind}}
198: \def\Index{{\rm Ind}}
199: \def\Ker{{\rm Ker}}
200: \def\Lie{{\rm Lie}}
201: \def\PGL{{\rm PGL}}
202: \def\PSL{{\rm PSL}}
203: \def\rank{{\rm rank}}
204: \def\Res{{\rm Res}}
205: \def\sign{{\rm sign}}
206: \def\SL{{\rm SL}}
207: \def\Spec{{\rm Spec}}
208: \def\Sp{{\rm Spec}}
209: \def\spin{{\rm spin}}
210: \def\Spin{{\rm Spin}}
211: \def\Trace{{\rm Tr}}
212: \def\Tr{{\rm Tr}}
213: \def\tr{{\rm tr}}
214:
215: \parindent 0in
216:
217: \title[The noncommutative ``Pardis'']{A walk in the noncommutative garden}
218: \author[Connes]{Alain Connes}
219: \author[Marcolli]{Matilde Marcolli}
220: \address{A.~Connes: Coll\`ege de France \\
221: 3, rue d'Ulm \\ Paris, F-75005 France
222: \\ I.H.E.S. and Vanderbilt
223: University} \email{alain\@@connes.org}
224: \address{M.~Marcolli: Max--Planck Institut f\"ur Mathematik \\
225: Vivatsgasse 7 \\
226: Bonn, D-53111 Germany} \email{marcolli\@@mpim-bonn.mpg.de}
227:
228: \begin{document}
229: \maketitle
230:
231: \tableofcontents
232:
233: \newpage
234:
235: \section{Introduction}\label{intro}
236:
237: \begin{verse}
238: {\em If you cleave the hearth of one drop of water \\
239: a hundred pure oceans emerge from it.} \\
240: \smallskip
241: (Mahmud Shabistari, {\em Gulshan-i-raz})
242: \end{verse}
243:
244: \bigskip
245:
246: We have decided to contribute to the volume of the IPM lectures on
247: noncommutative geometry a text that collects a list of examples of
248: noncommutative spaces. As the quote of the Sufi poet here above suggests,
249: it is often better to approach a new subject by analyzing specific
250: examples rather than presenting the general theory. We hope that the
251: diversity of examples the readers will encounter in this text will
252: suffice to convince them of the fact that noncommutative geometry is
253: a very rich field in rapid evolution, full of interesting and yet
254: unexplored landscapes. Many of the examples collected here have not
255: yet been fully explored from the point of view of the general
256: guidelines we propose in Section \ref{road} and the main point of
257: this text is to provide a great number of open questions. The reader
258: should interpret this survey as a suggestion of possible interesting
259: problems to investigate, both in the settings described here, as
260: well as in other examples that are available but did not fit in this
261: list, and in the many more that still await to be discovered.
262: Besides the existing books on NCG such as \cite{Co90},
263: \cite{Manin91NC}, \cite{Co94}, \cite{khal},
264: \cite{landi} \cite{FGV} \cite{Mar}, two new
265: books are being written: one by the two authors of this paper \cite{CoMarBook},
266: and one by Connes and Moscovici \cite{CoMoBook}.
267:
268:
269: \section{Handling noncommutative spaces in the wild:
270: basic tools}\label{road}
271:
272: We are going to see in many example how one obtains the algebra
273: of coordinates $\cA$ of a noncommutative space $X$.
274: Here we think of $\cA$ as being the algebra
275: of ``smooth functions'', which will usually be
276: a dense subalgebra of a $C^*$-algebra $\bar\cA$.
277:
278: \smallskip
279:
280: Here are some basic steps that one can perform in order to
281: acquire a good understanding of a given noncommutative space $X$
282: with algebra of coordinates $\cA$.
283:
284: \begin{itemize}
285:
286: \item[1)] Resolve the diagonal of $\cA$ and compute the cyclic cohomology.
287:
288: \item[2)] Find a geometric model of $X$ up to homotopy.
289:
290: \item[3)] Construct the spectral geometry $(\cA,\cH,D)$.
291:
292: \item[4)] Compute the time evolution and analyze the thermodynamics.
293:
294: \end{itemize}
295:
296: \medskip
297:
298: 1) The first step means finding a resolution of the $\cA$-bimodule
299: $\cA$ by projective $\cA$-bimodules making it possible to compute
300: the Hochschild homology of $\cA$ effectively. In general, such
301: resolutions will be of Kozsul type and the typical example is the
302: resolution of the diagonal for the algebra $C^\infty(X)$ of smooth
303: functions on a compact manifold as in the $C^\infty$ version
304: \cite{ConnesCH} of the Hochschild, Kostant, Rosenberg theorem (\cf
305: \cite{HKR}). It makes it possible to know what is the analogue of
306: differential forms and of de Rham currents on the space $X$ and to
307: take the next step of computing the cyclic homology and cyclic
308: cohomology of $\cA$, which are the natural replacements for the de
309: Rham theory. For foliation algebras
310: this was done long ago (\cf \cite{Co1}, \cite{BryNi}, \cite{Crai}).
311: It ties in with the natural double complex of transverse currents.
312:
313: It is not always easy to perform this step of finding
314: a resolution and computing Hochschild and cyclic (co)homology.
315: For instance, in the case of algebras given by generators and
316: relations this uses the whole theory of Kozsul duality,
317: which has been successfully extended to $N$-homogeneous algebras (\cf
318: \cite{D-V}, \cite{D-V1}, \cite{D-V2}, \cite{D-V3}, \cite{D-V4},
319: \cite{BDV}).
320:
321: One specific example in which it would be very interesting to
322: resolve the diagonal is the modular Hecke algebras (Section
323: \ref{hecke}). In essence, finding a resolution of the diagonal in
324: the algebra of modular forms of arbitrary level, equivariant with
325: respect to the action of the group $\GL_2(\A_f)$ of finite adeles,
326: would yield formulas for the compatibility of Hecke operators with
327: the algebra structure. This is a basic and hard problem of the
328: theory of modular forms.
329:
330: \medskip
331:
332: Cyclic cohomology (and homology) is a well developed theory which
333: was first designed to handle the leaf spaces of foliations as well
334: as group rings of discrete groups (\cf \cite{khal}). The theory
335: admits a purely algebraic version which is at center stage in
336: ``algebraic" noncommutative geometry, but it is crucial in the
337: analytic set-up to construct cyclic cocycles with good compatibility
338: properties with the topology of the algebra. For instance, when the
339: domain of definition of the cocycle is a dense subalgebra stable
340: under holomorphic functional calculus, it automatically gives an
341: invariant of the $K$-theory of the underlying $C^*$-algebra (\cf
342: \cite{Co1}).
343:
344:
345: \smallskip
346:
347: 2) The essence of the second step is that many noncommutative spaces
348: defined as a ``bad quotients'' (\cf Section \ref{quotients})
349: can be desingularized, provided one is
350: ready to work up to homotopy. Thus for instance if the space $X$ is
351: defined as the quotient
352: $$
353: X=\,Y/\,\sim
354: $$
355: of an ``ordinary" space $Y$ by an equivalence relation $\sim$ one
356: can often find a description of the same space $X$ as a quotient
357: $$
358: X=\,Z/\,\sim
359: $$
360: where the equivalence classes are now {\em contractible} spaces. The
361: homotopy type of $Z$ is then uniquely determined and serves as a
362: substitute for that of $X$ (see \cite{BauCo}).
363:
364: For instance, if the equivalence relation on $Y$ comes from the free
365: action of a torsion free discrete group $\Gamma$, the space $Z$ is
366: simply a product over $\Gamma$ of the form
367: $$
368: Z=\,Y\times_\Gamma\,E\Gamma ,
369: $$
370: where $E\Gamma$ is a contractible space on which $\Gamma$ acts
371: freely and properly.
372:
373: The main point of this second step is that it gives a starting point
374: for computing the $K$-theory of the space $X$ \ie of the
375: $C^*$-algebra $A=\bar\cA$ playing the role of the algebra of continuous
376: functions on $X$. Indeed, for each element of the $K$-homology
377: of the classifying space $Z$, there is a general construction
378: of an index problem for ``families
379: parameterized by $X$" that yields an assembly map (\cf \cite{BauCo})
380: \begin{equation}\label{mu-map}
381: \mu\;:\; K_*(Z)\to K_*(A)
382: \end{equation}
383: This Baum--Connes map is an isomorphism in a lot of cases (with
384: suitable care of torsion, \cf \cite{BCH}) including all
385: connected locally compact groups, all amenable groupoids
386: and all hyperbolic discrete groups.
387: It thus gives a computable guess for the $K$-theory of $X$.
388:
389: The next
390: step is not only to really compute $K(A)$ but also to get a good
391: model for the ``vector bundles" on $X$ \ie the finite projective
392: modules over $A$. This step should then be combined with the above
393: first step to compute the Chern character using connections,
394: curvature, and eventually computing moduli spaces of Yang-Mills
395: connections as was done for instance for the NC-torus in
396: \cite{CoRi}.
397:
398: \smallskip
399:
400: 3) The third step makes it possible to pass from the soft part of
401: differential
402: geometry to the harder ``Riemannian" metric aspect. The sought for
403: spectral geometry $(\cA,\cH,D)$ has three essential features:
404:
405: \begin{itemize}
406:
407: \item The $K$-homology class of $(\cA,\cH,D)$.
408:
409: \item The smooth structure.
410:
411: \item The metric.
412:
413: \end{itemize}
414:
415: One should always look for a spectral triple whose $K$-homology
416: class is as non-trivial as possible. Ideally it should extend to a
417: class for the double algebra $\cA\otimes \cA^o$ and then be a generator
418: for Poincar\'e duality. In general this is too much to ask for, since
419: many interesting spaces do not fulfill Poincar\'e duality. The main
420: tool for determining the stable homotopy class of the spectral
421: triple is Kasparov's bivariant $KK$-theory. Thus it is quite
422: important to already have taken step 2 and to look for classes whose
423: pairing with $K$-theory is as non-trivial as can be. For the smooth
424: structure, there is often a natural guess for a subalgebra
425: $A^\infty\subset A$ of the $C^*$-algebra $A=\bar\cA$ that will play the role
426: of the algebra of smooth functions. It should in general contain the
427: original algebra $\cA$ but should have the further property of being
428: stable under the holomorphic functional calculus. This ensures that
429: the inclusion $A^\infty\subset A$ is an isomorphism in $K$-theory
430: and makes it possible to complete the classification of smooth vector bundles.
431:
432: The role of the unbounded operator $D$ for the smooth structure is
433: that it defines the geodesic flow by the formula
434: $$
435: F_t(a)=\,e^{it|D|}\,a\,e^{-it|D|}\qqq a\in A^\infty
436: $$
437: and one expects that smoothness is governed by the smoothness of the
438: operator valued map $\R\ni t \mapsto F_t(a)$. The main result of the
439: general theory is the local index formula of \cite{CoMo}, which
440: provides the analogue of the Pontrjagin classes of smooth manifolds
441: in the noncommutative framework.
442:
443: The problem of determining $D$ from the knowledge of the
444: $K$-homology class is very similar to the choice of a connection on
445: a bundle. There are general results that assert the existence of an
446: unbounded selfadjoint $D$ with bounded commutators with $\cA$ from
447: estimates on the commutators with the phase $F$. The strongest is obtained
448: (\cf \cite{Co94}) just assuming that the $[F,a]$ are in an ideal
449: called ${\rm Li}\,\cH$ and it ensures the existence of a theta-summable
450: spectral triple which is what one needs to get started.
451:
452: It is not always possible to find a finitely-summable spectral
453: triple, first because of growth conditions on the algebra
454: \cite{Co-fr}, but also since the finitely-summable condition is very
455: analogous to type II in the theory of factors. In very general
456: cases, like the noncommutative space coming from foliations, one can
457: however go from type III to type II by passing to the total space of
458: the space of transverse metrics and then use the theory of
459: hypoelliptic operators \cite{CoMoHopf}.
460:
461:
462: \medskip
463:
464:
465:
466: Another way to attack the problem of determining $D$ is to consider
467: the larger algebra generated by $\cA$ and $D$, write a-priori
468: relations between $\cA$ and $D$ and then look for irreducible
469: representations that fall in the correct stable homotopy class.
470: Ideally one should minimize the spectral action functional
471: \cite{C-C2} in this homotopy class thus coming close to gravity. In
472: practise one should use anything available and the example of the
473: NC-space given by the quantum group $SU_q(2)$ shows that things can
474: be quite subtle \cite{SDLSV2}.
475:
476: Once the spectral triple $(\cA,\cH,D)$ has been determined, the basic
477: steps are the following, one should compute
478:
479: \begin{itemize}
480:
481: \item The dimension spectrum $\Sigma\subset \C$.
482:
483: \item The local index formula.
484:
485: \item The inner fluctuations, scalar curvature, and spectral action.
486:
487: \end{itemize}
488:
489:
490: \bigskip
491:
492: 4) Often a noncommutative space comes with a measure class, which in
493: turn determines a time evolution $\sigma_t$, namely a 1-parameter
494: family of automorphisms of the $C^*$-algebra $A=\bar\cA$. In the type II
495: situation one can apply the discussion of step 3 above and, in the
496: finite dimensional case, use the operator $D$ to represent
497: functionals in the measure class in the form
498: $$
499: \varphi(a)=\, \int\!\!\!\!\!\!-\ \,a \,|D|^{-p} \qqq \,a\in \cA
500: $$
501: where $\int\!\!\!\!\!\!-\ $ is the noncommutative integral \ie the
502: Dixmier trace and $p$ is the ``dimension".
503:
504: In the general case one should expect to be in the type III
505: situation in which the time evolution $\sigma_t$ is highly
506: non-trivial. We shall see some examples, for instance in Section
507: \ref{Qlatt}. Given the data $(\bar\cA,\sigma_t)$ it is natural to regard
508: it as a quantum statistical mechanical system, with $\bar\cA$ as algebra
509: of observables and $\sigma_t$ as time evolution. One can then look
510: for equilibrium states for the system, for a given value $\beta$ of
511: the thermodynamic parameter (inverse temperature).
512:
513: If the algebra $\bar\cA$ is concretely realized as an algebra of bounded
514: operators on a Hilbert space $\cH$, then one can consider the
515: Hamiltonian $H$, namely the (unbounded) operator on $\cH$ that is the
516: infinitesimal generator of the time evolution. If the operator
517: $\exp(-\beta H)$ is trace class, then one has equilibrium states for
518: the system $(\bar\cA,\sigma_t)$ written in the usual Gibbs form
519: $$
520: \varphi_\beta(a)=\frac{\Tr(a\exp(-\beta H))}{\Tr(\exp(-\beta H))},
521: $$
522: where $Z(\beta)=\Tr(\exp(-\beta H))$ is the partition function of the
523: system. The notion of equilibrium state continues to make sense
524: when $\exp(-\beta H)$ is not necessarily trace class, and is given by
525: the more subtle notion of KMS (Kubo--Martin--Schwinger) states.
526:
527: These are states on $\bar\cA$, namely continuous functionals
528: $\varphi: \bar\cA\to \C$ with $\varphi(1)=1$ and $\varphi(a^*a)\geq
529: 0$, satisfying the KMS$_\beta$ condition that, for all $a,b\in
530: \bar\cA$ there exists a function $F_{a,b}(z)$ which is holomorphic
531: on the strip $0< \Re(z) <i\beta$ continuous and bounded on the
532: closed strip and such that, for all $t\in\R$,
533: \begin{equation}\label{KMScond}
534: F_{a,b}(t)=\varphi(a\sigma_t(b)) \ \ \text { and } \ \
535: F_{a,b}(t+i\beta)=\varphi(\sigma_t(b)a).
536: \end{equation}
537: KMS states at zero temperature can be defined as weak limits as
538: $\beta\to \infty$ of KMS$_\beta$ states.
539:
540:
541: \begin{figure}
542: \begin{center}
543: \includegraphics[scale=0.9]{kmss.eps}
544: \end{center}
545: \caption{The KMS condition \label{FigKMS}}
546: \end{figure}
547:
548:
549: One can construct using KMS states very refined invariants of
550: noncommutative spaces. For a fixed $\beta$, the KMS$_\beta$ states
551: form a simplex, hence one can consider only the extremal KMS$_\beta$
552: states $\cE_\beta$, from which one recovers all the others by convex
553: combinations. An extremal KMS$_\beta$ state is always factorial and
554: the type of the factor is an invariant of the state. The simplest
555: situation is the type I. One can show under minimal hypothesis
556: (\cite{CCM}) that extremal KMS$_\beta$ states continue to survive
557: when one lowers the temperature \ie one increases $\beta$. Thus, in
558: essence, when cooling down the system this tends to become more and
559: more ``classical" and in the $0$-temperature limit $\cE_\beta$ gives
560: a good replacement of the notion of classical points for a
561: noncommutative space. We shall see in Section \ref{shimura} how, in
562: examples related to arithmetic, the ``classical points'' described
563: by the zero temperature KMS states of certain quantum statistical
564: mechanical systems recover classical arithmetic varieties. The
565: extremal KMS states at zero temperature, evaluated on suitable
566: arithmetic elements in the noncommutative algebra, can be shown in
567: significant cases to have an interesting Galois action, related to
568: interesting questions in number theory (\cf \cite{BC}, \cite{CMR},
569: \cite{CoMa}).
570:
571: In joint work with Consani, we showed in \cite{CCM}
572: how to define an analog in characteristic
573: zero of the action of the Frobenius on the etale cohomology by a
574: process involving the above thermodynamics. One key feature is that
575: the analogue of the Frobenius is the ``dual" of the above time
576: evolution $\sigma_t$. The process involves cyclic homology and its
577: three basic steps are (\cite{CCM})
578:
579: \begin{itemize}
580:
581:
582: \item Cooling.
583:
584: \item Distillation.
585:
586: \item Dual action of $\R_+^*$ on the cyclic homology of the
587: distilled space.
588:
589: \end{itemize}
590:
591:
592: \medskip
593: When applied to the simplest system
594: (the Bost--Connes system of \cite{BC})
595: this yields a cohomological interpretation of the spectral
596: realization of the zeros of the Riemann zeta function
597: (\cite{CoRH}, \cite{CCM}).
598:
599:
600:
601:
602: \bigskip
603:
604:
605: \section{Phase spaces of microscopic systems}\label{heisenberg}
606:
607: What can be regarded historically as the first example of a
608: non--commutative space is the Heisenberg formulation of the
609: observational Ritz-Rydberg law of spectrocopy. In fact, quantum
610: mechanics showed that indeed the parameter space, or phase space
611: of the mechanical system given by a single atom fails to be a
612: manifold. It is important to convince oneself of this fact and to
613: understand that this conclusion is indeed dictated by the {\it
614: experimental findings of spectroscopy}.
615:
616: At the beginning of the twentieth century a wealth of experimental
617: data was being collected on the spectra of various chemical
618: elements. These spectra obey experimentally discovered laws, the
619: most notable being the Ritz-Rydberg combination principle. The
620: principle can be stated as follows; spectral lines are indexed by
621: pairs of labels. The statement of the principle then is that
622: certain pairs of spectral lines, when expressed in terms of
623: frequencies, do add up to give another line in the spectrum.
624: Moreover, this happens precisely when the labels are of the form
625: $i,j$ and $j,k$.
626:
627: \begin{center}
628: \begin{figure}
629: \includegraphics[scale=0.9]{Ritz2.eps}
630: \caption{Spectral lines and the Ritz-Rydberg law.
631: \label{FigRitz}}
632: \end{figure}
633: \end{center}
634:
635: In the seminal paper \cite{Heisenberg} of 1925, Heisenberg
636: considers the classical prediction for the radiation emitted by a
637: moving electron in a field, where the observable dipole moment can
638: be computed, with the motion of the electron given in Fourier
639: expansion. The classical model would predict (in his notation)
640: frequencies distributed according to the law
641: \begin{equation}\label{emissionCM}
642: \nu(n,\alpha) = \alpha \nu(n) = \alpha \frac{1}{h} \frac{dW}{dn}.
643: \end{equation}
644: When comparing the frequencies obtained in this classical model
645: with the data, Heisenberg noticed that the classical law
646: \eqref{emissionCM} did not match the phenomenon observed.
647:
648: The spectral rays provide a `picture' of an atom: if atoms were
649: classical systems, then the picture formed by the spectral lines
650: would be (in our modern mathematical language) a {\it group},
651: which is what \eqref{emissionCM} predicts. That is, the classical
652: model predicts that the observed frequencies should simply add,
653: obeying a group law, or, in Heisenberg's notation, that
654: \begin{equation}\label{emission2}
655: \nu(n,\alpha)+ \nu(n,\beta)= \nu(n,\alpha+\beta).
656: \end{equation}
657: Correspondingly, observables would form the convolution algebra of
658: a group.
659:
660: What the spectral lines were instead providing was the picture of
661: a {\it groupoid}. Heisenberg realized that the classical law of
662: (\ref{emissionCM}) \eqref{emission2} would have to be replaced by
663: the quantum--mechanical
664: \begin{equation}\label{emissionQM}
665: \nu(n,n-\alpha) = \frac{1}{h} \left( W(n)- W(n-\alpha) \right).
666: \end{equation}
667: This replaces the group law with that of a groupoid, replacing the
668: classical \eqref{emission2} by the quantum--mechanical
669: \begin{equation}\label{emissionQ}
670: \nu(n,n-\alpha)+ \nu(n-\alpha, n-\alpha-\beta)=
671: \nu(n,n-\alpha-\beta).
672: \end{equation}
673: Similarly, the classical Fourier modes $\fU_{\alpha}(n)
674: e^{i\omega(n)\alpha t}$ were replaced by $ \fU
675: (n,n-\alpha)e^{i\omega(n,n-\alpha)t}$.
676:
677: The analysis of the emission spectrum given by Heisenberg was in
678: very good agreement with the Ritz--Rydberg law, or combination
679: principle, for spectral lines in emission or absorption spectra.
680:
681: In the same paper, Heisenberg also extends his redefinition of the
682: multiplication law for the Fourier coefficients to coordinates and
683: momenta, by introducing transition amplitudes that satisfy similar
684: product rules. This is, in Born's words, ``his most audacious
685: step'': in fact, it is precisely this step that brings
686: non--commutative geometry on the scene.
687:
688: It was Born who realized that what Heisenberg described in his
689: paper corresponded to replacing classical coordinates with
690: coordinates which no longer commute, but which obey the laws of
691: matrix multiplication. In his own words reported in
692: \cite{vdWaerden},
693:
694: \begin{quote}
695: After having sent Heisenberg's paper to the Zeitschrift f\"ur
696: Physik for publication, I began to ponder about his symbolic
697: multiplication, and was soon so involved in it that I thought the
698: whole day and could hardly sleep at night. For I felt there was
699: something fundamental behind it ... And one morning ... I suddenly
700: saw light: Heisenberg's symbolic multiplication was nothing but
701: the matrix calculus.
702: \end{quote}
703:
704: \bigskip
705:
706: Thus, spectral lines are parameterized by two indices
707: $L_{\alpha\beta}$ satisfying a cocycle relation
708: $L_{\alpha\beta}+L_{\beta\gamma}= L_{\alpha\gamma}$, and a
709: coboundary relation expresses each spectral line as a difference
710: $L_{\alpha\beta}=\nu_\alpha -\nu_\beta$. In other words, the
711: Ritz--Rydberg law gives the groupoid law \eqref{emissionQ}, or
712: equivalently,
713: $$ (i,j)\bullet (j,k)= (i,k) $$
714: and the convolution algebra of the group is replaced by
715: observables satisfying the matrix product
716: $$ (AB)_{ik} = \sum_j A_{ij} B_{jk}, $$
717: for which in general commutativity is lost:
718: $$ AB\neq BA. $$
719: The Hamiltonian $H$ is a matrix with the frequencies on the
720: diagonal, and observables obey the evolution equation
721: $$ \frac{d}{dt} A = i [H,A]. $$
722:
723: Out of Heisenberg's paper and Born's interpretation of the same in
724: terms of matrix calculus, emerged the statement of Heisenberg's
725: uncertainty principle in the form of a commutation relation of
726: matrices
727: $$ [P,Q] = \frac{h}{2\pi i} I. $$
728: The matrix calculus and the uncertainty principle were formulated
729: in the subsequent paper of Born and Jordan \cite{Born}, also
730: published in 1925. This viewpoint on quantum mechanics was later
731: somewhat obscured by the advent of the Schr\"odinger equation.
732: The Schr\"odinger approach shifted the emphasis back to the more
733: traditional technique
734: of solving partial differential equations, while the more modern
735: viewpoint of Heisenberg implied a much more serious change
736: of paradigm, affecting our most basic understanding of the notion of
737: space. Heisenberg's approach can be regarded as the historic origin of
738: noncommutative geometry.
739:
740: \bigskip
741:
742:
743: \section{Noncommutative quotients}\label{quotients}
744:
745: A large source of examples of noncommutative spaces is given by
746: quotients of equivalence relations. One starts by an ordinary
747: commutative space $X$ (\eg a smooth manifold or more generally a
748: locally compact Hausdorff topological space). This can be described
749: via its algebra of functions $C(X)$, and abelian $C^*$-algebra.
750: Suppose then that we are interested in taking a quotient $Y=X/\sim$
751: of $X$ with respect to an equivalence relation. In general, one
752: should not expect the quotient to be nice. Even when $X$ is a smooth
753: manifold, the quotient $Y$ need not even be a Hausdorff space. In
754: general, one would like to still be able to characterize $Y$ through
755: its ring of functions. One usually defines $C(Y)$ to be functions
756: on $X$ that are invariant under the equivalence relation ,
757: \begin{equation}\label{commCY}
758: C(Y)= \{ f\in C(X): f(a)=f(b), \forall a\sim b \} .
759: \end{equation}
760: Clearly, for a ``bad'' equivalence relation one typically
761: gets this way only constant functions $C(Y)=\C$.
762:
763: There is a better way to
764: associate to the quotient space $Y$ a ring of functions which is
765: nontrivial for any equivalence relation. This requires dropping the
766: commutativity requirement. One can then consider functions of two
767: variables $f_{ab}$ defined on the graph of the equivalence relation,
768: with a product which is no longer the commutative pointwise product,
769: but the noncommutative convolution product dictated by the groupoid of
770: the equivalence relation.
771: In general the elements in the algebra of functions
772: \begin{equation}\label{ncCY}
773: ``C(Y)\text{''}= \{ (f_{ab}) : a\sim b \}
774: \end{equation}
775: act as bounded operators on the Hilbert space $L^2$ of the
776: equivalence class. This also guarantees the convergence in the
777: operator norm of the convolution product
778: $$ \sum_{a\sim b\sim c} f_{ab} g_{bc}. $$
779:
780: We give a few examples to illustrate the difference between the
781: traditional construction and the one of noncommutative geometry.
782:
783: \begin{ex} {\em
784: Consider the space $Y=\{ x_0, x_1\}$ with the equivalence relation
785: $x_0\sim x_1$. With the first point of view the algebra of
786: functions on the quotient is $\C$, and in the second point of view
787: it is $\cB =M_2(\C)$, that is
788: \begin{equation}
789: \cB = \left\{ f = \left( \begin{array}{cc} f_{aa} & f_{ab} \\
790: f_{ba} & f_{bb} \end{array} \right) \right\} \, . \label{twopoint}
791: \end{equation}
792: These two algebras are not the same, though in this case they are
793: Morita equivalent. }\end{ex}
794:
795: Notice that, when one computes the spectrum of the algebra
796: (\ref{twopoint}), it turns out that it is composed of only one
797: point, so the two points $a$ and $b$ have been identified.
798: This first trivial example represents the typical situation where
799: the quotient space is ``nice'': the two constructions give Morita
800: equivalent algebras. In this sense, Morita
801: equivalent algebras are regarded as ``the same'' (or better
802: isomorphic) spaces in non--commutative geometry.
803:
804: \begin{ex} {\em
805: Consider the space $Y=[0,1] \times \{ 0,1\}$ with the equivalence
806: relation $(x,0)\sim (x,1)$ for $x\in (0,1)$. Then in the first
807: viewpoint the algebra of functions is again given just by the
808: constant functions $\C$, but in the second case we obtain
809: \begin{equation} \{ f\in C([0,1])\otimes M_2(\C) : \, f(0) \text{ and
810: } f(1) \text{ diagonal } \}. \label{alg1} \end{equation} }
811: \label{ex2} \end{ex}
812:
813: In this case, these algebras are not Morita equivalent. This can be
814: seen by computing their $K$--theory. This means that the approach of
815: non--commutative spaces produces something genuinely new, as soon as
816: the quotient space ceases to be ``nice'' in the classical sense.
817:
818: In general, the first kind of construction of functions on the
819: quotient space is cohomological in nature: one seeks for functions
820: satisfying certain equations or constraints. Usually there are
821: very few solutions. The second approach, instead, typically
822: produces a very large class of functions.
823:
824: \bigskip
825:
826: \section{Spaces of leaves of foliations}\label{foliations}
827:
828: There is a very rich collection of examples of noncommutative spaces
829: given by the leaf spaces of foliations. The connection thus
830: obtained between noncommutative geometry and the geometric theory of
831: foliations is very far reaching for instance through the role of
832: Gelfand-Fuchs cohomology, of the Godbillon-Vey invariant and of the
833: passage from type III to type II using the transverse frame bundle.
834: It is this class of examples that triggered the initial development
835: of cyclic cohomology (\cf \cite{khal} section 4), of the local index
836: formula in noncommutative geometry as well as the theory of
837: characteristic classes for Hopf algebra actions.
838:
839:
840: \medskip
841:
842: The construction of the algebra associated to a foliation is a
843: special case of the construction of section \ref{quotients} but both
844: the presence of holonomy and the case when the graph of the
845: foliation is non-hausdorff require special care, so we shall recall
846: the basic steps below.
847:
848: Let $V$ be a smooth manifold and $TV$ its tangent bundle, so that
849: for each $x \in V$, $T_x V$ is the tangent space of $V$ at $x$. A
850: smooth subbundle $F$ of $TV$ is called {\it integrable} iff one of
851: the following equivalent conditions is satisfied:
852:
853: \smallskip
854:
855: \begin{enumerate}
856:
857: \item[a)] Every $x \in V$ is contained in a submanifold $W$ of $V$ such that
858: $$
859: T_y (W) = F_y \qquad \forall \, y \in W \, .
860: $$
861:
862: \smallskip
863:
864: \item[b)] Every $x \in V$ is in the domain $U \subset V$ of a
865: submersion $p : U \to {\mathbb R}^q$ ($q = {\rm codim} \, F$) with
866: $$
867: F_y = {\rm Ker} (p_*)_y \qquad \forall \, y \in U \, .
868: $$
869:
870: \smallskip
871: \item[c)] $C^{\infty} (F) = \{ X \in C^{\infty} (TV) \, , \ X_x \in
872: F_x \quad \forall \, x \in V \}$ is a Lie algebra.
873:
874: \smallskip
875:
876: \item[d)] The ideal $J(F)$ of smooth exterior differential forms which
877: vanish on $F$ is stable by exterior differentiation.
878: \end{enumerate}
879:
880: \medskip
881:
882: Any $1$-dimensional subbundle $F$ of $TV$ is integrable, but for
883: $\dim F \geqq 2$ the condition is non trivial, for instance if $P
884: \overset{p}{\to} B$ is a principal $H$-bundle (with compact
885: structure group $H$) the bundle of horizontal vectors for a given
886: connection is integrable iff this connection is flat.
887:
888: \smallskip
889:
890: A foliation of $V$ is given by an integrable subbundle $F$ of $TV$.
891: The leaves of the foliation $(V,F)$ are the maximal connected
892: submanifolds $L$ of $V$ with $T_x (L) = F_x $, $\forall \, x \in L$,
893: and the partition of $V$ in leaves $$V = \cup
894: L_{\alpha}\,,\quad\alpha \in X$$ is characterized geometrically by
895: its ``local triviality'': every point $x \in V$ has a neighborhood
896: $U$ and a system of local coordinates
897: $(x^j)_{j = 1 , \ldots , \dim V}$ called
898: {\it foliation charts}, so
899: that the partition of $U$ in connected components of
900: leaves corresponds to the partition of $${\mathbb
901: R}^{\dim V} = {\mathbb R}^{\dim F} \times {\mathbb R}^{\text{codim}
902: \, F}$$ in the parallel affine subspaces ${\mathbb R}^{\dim F}
903: \times {\rm pt}$. These are the leaves of the
904: restriction of $F$ and are called {\it plaques}.
905:
906: The set $X=V/F$ of leaves of a foliation $(V,F)$
907: is in most cases a noncommutative space. In other words even though
908: as a set it has the cardinality of the continuum it is in general
909: not so at the effective level and it is in general impossible to
910: construct a countable set of measurable functions on $V$ that form a
911: complete set of invariants for the equivalence relation coming from
912: the partition of $V$ in leaves $V = \cup L_{\alpha}$. Even in the
913: simple cases in which the set $X=V/F$ of leaves is classical it
914: helps to introduce the associated algebraic tools in order to get a
915: feeling for their role in the general singular case.
916:
917: \smallskip
918:
919: To each foliation $(V,F)$ is canonically associated a $C^*$ algebra
920: $C^* (V,F)$ which encodes the topology of the space of leaves. The
921: construction is basically the same as the general one for quotient
922: spaces of Section \ref{quotients}, but there are interesting nuances
923: coming from the presence of holonomy in the foliation context. To
924: take this into account one first constructs a manifold $G$, $\dim
925: \,G = \dim \,V + \dim \,F$, called the graph (or holonomy groupoid)
926: of the foliation, which refines the equivalence relation coming from
927: the partition of $V$ in leaves $V = \cup L_{\alpha}$. This
928: construction is due to Thom, Pradines and Winkelnkemper,
929: see \cite{Winkel}.
930:
931: An element $\gamma$ of $G$ is given by two points $x = s(\gamma)$,
932: $y = r(\gamma)$ of $V$ together with an equivalence class of smooth
933: paths: $\gamma (t)\in V$, $t \in [0,1]$; $\gamma (0) = x$, $\gamma
934: (1) = y$, tangent to the bundle $F$ ({\it i.e.} with $\dot\gamma (t)
935: \in F_{\gamma (t)}$, $\forall \, t \in {\mathbb R}$) up to the
936: following equivalence: $\gamma_1$ and $\gamma_2$ are equivalent iff
937: the {\it holonomy} of the path $\gamma_2 \circ \gamma_1^{-1}$ at the
938: point $x$ is the {\it identity}. The graph $G$ has an obvious
939: composition law. For $\gamma , \gamma' \in G$, the composition
940: $\gamma \circ \gamma'$ makes sense if $s(\gamma) = r(\gamma')$. If
941: the leaf $L$ which contains both $x$ and $y$ has no holonomy, then
942: the class in $G$ of the path $\gamma (t)$ only depends on the pair
943: $(y,x)$. The condition of trivial holonomy
944: is generic in the topological sense of dense $G_{\delta}$'s.
945: In general, if one fixes $x = s(\gamma)$, the map from $G_x
946: = \{ \gamma , s(\gamma) = x \}$ to the leaf $L$ through $x$, given
947: by $\gamma \in G_x \mapsto y = r(\gamma)$, is the holonomy covering
948: of $L$.
949:
950: Both maps $r$ and $s$ from the manifold $G$ to $V$ are smooth
951: submersions and the map $(r,s)$ to $V \times V$ is an immersion
952: whose image in $V \times V$ is the (often singular) subset
953: \begin{equation}\label{subset}
954: \{ (y,x)\in V \times V: \, \text{ $y$ and $x$ are on the same leaf}\}.
955: \end{equation}
956: In first
957: approximation one can think of elements of $C^* (V,F)$ as continuous
958: matrices $k(x,y)$, where $(x,y)$ varies in the set \eqref{subset}.
959: We now describe this $C^*$ algebra in all details. We
960: assume, for notational convenience, that the manifold $G$ is
961: Hausdorff. Since this fails to be the case in very interesting
962: examples, we also explain briefly how to remove this
963: hypothesis.
964:
965: \smallskip
966:
967: The basic elements of $C^* (V,F)$ are smooth half densities
968: $f \in C_c^{\infty} (G , \Omega^{1/2})$ with
969: compact support on $G$. The
970: bundle $\Omega_G^{1/2}$ of half densities on $G$ is obtained as
971: follows. One first defines a line bundle $\Omega_V^{1/2}$ on V. For
972: $x\in V$ one lets $\Omega_x^{1/2}$ be the one dimensional complex
973: vector space of maps from the exterior power $\wedge^k \, F_x$, $k =
974: \dim F$, to ${\mathbb C}$ such that
975: $$
976: \rho \, (\lambda \, v) = \vert \lambda \vert^{1/2} \, \rho \, (v)
977: \qquad \forall \, v \in \wedge^k \, F_x \, , \quad \forall \,
978: \lambda \in {\mathbb R} \, .
979: $$
980: Then, for $\gamma \in G$, one can identify $\Omega_{\gamma}^{1/2}$ with the one
981: dimensional complex vector space $\Omega_y^{1/2} \otimes
982: \Omega_x^{1/2}$, where $\gamma : x \to y$. In other words
983: $$
984: \Omega_G^{1/2}=\, r^*(\Omega_V^{1/2})\otimes s^*(\Omega_V^{1/2})\,.
985: $$
986: Of course the bundle $\Omega_V^{1/2}$ is trivial on $V$, and we
987: could choose once and for all a trivialisation $\nu$ turning
988: elements of $C_c^{\infty} (G , \Omega^{1/2})$ into functions.
989: Let us
990: however stress that the use of half densities makes all the
991: construction completely canonical.
992:
993: For $f,g \in C_c^{\infty} (G , \Omega^{1/2})$, the convolution
994: product $f * g$ is defined by the equality
995: $$
996: (f * g) (\gamma) = \int_{\gamma_1 \circ \gamma_2 = \gamma}
997: f(\gamma_1) \, g(\gamma_2) \, .
998: $$
999: This makes sense because, for fixed $\gamma : x \to y$ and fixing $v_x
1000: \in \wedge^k \, F_x$ and $v_y \in \wedge^k \, F_y$, the product
1001: $f(\gamma_1) \, g(\gamma_1^{-1} \gamma)$ defines a $1$-density on
1002: $G^y = \{ \gamma_1 \in G , \, r (\gamma_1) = y \}$, which is smooth
1003: with compact support (it vanishes if $\gamma_1 \notin$ support $f$),
1004: and hence can be integrated over $G^y$ to give a scalar, namely $(f * g)
1005: (\gamma)$ evaluated on $v_x , v_y$.
1006:
1007: The $*$ operation is defined by $f^* (\gamma) =
1008: \overline{f(\gamma^{-1})}$, {\it i.e.} if $\gamma : x \to y$ and
1009: $v_x \in \wedge^k \, F_x$, $v_y \in \wedge^k \, F_y$ then $f^*
1010: (\gamma)$ evaluated on $v_x , v_y$ is equal to
1011: $\overline{f(\gamma^{-1})}$ evaluated on $v_y , v_x$. We thus get a
1012: $*$ algebra $C_c^{\infty} (G , \Omega^{1/2})$. For each leaf $L$ of
1013: $(V,F)$ one has a natural representation of this $*$ algebra on the
1014: $L^2$ space of the holonomy covering $\tilde L$ of $L$. Fixing a
1015: base point $x \in L$, one identifies $\tilde L$ with $G_x = \{
1016: \gamma , s(\gamma) = x \}$ and defines
1017: $$
1018: (\pi_x (f) \, \xi) \, (\gamma) = \int_{\gamma_1 \circ \gamma_2 =
1019: \gamma} f(\gamma_1) \, \xi (\gamma_2) \qquad \forall \, \xi \in L^2
1020: (G_x),
1021: $$
1022: where $\xi$ is a square integrable half density on $G_x$. Given
1023: $\gamma : x \to y$ one has a natural isometry of $L^2 (G_x)$ on $L^2
1024: (G_y)$ which transforms the representation $\pi_x$ in $\pi_y$.
1025:
1026:
1027: By definition $C^* (V,F)$ is the $C^*$ algebra completion of
1028: $C_c^{\infty} (G , \Omega^{1/2})$ with the norm $$\Vert f \Vert =
1029: \sup_{x \in V} \, \Vert \pi_x (f) \Vert\,.$$
1030:
1031:
1032: Note that $C^* (V,F)$ is always norm separable and admits a natural
1033: smooth subalgebra, namely the algebra $C_c^\infty (V,F)=C_c^{\infty} (G ,
1034: \Omega^{1/2})$ of smooth compactly supported half densities.
1035:
1036: If the leaf $L$ has trivial holonomy then the representation
1037: $\pi_x$, $x \in L$, is irreducible. In general, its commutant is
1038: generated by the action of the (discrete) holonomy group $G_x^x$ in
1039: $L^2 (G_x)$. If the foliation comes from a submersion $p : V \to B$,
1040: then its graph $G$ is $\{ (x,y) \in V \times V , \, p(x) = p(y)\}$
1041: which is a submanifold of $V \times V$, and $C^* (V,F)$ is identical
1042: with the algebra of the continuous field of Hilbert spaces $L^2
1043: (p^{-1} \{ x \})_{x \in B}$. Thus (unless $\dim \,F = 0$) it is
1044: isomorphic to the tensor product of $C_0 (B)$ with the elementary
1045: $C^*$ algebra of compact operators. If the foliation comes from an
1046: action of a Lie group $H$ in such a way that the graph is identical
1047: with $V \times H$ (this is not always true even for flows) then
1048: $C^* (V,F)$ is identical with the reduced crossed product of $C_0
1049: (V)$ by $H$. Moreover the construction of $C^* (V,F)$ is local in
1050: the following sense.
1051:
1052:
1053: If $V' \subset V$ is an open set and $F'$ is the restriction of $F$
1054: to $V'$, then the graph $G'$ of $(V' , F')$ is an open set in the
1055: graph $G$ of $(V,F)$, and the inclusion $C_c^{\infty} (G' ,
1056: \Omega^{1/2}) \subset C_c^{\infty} (G , \Omega^{1/2})$ extends to an
1057: isometric $*$ homomorphism of $C^* (V' , F')$ in $C^* (V,F)$. The
1058: proof is straightforward and also applies in the case of non
1059: Hausdorff graph.
1060:
1061:
1062: Let us now briefly explain how the construction of the $C^*$ algebra
1063: $C^* (V,F)$ has to be done in the case when the graph of the
1064: foliation is not Hausdorff. This case is rather rare, since it never
1065: occurs if the foliation is real analytic. However, it does occur in
1066: cases which are topologically interesting for foliations, such as
1067: the Reeb foliation of the 3 sphere, which are constructed by
1068: patching together foliations of manifolds with boundaries $(V_i ,
1069: F_i)$ where the boundary $\partial V_i$ is a leaf of $F_i$. In fact
1070: most of the constructions done in geometry to produce smooth
1071: foliations of given codimension on a given manifold give a non
1072: Hausdorff graph. The $C^*$-algebra $C^* (V,F)$ turns out in this
1073: case to be obtained as a fibered product of the $C^* (V_i , F_i)$.
1074:
1075: \begin{figure}
1076: \begin{center}
1077: \includegraphics[scale=0.6]{reeb.eps}
1078: \end{center}
1079: \caption{The Reeb foliation \label{FigReeb}}
1080: \end{figure}
1081:
1082:
1083: In the general non-Hausdorff case the graph $G$ of $(V,F)$, being
1084: non Hausdorff may have only very few continuous fonctions with
1085: compact support. However, being a manifold, we can give a local chart
1086: $U \overset{\chi}{\to} {\mathbb R}^{\dim G}$. Take a smooth function
1087: $\varphi \in C_c^{\infty} ({\mathbb R}^{\dim G})$, ${\rm Supp} \,
1088: \varphi \subset \chi (U)$ and consider the function on $G$ equal to
1089: $\varphi \circ \chi$ on $U$ and to $0$ outside $U$. If $G$ were
1090: Hausdorff, this would generate all of $C_c^{\infty} (G)$ by taking
1091: linear combinations, and in general we take this linear span as the
1092: {\it definition} of $C_c^{\infty} (G)$. Note that we do not get {\it
1093: continuous} functions, since there may well be a sequence $\gamma_n
1094: \in U$ with two limits, one in ${\rm Supp} \, \varphi \circ \chi$
1095: one in the complement of $U$. The above definition of $C_c^{\infty}
1096: (G)$ obviously extends to get $C_c^{\infty} (G,\Omega^{1/2})$ the
1097: space of smooth $\frac{1}{2}$ densities on $G$. One then shows that
1098: the convolution $\varphi_1 * \varphi_2$ of $\varphi_1 , \varphi_2
1099: \in C_c^{\infty} (G , \Omega^{1/2})$ is in $C_c^{\infty} (G ,
1100: \Omega^{1/2})$.
1101:
1102:
1103: Then we proceed exactly as in the Hausdorff case, and construct the
1104: representation $\pi_x$ of the $*$ algebra $C_c^{\infty} (G ,
1105: \Omega^{1/2})$ in the Hilbert space $L^2 (G_x)$. We note that though
1106: $G$ is not Hausdorff, each $G_x$ is {\it Hausdorff}, being the
1107: holonomy covering of the leaf through $x$.
1108:
1109:
1110: For each $\varphi \in C_c^{\infty} (G , \Omega^{1/2})$ and $x \in
1111: V$, $\pi_x (\varphi)$ is an ordinary smoothing operator, bounded in
1112: $L^2 (G_x)$.
1113:
1114: Exactly as in the Hausdorff case $C^* (V,F)$ is defined as the
1115: $C^*$ completion of $C_c^{\infty} (G , \Omega^{1/2})$ with norm
1116: $\sup_{x \in V} \, \Vert \pi_x (\varphi) \Vert$.
1117:
1118: \medskip
1119:
1120: The obtained functor from foliations to $C^*$-algebras makes it
1121: possible first
1122: of all to translate from basic geometric properties to algebraic
1123: ones and the simplest examples of foliations already exhibit
1124: remarkable $C^*$-algebras. For instance the horocycle foliation of
1125: the unit sphere bundle of a Riemann surface of genus $>1$ gives a
1126: simple $C^*$-algebra without idempotent. The Kronecker foliation
1127: gives rise to the noncommutative torus, which we describe
1128: in more detail in Section \ref{nctori}.
1129:
1130: In the type II situation \ie in the presence of a holonomy invariant
1131: transverse measure $\Lambda$ the basic result of the theory is the
1132: longitudinal index theorem which computes the $L^2$-index of
1133: differential operators $D$ on the foliated manifold $(V,F)$ which
1134: are elliptic in the longitudinal direction (\ie $D$ restrict to
1135: the leaves $L$ as elliptic operators $D_L$). One starts with a pair of
1136: smooth vector bundles $E_1$, $E_2$ on $V$ together with a
1137: differential operator $D$ on $V$ from sections of $E_1$ to sections
1138: of $E_2$ such that:
1139: \begin{enumerate}
1140: \item[1)] $D$ restricts to leaves, \ie $(D \xi)_x$
1141: only depends on the restriction of $\xi$ to a neighborhood of
1142: $x$ in the leaf of $x$ ({\it i.e.} $D$ only uses partial
1143: differentiation in the leaf direction).
1144: \item[2)] $D$ is elliptic when restricted to any leaf.
1145: \end{enumerate}
1146:
1147: \begin{thm}\label{longmeasindex} \cite{Co78}
1148:
1149: {\rm a)} There exists a Borel transversal $B$ (resp. $B'$) such that
1150: the bundle $(\ell^2 (L \cap B))_{L \in V/F}$ is measurably
1151: isomorphic to the bundle $({\rm Ker} \, D_L)_{L \in V/F}$ (resp. to
1152: $({\rm Ker} \, D_L^*)_{L \in V/F})$.
1153:
1154: \smallskip
1155:
1156: \noindent {\rm b)} The scalar $\Lambda (B) < \infty$ is independent
1157: of the choice of $B$ and noted $\dim_{\Lambda} ({\rm Ker} \, (D))$.
1158:
1159: \smallskip
1160:
1161: \noindent ${\rm c)}\qquad\dim_{\Lambda} ({\rm Ker} \, (D)) -
1162: \dim_{\Lambda} ({\rm Ker} \, (D^*)) = \,\varepsilon \, \langle {\rm
1163: ch} \, \sigma_D \, {\rm Td} \, (F_{\mathbb C}) , [C] \rangle$
1164:
1165:
1166: ($\varepsilon = (-1) \, \frac{k(k+1)}{2}$, $k = \dim F$,
1167: ${\rm Td} \, (F_{\mathbb C})=$ Todd genus,
1168: $\sigma_D=$ symbol of $D$) .
1169:
1170:
1171: \end{thm}
1172:
1173: Here $[C]\in H_k(V,\C)$ is the homology class of the Ruelle-Sullivan
1174: current, a closed de-Rham current of dimension $k = \dim F$ which
1175: encodes the transverse measure $\Lambda$ by integration of a
1176: $k$-dimensional differential form $\omega$ on $V$ along the plaques
1177: of foliation charts.
1178:
1179: \bigskip
1180:
1181:
1182: In particular the Betti numbers $\beta_j$ of a measured foliation
1183: were defined in \cite{Co78} and give the $L^2$-dimension of the
1184: space of $L^2$-harmonic forms along the leaves, more precisely one
1185: has the following result.
1186:
1187: {\rm a)} For each $j = 0,1,2,\ldots ,\dim F$, there exists a Borel
1188: transversal $B_j$ such that the bundle $(H^j (L,{\mathbb C}))_{L \in
1189: V/F}$ of $j$-th square integrable harmonic forms on $L$ is
1190: measurably isomorphic to $(\ell^2 (L \cap B))_{L \in V/F}$.
1191:
1192: \smallskip
1193:
1194: \noindent {\rm b)} The scalar $\beta_j = \Lambda (B_j)$ is finite,
1195: independent of the choice of $B_j$, of the choice of the Euclidean
1196: structure on $F$.
1197:
1198: \smallskip
1199:
1200: \noindent {\rm c)} One has $\Sigma \, (-1)^j \, \beta_j = \chi
1201: (F,\Lambda)$.
1202:
1203: Here the Euler characteristic is simply given by the pairing of the
1204: Ruelle-Sullivan current with the Euler class $e(F)$ of the oriented
1205: bundle $F$ on $V$.
1206:
1207: Extending ideas of Cheeger and Gromov \cite{ChGr} in the case of discrete groups,
1208: D.~Gaboriau has shown in a remarkable recent work
1209: (\cf \cite{Gabo1}, \cite{Gabo2}) that the Betti
1210: numbers $\beta_j(F,\Lambda)$ of a foliation with contractible leaves
1211: are invariants of the measured equivalence relation $\mathcal
1212: R=\,\{(x,y)\,|\,y \in {\rm leaf}(x)\}$.
1213:
1214: \medskip
1215:
1216: In the general case one cannot expect to have a holonomy invariant
1217: transverse measure and in fact the simplest foliations are of type
1218: III from the measure theoretic point of view. Obtaining an analogue
1219: in general of Theorem \ref{longmeasindex} was the basic motivation
1220: for the construction of the assembly map (the second step of section
1221: \ref{road}). Let us now briefly state the longitudinal index
1222: theorem.
1223:
1224:
1225: Let $D$ be as above an elliptic differential operator along the
1226: leaves of the foliation $(V,F)$. Since $D$ is elliptic it has an
1227: inverse modulo $C^* (V,F)$ hence it gives an element ${\rm Ind}_a
1228: (D)$ of $K_0 (C^* (V,F))$ which is the {\em analytic index} of $D$.
1229: The {\em topological index} is obtained as follows. Let $i$ be an
1230: auxiliary imbedding of the manifold $V$ in ${\mathbb R}^{2n}$. Let
1231: $N$ be the total space of the normal bundle to the leaves: $N_x =
1232: (i_* (F_x))^{\perp} \subset {\mathbb R}^{2n}$. Let us foliate
1233: $\tilde V = V \times \R^{2n}$ by $\tilde F$, $\tilde
1234: F_{(x,t)} = F_x \times \{ 0 \}$, so that the leaves of $(\tilde V ,
1235: \tilde F)$ are just $\tilde L = L \times \{ t \}$, where $L$ is a
1236: leaf of $(V,F)$ and $t \in {\mathbb R}^{2n}$. The map $(x,\xi) \to
1237: (x,i(x) + \xi)$ turns an open neighborhood of the $0$-section in $N$
1238: into an open transversal $T$ of the foliation $(\tilde V , \tilde
1239: F)$. For a suitable open neighborhood $\Omega$ of $T$ in $\tilde V$,
1240: the $C^*$-algebra $C^* (\Omega , \tilde F)$ of the restriction of
1241: $\tilde F$ to $\Omega$ is (Morita) equivalent to $C_0 (T)$, hence
1242: the inclusion $C^* (\Omega , \tilde F) \subset C^* (\tilde V ,
1243: \tilde F)$ yields a $K$-theory map: $K^0 (N) \to K_0 (C^* (\tilde V
1244: , \tilde F))$. Since $C^* (\tilde V , \tilde F) = C^* (V,F) \otimes
1245: C_0 (\R^{2n})$, one has, by Bott periodicity, the equality
1246: $K_0 (C^* (\tilde V , \tilde F)) = K_0 (C^* (V,F))$.
1247:
1248: Using the Thom isomorphism, $K^0 (F^*)$ is identified with $K^0 (N)$
1249: so that one gets by the above construction the topological index:
1250: $$
1251: {\rm Ind}_t : K^0 (F^*) \to K_0 (C^* (V,F)) \, .
1252: $$
1253: The {\em longitudinal index theorem} \cite{CoSk} is the equality
1254: \begin{equation}\label{longind}
1255: {\rm Ind}_a (D) = {\rm Ind}_t ([\sigma_D]),
1256: \end{equation}
1257: where $\sigma_D$ is the
1258: longitudinal symbol of $D$ and $[\sigma_D]$ is its class in $K^0
1259: (F^*)$.
1260:
1261: Since the group $K_0 (C^* (V,F))$ is still fairly hard to compute
1262: one needs computable invariants of its elements and this is where
1263: cyclic cohomology enters the scene. In fact its early development
1264: was already fully completed in 1981 for that precise goal (\cf
1265: \cite{khal}). The role of the trace on $C^* (V,F)$ associated to the
1266: transverse measure $\Lambda$ is now played by cyclic cocycles on a
1267: dense subalgebra of $C^* (V,F)$. The hard analytic problem is to
1268: show that these cocycles have enough semi-continuity properties to
1269: define invariants of $K_0 (C^* (V,F))$. This was achieved for some
1270: of them in \cite{Co1} and makes it possible to formulate corollaries whose
1271: statements are independent of the general theory, such as
1272:
1273: \smallskip
1274: \begin{thm}\label{transahat} \cite{Co1}
1275: Let $M$ be a compact oriented manifold and assume that the $\hat
1276: A$-genus
1277: $\hat A (M)$ is non-zero (since $M$ is not
1278: assumed to be a spin manifold $\hat A (M)$ need not be an integer).
1279: Let then $F$ be an integrable Spin sub-bundle of $TM$. There exists
1280: no metric on $F$ for which the scalar curvature (of the leaves) is
1281: strictly positive $(\geq \varepsilon > 0)$ on $M$.
1282: \end{thm}
1283:
1284: There is a very rich interplay between the theory of foliations and
1285: their characteristic classes and operator algebras even at the
1286: purely measure theoretic level \ie the classification of factors.
1287:
1288: In a remarkable series of papers (see \cite{Hu} for references),
1289: J.~Heitsch and S.~Hurder have analyzed the interplay between the
1290: vanishing of the Godbillon-Vey invariant of a compact foliated
1291: manifold $(V,F)$ and the type of the von Neumann algebra of the
1292: foliation. Their work culminates in the following beautiful result
1293: of S.~Hurder (\cite{Hu}). If the von Neumann algebra is {\it
1294: semi-finite}, then the Godbillon-Vey invariant {\it vanishes}. We
1295: have shown, in fact, that cyclic cohomology yields a stronger
1296: result, proving that, if ${\rm GV} \ne 0$, then the central
1297: decomposition of $M$ contains necessarily factors $M$, whose virtual
1298: modular spectrum is of finite covolume in ${\mathbb R}_+^*$.
1299:
1300: \begin{thm}\label{th3} \cite{Co1}
1301: Let $(V,F)$ be an oriented, transversally oriented, compact,
1302: foliated manifold, $({\rm codim} \, F = 1)$. Let $M$ be the
1303: associated von Neumann algebra, and ${\rm Mod}(M)$ be its flow of
1304: weights.
1305: Then, if the Godbillon-Vey class of
1306: $(V,F)$ is different from $0$, there exists an invariant probability
1307: measure for the flow ${\rm Mod}(M)$.
1308: \end{thm}
1309:
1310: One actually constructs an invariant measure for the flow ${\rm
1311: Mod}(M)$, exploiting the following remarkable property of the
1312: natural cyclic 1-cocycle $\tau$ on the algebra $\cA$ of the
1313: transverse 1-jet bundle for the foliation. When viewed as a linear
1314: map $\delta$ from $\cA$ to its dual, $\delta$ is an unbounded
1315: derivation, which is {\em closable}, and whose domain extends to the
1316: center $Z$ of the von-Neumann algebra generated by $\cA$. Moreover,
1317: $\delta$ vanishes on this center, whose elements $h\in Z$ can then
1318: be used to obtain new cyclic cocycles $\tau_h$ on $\cA$. The pairing
1319: $$
1320: L(h)=\,\langle \tau_h,\, \mu(x)\rangle
1321: $$
1322: with the K-theory classes $\mu(x)$ obtained from the assembly map
1323: $\mu$, which we had constructed with P. Baum \cite{BauCo},
1324: then gives a measure on $Z$, whose invariance under the
1325: flow of weights follows from the discreteness of the K-group. To
1326: show that it is non-zero, one uses an index formula that evaluates
1327: the cyclic cocycles, associated as above to the Gelfand-Fuchs
1328: classes, on the range of the assembly map $\mu$.
1329:
1330: \medskip
1331:
1332: The central question in the analysis of the noncommutative leaf
1333: space of a foliation is step 3) (of section \ref{road}), namely the
1334: metric aspect which entails in particular constructing a spectral
1335: triple describing the transverse geometry. The reason why the
1336: problem is really difficult is that it essentially amounts to doing
1337: ``metric" geometry on manifolds in a way which is ``background
1338: independent" to use the terminology of physicists \ie which is {\em
1339: invariant} under diffeomorphisms rather than covariant as in
1340: traditional Riemannian geometry. Indeed the transverse space of a
1341: foliation is a manifold endowed with the action of a large pseudo
1342: group of partial diffeomorphisms implementing the holonomy. Thus in
1343: particular no invariant metric exists in the general case and the
1344: situation is very similar to trying to develop gravity without
1345: making use of any particular ``background" metric that automatically
1346: destroys the invariance under the action of diffeomeorphisms (\cf
1347: \cite{CoMo-back}). Using the theory of hypoelliptic differential
1348: operators and the basic technique of reduction from type III to type
1349: II, a general construction of a spectral triple was done by
1350: Connes-Moscovici in \cite{CoMo}. The remaining problem of the
1351: computation of the local index formula in cyclic cohomology was
1352: solved in \cite{CoMoHopf} and led in particular to the discovery of
1353: new symmetries given by an action of a Hopf algebra which only
1354: depends upon the transverse dimension of the foliation.
1355:
1356: This also led to the development of the noncommutative analogue of
1357: the Chern-Weil theory of characteristic classes \cite{CoMoHopf2} in
1358: the general context of Hopf algebra actions on noncommutative spaces
1359: and cyclic cohomology, a subject which is undergoing rapid progress,
1360: in particular thanks to the recent works of M. Khalkhali and
1361: collaborators \cite{kr1}, \cite{kr2}, \cite{kr3}, \cite{Hajac}.
1362:
1363: \bigskip
1364:
1365: \section{The noncommutative tori}\label{nctori}
1366:
1367: This is perhaps considered as the prototype example of a
1368: noncommutative space, since it illustrates very clearly
1369: the properties and structures of noncommutative geometries.
1370: Noncommutative tori played a key role in
1371: the early developments of the theory in the 1980's (\cf
1372: \cite{C3}), giving rise to noncommutative analogues of
1373: vector bundles, connections, curvature, etc.
1374:
1375: \smallskip
1376:
1377: One can regard noncommutative tori as a special case of
1378: noncommutative spaces arising from foliations. In this case, one
1379: considers certain vector fields on the ordinary 2-dimensional
1380: real torus $T^2=\R^2/\Z^2$. In fact, one considers on $T^2$ the
1381: Kronecker foliation $dx =\theta dy$, where $\theta$ is a given real
1382: number. We are especially interested in the case where $\theta$ is
1383: irrational. That is, we consider the
1384: space of solutions of the differential
1385: equation,
1386: \begin{equation}
1387: dx = \theta dy, \ \ \forall x,y \in \R/\Z \label{Krfol}
1388: \end{equation}
1389: for $\theta \in ]0,1[$ is a fixed irrational number. In other words, we
1390: are considering the space of
1391: leaves of the Kronecker foliation on the torus (\cf Figure \ref{FigNCtori}).
1392:
1393: \begin{figure}
1394: \begin{center}
1395: \includegraphics[scale=0.8]{NCtorus1.eps}
1396: \end{center}
1397: \caption{The Kronecker foliation and the noncommutative
1398: torus \label{FigNCtori}}
1399: \end{figure}
1400:
1401: We can choose a transversal $T$ to the foliation, given by
1402: $$ T=\{ y=0 \}, \ \ \ \ T\cong S^1\cong\R/\Z. $$
1403: Two points of the transversal which differ by an integer multiple
1404: of $\theta$ give rise to the same leaf.
1405: We want to describe the further quotient
1406: \begin{equation}
1407: S^1 / \theta \Z \label{quottheta}
1408: \end{equation}
1409: by the equivalence relation which identifies any two points on the
1410: orbits of the irrational rotation
1411: \begin{equation}
1412: R_{\theta} x = x + \theta \mod 1 \, .
1413: \label{Rtheta}
1414: \end{equation}
1415:
1416: \smallskip
1417:
1418: We can regard the circle $S^1$ and the quotient space
1419: \eqref{quottheta} at various levels of
1420: regularity (smooth, topological, measurable). This corresponds to
1421: different algebras of functions on the circle,
1422: \begin{equation}
1423: C^{\infty} (S^1) \subset C(S^1) \subset L^{\infty} (S^1) \, .
1424: \label{S1}
1425: \end{equation}
1426:
1427: When passing to the quotient \eqref{quottheta}, if we just consider
1428: invariant functions we obtain a very poor algebra of functions, since,
1429: even at the measurable level, we would only have constant functions.
1430: If instead we consider the noncommutative algebra of functions
1431: obtained by the general recipe of ``noncommutative quotients''
1432: (functions on the graph of the equivalence relation with the
1433: convolution product), we obtain a very interesting and highly
1434: non--trivial algebra of functions describing the space of leaves
1435: of the foliation. This is given (in the topological category) by the
1436: ``irrational rotation algebra'', \ie the $C^*$-algebra
1437: \begin{equation} \label{algNCT}
1438: \cA_\theta: = \{ \left( a_{ij} \right)
1439: \, \, \, \, i,j\in T \, \, \text{ in the same leaf } \}.
1440: \end{equation}
1441: Namely, elements in the algebra $\cA_\theta$ associated to the
1442: transversal $T\simeq S^1$ are just matrices $(a_{ij})$ where the
1443: indices are arbitrary pairs of elements $i,j$ of $S^1$ belonging
1444: to the same leaf.
1445:
1446: The algebraic rules are the same as for ordinary matrices. In the
1447: above situation, since the equivalence is given by a group action,
1448: the construction coincides with the crossed product. For
1449: instance, in the topological category,
1450: $\cA_\theta$ is identified with the crossed product $C^*$-algebra
1451: \begin{equation}\label{irr-rot-alg}
1452: \cA_\theta = C(S^1) \rtimes_{R_\theta} \Z.
1453: \end{equation}
1454:
1455: \smallskip
1456:
1457: The algebra (\ref{algNCT}) has two natural generators:
1458: \begin{equation} \label{U} U=\left\{ \begin{array}{ll}
1459: 1 & n=1 \\ 0 & \text{otherwise} \end{array}\right. \end{equation}
1460: and
1461: \begin{equation} \label{VV} V=\left\{ \begin{array}{ll}
1462: e^{2\pi ia} & n=0 \\ 0 & \text{otherwise}
1463: \end{array} \right. \end{equation}
1464: In fact, an element $b=(a_{ij})$ of $\cA_\theta$ can be written as
1465: power series
1466: \begin{equation}
1467: b = \sum_{n \in \Z} b_n U^n ,
1468: \label{powerU}
1469: \end{equation}
1470: where each $b_n$ is an element of the algebra (\ref{S1}), with
1471: the multiplication rule given by
1472: \begin{equation}
1473: U h U^{-1} = h \circ R_{\theta}^{-1} \, . \label{RthetaU}
1474: \end{equation}
1475: The algebra \eqref{S1} is generated by the function $V$ on
1476: $S^1$,
1477: \begin{equation}
1478: V(\alpha) = \exp (2\pi i \alpha) \ \ \ \forall \alpha \in S^1
1479: \label{Valpha}
1480: \end{equation}
1481: and it follows that $\cA_\theta$ is generated by two unitaries
1482: $(U,V)$ with presentation given by the relation
1483: \begin{equation}
1484: VU = \lambda \, U V, \ \ \text{ with } \ \ \lambda = \exp (2\pi i \theta) \, .
1485: \label{[U,V]}
1486: \end{equation}
1487:
1488: \smallskip
1489:
1490: If we work in the smooth category, then a generic element
1491: $b$ in \eqref{S1} is given by a power series
1492: \begin{equation}
1493: b = \sum_{\Z^2} b_{nm} U^n V^m \ \in \cS (\Z^2)
1494: \label{smoothNCt}
1495: \end{equation}
1496: where $\cS (\Z^2)$ is the Schwartz space of sequences of rapid
1497: decay on $\Z^2$. We refer to the algebra of smooth functions \eqref{smoothNCt}
1498: as $\cC^\infty (\bT^2_\theta)$, where we think of
1499: $\bT^2_\theta$ as the (smooth) non--commutative torus.
1500:
1501: \smallskip
1502:
1503: Notice that in the definition (\ref{algNCT}) it is not necessary
1504: to restrict to the condition that $i,j$ lie on the transversal
1505: $T$. It is possible to also form an algebra
1506: \begin{equation} \cB_\theta=\{ \left( a_{ij}
1507: \right) \, \, \, \, i,j\in T^2 \, \, \text{ in the same leaf } \},
1508: \end{equation}
1509: where now the parameter of integration is no longer discrete. This
1510: ought to correspond to the same non--commutative space, and in
1511: fact the algebras are related by
1512: $$ \cB_\theta = \cA_\theta \otimes \cK, $$
1513: where $\cK$ is the algebra of all compact operators.
1514:
1515: \medskip
1516:
1517: The tangent space to the ordinary torus $T^2$ is spanned by the
1518: tangent directions $\frac{\partial}{\partial x}$ and
1519: $\frac{\partial}{\partial y}$. By choosing coordinates $U,V$, with
1520: $U=e^{2\pi i x}$ and $V=e^{2\pi i y}$, the tangent vectors are
1521: given by $\frac{\partial}{\partial x}=2\pi i
1522: U\frac{\partial}{\partial U}$ and $\frac{\partial}{\partial
1523: y}=2\pi i V\frac{\partial}{\partial V}$. These have analogs in
1524: terms of derivations of the algebra of the non--commutative torus.
1525: The two commuting vector fields which span the tangent space for
1526: an ordinary (commutative) 2-torus correspond algebraically to two
1527: commuting derivations of the algebra of smooth functions.
1528:
1529:
1530: These derivations continue to make sense when we replace the
1531: generators $U$ and $V$ of $C^{\infty} (\bT^2)$ by the generators of
1532: the algebra $\cC^{\infty} (\bT_{\theta}^2)$, which no longer
1533: commute, as shown in \eqref{[U,V]}. The derivations are still
1534: given by the same formulas as in the commutative case,
1535: \begin{equation}
1536: \delta_1 = 2\pi i U \, \frac{\partial}{\partial U} \ \ \ \
1537: \delta_2 = 2\pi i V \, \frac{\partial}{\partial V}
1538: \label{derivationsUV}
1539: \end{equation}
1540: so that $\delta_1 \left( \sum b_{nm} U^n V^m \right) = 2\pi i \sum
1541: n b_{nm} U^n V^m$, and similarly for $\delta_2$.
1542:
1543:
1544: The operators (\ref{derivationsUV}) are commuting derivations of
1545: the algebra $\cC^{\infty}(\bT_{\theta}^2)$.
1546:
1547:
1548: In fact, it is straightforward to verify that $\delta_1$ and
1549: $\delta_2$ satisfy
1550: \begin{equation}
1551: \delta_1 \delta_2 = \delta_2 \delta_1
1552: \label{der-commuting}
1553: \end{equation}
1554: and
1555: \begin{equation}
1556: \delta_j (bb') = \delta_j (b) b' + b \delta_j (b') \quad \forall b
1557: , b' \in \cA_\theta. \label{der-rule}
1558: \end{equation}
1559:
1560: Just as in the classical case of a (commutative)
1561: manifold, what ensures that the derivations considered are enough
1562: to span the whole tangent space is the condition of ellipticity
1563: for the Laplacian
1564: $$ \Delta = \delta_1^2 + \delta_2^2. $$
1565: In Fourier modes the Laplacian is of the form $n^2 + m^2$, hence
1566: $\Delta^{-1}$ is a compact operator.
1567:
1568:
1569: \medskip
1570:
1571: The geometry of the Kronecker foliation is closely related to the
1572: structure of the algebra. In fact, a choice of a {\it closed
1573: transversal} $T$ of the foliation corresponds
1574: canonically to a {\it finite projective module} over the algebra
1575: $\cA_\theta$.
1576:
1577: In fact, the main result on finite projective module over the
1578: non--commutative tori $\bT_{\theta}^2$ is the following
1579: classification, which is obtained by combining the results of
1580: \cite{PV}, \cite{C3}, \cite{Rieffel}.
1581:
1582:
1583: \begin{thm}\label{NCTmodules}
1584: Finite projective modules over $\cA_{\theta}$ are classified up to
1585: isomorphism by a pair of integers $(p,q)$ such that $p+q \theta
1586: \geq 0$. For a choice of such pair, the corresponding module
1587: ${\mathcal H}_{p,q}^{\theta}$ is obtained from the transversal
1588: $T_{p,q}$ given by the closed geodesic of the torus $T^2$
1589: specified by $(p,q)$, via the following construction. Elements of
1590: the module associated to the transversal $T_{p,q}$ are rectangular
1591: matrices, $(\xi_{i,j})$ with $(i,j) \in T \times S^1$, and with
1592: $i$ and $j$ belonging to the same leaf. The right action of
1593: $(a_{i,j}) \in \cA_{\theta}$ is by matrix multiplication.
1594: \end{thm}
1595:
1596: For instance, from the transversal $x=0$ one obtains the following
1597: right module over $\cA_\theta$. The underlying linear space is the
1598: usual Schwartz space
1599: \begin{equation}
1600: \cS (\R) = \{ \xi : \, \xi (s) \in \C, \quad \forall s \in \R
1601: \} \label{SRmodule}
1602: \end{equation}
1603: of complex valued smooth functions on $\R$, all of whose
1604: derivatives are of rapid decay.
1605: The right module structure is given by the action of
1606: the generators $U,V$
1607: \begin{equation}
1608: (\xi U) (s) = \xi (s+\theta) \ \ \ \ (\xi V) (s) = e^{2\pi is} \xi
1609: (s) \ \ \ \forall s \in \R \, .
1610: \label{actionUV}
1611: \end{equation}
1612: One of course checks that the relation \eqref{[U,V]} is satisfied,
1613: and that, as a right module over
1614: $\cA_\theta$, the space $\cS (\R)$ is {\it finitely generated}
1615: and {\it projective} (\ie it complements to a free module).
1616:
1617: \smallskip
1618:
1619: Finitely generated projective modules play an important role in
1620: noncommutative geometry, as they replace {\em vector bundles} in the
1621: commutative setting. In fact, in ordinary commutative geometry,
1622: one can equivalently describe vector bundles through their sections,
1623: which in turn form a finite projective module over the algebra of
1624: smooth functions. The
1625: notion of finite projective module continues to make sense in the
1626: non--commutative setting, and provides this way a good notion of
1627: ``non--commutative vector bundles''.
1628:
1629: \smallskip
1630:
1631: Suppose given a vector bundle $E$, described algebraically through its
1632: space of smooth sections $C^{\infty} (X,E)$. One can
1633: compute the dimension of $E$ by computing the trace of the identity
1634: endomorphism. In terms of the space of smooth sections, hence of finite
1635: projective modules, it is possible to recover the dimension of the
1636: vector bundle as a limit
1637: \begin{equation} \label{dimE}
1638: \dim_{\cA}(\cE)= \lim_{N\to \infty}\frac{1}{N}\left( \#
1639: \text{Generators of} \, \, \underbrace{\cE \oplus \cdots \oplus
1640: \cE}_{N \text{times}} \right).
1641: \end{equation}
1642: This method applies to the noncommutative setting. In the case
1643: of noncommutative tori, one finds that the Schwartz space $\cS(\R)$
1644: has dimension the real number
1645: \begin{equation}
1646: \dim_{\cB} (\cS) = \theta \, .
1647: \label{dimtheta}
1648: \end{equation}
1649: One similarly finds values $p+q\theta$ for the more general
1650: case of Theorem \ref{NCTmodules}.
1651:
1652: \smallskip
1653:
1654: The appearance of a real values dimension is related to the {\em
1655: density} of transversals in the leaves, that is, the limit of
1656: $$ \frac{\# B_R \cap S}{\text{size of}\, \, B_R}, $$
1657: for a ball $B_R$ of radius $R$ in the leaf. In this sense, the
1658: dimension $\theta$ of the Schwartz space measure the relative
1659: densities of the two transversals $S=\{ x=0 \}$ and $T=\{ y=0 \}$.
1660:
1661:
1662: In general, the appearance of non integral dimension is a basic
1663: feature of von
1664: Neumann algebras of type II. The dimension of a vector bundle
1665: is the only invariant that remains when one looks from the measure
1666: theoretic point of view, using the algebra of measurable functions
1667: $L^\infty(S^1)$ in (\ref{S1}). The von Neumann algebra which
1668: describes the quotient space from the measure theoretic point of
1669: view is the crossed product
1670: \begin{equation}
1671: R = L^{\infty} (S^1) \rtimes_{R_{\theta}} \Z.
1672: \label{measNCT}
1673: \end{equation}
1674: This is the well known hyperfinite factor of type II$_1$. In
1675: particular the classification of finite projective modules $\cE$
1676: over $R$ is given by a positive real number, the Murray and von
1677: Neumann dimension
1678: \begin{equation}
1679: \dim_R (\cE) \in \R_+ \, . \label{MvNR}
1680: \end{equation}
1681:
1682: \medskip
1683:
1684: The simplest way to describe the phenomenon of Morita equivalence
1685: for non--commutative tori is in terms of the Kronecker foliation,
1686: where it corresponds to reparameterizing the leaves space in terms
1687: of a different closed transversal. Thus, Morita equivalence of the
1688: algebras $\cA_{\theta}$ and $\cA_{\theta'}$ for $\theta$ and
1689: $\theta'$ in the same $\PGL(2,\Z)$ orbit becomes simply a
1690: statement that the leaf--space of the original foliation is
1691: independent of the transversal used to parameterize it. For
1692: instance, Morita equivalence between $\cA_{\theta}$ and
1693: $\cA_{-1/\theta}$ corresponds to changing the parameterization of
1694: the space of leaves from the transversal $T=\{ y=0 \}$ to the
1695: transversal $S=\{ x=0 \}$.
1696:
1697: More generally, an explicit construction of bimodules
1698: $\cM_{\theta,\theta'}$ was obtained in \cite{C3}. These are
1699: given by the Schwartz space ${\mathcal S}(\R \times \Z/c)$, with
1700: the right action of $\cA_\theta$ given by
1701: $$ Uf\, (x,u)=f\left( x-\frac{c\theta+d}{c}, u-1\right) $$
1702: $$ Vf\, (x,u)=\exp(2\pi i(x-u d/c)) f(x,u) $$
1703: and the left action of $\cA_{\theta'}$
1704: $$ U'f\, (x,u)=f\left( x-\frac{1}{c}, u-a \right) $$
1705: $$ V'f\, (x,u)=\exp\left( 2\pi i \left(\frac{x}{c\theta+d}
1706: -\frac{u}{c}\right)\right) f(x,u). $$
1707: The bimodule $\cM_{\theta,\theta'}$ realizes the
1708: Morita equivalences between $\cA_\theta$ and
1709: $\cA_{\theta'}$ for
1710: $$ \theta'=\frac{a\theta + b}{c\theta + d}=g\theta $$
1711: with $g\in \PGL(2,\Z)$, \cf \cite{C3}, \cite{Ri2}.
1712:
1713: \bigskip
1714:
1715:
1716: \section{Duals of discrete groups}\label{discgr}
1717:
1718:
1719: Noncommutative geometry provides naturally
1720: a generalization of Pontrjagin duality for
1721: discrete groups. While the Pontrjagin dual $\hat\Gamma$ of a
1722: finitely generated discrete abelian group is a compact abelian group,
1723: the dual of a more general finitely generated discrete group
1724: is a noncommutative space.
1725:
1726: To see this, recall that the usual Pontrjagin duality assigns to
1727: a finitely generated discrete abelian group $\Gamma$ its group of
1728: characters $\hat\Gamma=\Hom(\Gamma,U(1))$. The duality
1729: is given by Fourier transform $e^{i\langle k, \gamma\rangle}$, for
1730: $\gamma\in \Gamma$ and $k\in \hat\Gamma$.
1731:
1732:
1733: In particular, Fourier transform gives an identification between
1734: the algebra of functions on
1735: $\hat\Gamma$ and the (reduced) $C^*$-algebra of
1736: the group $\Gamma$,
1737: \begin{equation}\label{CGammadual}
1738: C(\hat\Gamma) \cong C^*_r(\Gamma),
1739: \end{equation}
1740: where the reduced $C^*$-algebra $C^*_r(\Gamma)$ is the
1741: $C^*$-algebra generated by $\Gamma$ in the regular representation
1742: on $\ell^2(\Gamma)$.
1743:
1744:
1745: When $\Gamma$ is non-abelian Pontrjagin duality no
1746: longer applies in the classical sense. However, the left hand side of
1747: \eqref{CGammadual} still makes sense and it behaves ``like''
1748: the algebra of functions on the dual group. One can then
1749: say that, for a non-abelian group, the Pontrjagin dual
1750: $\hat\Gamma$ still exists as a noncommutative space whose algebra
1751: of coordinates is the $C^*$-algebra $C^*_r(\Gamma)$.
1752:
1753: \medskip
1754:
1755: As an example that illustrates this general philosophy we
1756: give a different version of Example \ref{ex2}.
1757:
1758: \begin{ex}{\em
1759: The algebra (\ref{alg1}) of Example \ref{ex2}, is the group ring of
1760: the dihedral group $\Z \rtimes \Z/2 \cong \Z/2 * \Z/2$. }
1761: \label{dihedral}
1762: \end{ex}
1763:
1764: In fact, first notice that to a representation of the group $\Z/2 *
1765: \Z/2$ (free product of two copies of the group with two elements) is
1766: the same thing as a pair of subspaces in the Hilbert space,
1767: $E,F\subset \cH$. The corresponding operators are $U=I-2P_E$,
1768: $V=I-2P_V$, with $P_E,P_V$ the projections. The operators $U,V$
1769: represent reflections, since $U=U^*$, $U^2=I$, $V=V^*$, $V^2=I$. The
1770: group $\Gamma =\Z/2 * \Z/2$ realized as words in the generators $U$
1771: and $V$ can equivalently be described as the semi-direct product
1772: $\Gamma =\Z \rtimes \Z/2$, by setting $X=UV$, with the action
1773: $UXU^{-1}=X^{-1}$. The regular representation of $\Gamma$ is
1774: analyzed using Mackey's theory for semi-direct products. 0ne
1775: considers first representations of the normal subgroup, and then
1776: orbits of the action of $\Z/2$. The irreducible representations of
1777: the normal subgroup $\Z$ are labeled by $S^1=\{z\in\C\,|z|=1\}$ and
1778: given by $X^n\mapsto z^n$. The action of $\Z/2$ is the involution given
1779: by conjugation $z \mapsto \bar z$. The quotient of $S^1$ by the
1780: $\Z/2$ action is identified with the interval $[-1,1]$ by the map
1781: $z\mapsto \Re( z)$. For points inside the interval the corresponding
1782: irreducible representation of $\Gamma$ is two dimensional. At each
1783: of the two endpoints $\pm 1$ one gets two inequivalent irreducible
1784: representations of $\Gamma$. Thus we recover the picture of Example
1785: \ref{ex2} and an isomorphism $C^*(\Gamma)\sim A$ where $A$ is the
1786: algebra \eqref{alg1}.
1787:
1788: \medskip
1789: The first two basic steps of the general theory are known
1790: for arbitrary discrete groups $\Gamma$, namely
1791:
1792: 1) The
1793: resolution of the diagonal and computation of the cyclic cohomology
1794: are provided by the geometric model (due to Burghelea \cite{Burgh}) given by the
1795: free loop space of the classifying space $B\Gamma$.
1796:
1797: 2) The assembly map (BC-map) of \cite{BauCo}
1798: from the K-homology of the classifying
1799: space $B\Gamma$ to the K-theory of the reduced $C^*$-algebra
1800: $C^*_r(\Gamma)$ is refined in \cite{BCH} to take care of torsion
1801: in the group
1802: $\Gamma$ and gives a pretty good approximation to the K-theory of
1803: $C^*_r(\Gamma)$ (\cf \cite{Skan1} and references therein).
1804:
1805: In the presence of a natural
1806: smooth subalgebra of $C^*_r(\Gamma)$ containing the group ring and
1807: stable under holomorphic functional calculus, the combination of
1808: the two steps described above makes it possible to prove an index
1809: theorem which is an higher dimensional form of Atiyah's $L^2$-index
1810: theorem for coverings. This gave the first proof of the Novikov
1811: conjecture for hyperbolic groups (\cite{CoMoind}). Since then the
1812: analysis of dense smooth subalgebras has played a key role, in
1813: particular in the ground breaking work of Vincent Lafforgue.
1814: See \cite{BauCo}, \cite{Julg}, \cite{Laff}, \cite{Skan1}, \cite{Skan2}.
1815:
1816:
1817: The next step, \ie the construction of a spectral geometry, is
1818: directly related to geometric group theory. In general one cannot
1819: expect to get a finite dimensional spectral triple since the growth
1820: properties of the group give (except for groups of polynomial
1821: growth) a basic obstruction (\cf \cite{Co-fr}). A general
1822: construction of a theta summable spectral triple was given in
1823: \cite{Co94} Section IV.9. Basically the transition from finitely
1824: summable spectral triples to the theta summable ones is the
1825: transition from finite dimensional geometry to the infinite
1826: dimensional case. In the theta summable case the Chern character is
1827: no longer a finite dimensional cyclic cocycle and one needs to
1828: extend cyclic cohomology using cocycles with infinite support in the
1829: $(b,B)$ bicomplex fulfilling a subtle growth condition. The general
1830: theory of {\em entire} cyclic cohomology was developed in
1831: \cite{Cotheta}. It is in general quite difficult to compute the
1832: Chern character in the theta summable case and one had to wait quite
1833: a long time until it was done for the basic example of discrete
1834: subgroups of semi-simple Lie groups. This has been achieved in a
1835: remarkable recent paper of M. Puschnigg \cite{Pusch} in the case of
1836: real rank one.
1837:
1838: The fourth step \ie the thermodynamics might seem irrelevant in the
1839: type II context of discrete groups. However a small variant of the
1840: construction of the group ring, namely the Hecke algebra associated
1841: to an almost normal inclusion of discrete groups (in the sense
1842: considered in \cite{BC})
1843: suffices to meet the type III world. One of the open fields is to
1844: extend the above steps 1), 2) and 3) in the general context of
1845: almost normal inclusions of discrete groups, and to perform the
1846: thermodynamical analysis in the spirit of \cite{CCM} in that
1847: context.
1848:
1849:
1850: \bigskip
1851:
1852:
1853: \section{Brillouin zone and the quantum Hall effect}\label{qhe}
1854:
1855:
1856:
1857: An important application to physics of the theory of
1858: non--commutative tori was the development of a rigorous
1859: mathematical model for the Integer Quantum Hall Effect (IQHE)
1860: obtained by Bellissard and collaborators \cite{Bellissard1},
1861: \cite{Bellissard2}, \cite{Co94}.
1862:
1863: The classical Hall effect is a physical phenomenon first observed
1864: in the XIX century \cite{Hall}. A very thin metal sample is
1865: immersed in a constant uniform strong magnetic field orthogonal to
1866: the surface of the sample. By forcing a constant current to flow
1867: through the sample, the flow of charge carriers in the metal is
1868: subject to a Lorentz force perpendicular to the current and the
1869: magnetic field. The equation for the equilibrium of forces in the
1870: sample is of the form
1871: \begin{equation} \label{clHall} N e {\bf E} + {\bf j} \wedge {\bf B}
1872: =0, \end{equation} where ${\bf E}$ is the electric field, $e$ and
1873: $N$ the charge and number of the charge carriers in the metal,
1874: ${\bf B}$ the magnetic field, and ${\bf j}$ the current.
1875:
1876:
1877: \begin{center}
1878: \begin{figure}
1879: \includegraphics[scale=1.1]{qhe.eps}
1880: \caption{The quantum Hall effect experiment.
1881: \label{FigQHE}}
1882: \end{figure}
1883: \end{center}
1884:
1885:
1886: The equation \eqref{clHall} defines a linear relation: the ratio
1887: of the intensity of the Hall current to the intensity of the
1888: electric field defines the Hall conductance,
1889: \begin{equation} \label{condHall} \sigma_H = \frac{N e \delta}{B},
1890: \end{equation}
1891: with $B=|{\bf B}|$ the intensity of the magnetic field and
1892: $\delta$ the sample width. The dimensionless quantity
1893: \begin{equation} \label{filling} \nu_H = \frac{N\delta h}{B e} =
1894: \sigma_H R_H \end{equation} is called the filling factor, while
1895: the quantity $R_H = h/e^2$ is the Hall resistance. The filling
1896: factor measures the fraction of Landau level filled by conducting
1897: electrons in the sample. Thus, classically, the Hall conductance,
1898: measured in units of $e^2/h$, equals the filling factor.
1899:
1900: In 1980, about a century after the classical Hall effect was
1901: observed, von Klitzing's experiment showed that, lowering the
1902: temperature below 1 K, the relation of Hall conductance to filling
1903: factor shows plateaux at integer values, \cite{vonKlitz}. The
1904: integer values of the Hall conductance are observed with a
1905: surprising experimental accuracy of the order of $10^{-8}$. This
1906: phenomenon of quantization of the Hall conductance is known as
1907: Integer Quantum Hall Effect (IQHE).
1908:
1909: Laughlin first suggested that IQHE should be of a geometric origin
1910: \cite{Lau}. A detailed mathematical model of the IQHE, which
1911: accounts for all the important features of the experiment
1912: (quantization, localization, insensitivity to the presence of
1913: disorder, vanishing of direct conductance at plateaux levels)
1914: improving over the earlier Laughlin model, was developed by
1915: Bellissard and collaborators \cite{Bellissard1},
1916: \cite{Bellissard2}.
1917:
1918: Bellissard's approach to the IQHE is based on non--commutative geometry.
1919: The quantization of the Hall conductance at integer values is
1920: indeed geometric in nature: it resembles another well known
1921: ``quantization'' phenomenon that happens in the more familiar
1922: setting of the geometry of compact 2--dimensional manifolds,
1923: namely the Gauss--Bonnet theorem, where the integral of the
1924: curvature is an integer multiple of $2\pi$, a property that is
1925: stable under deformations. In the same spirit, the values of the
1926: Hall conductance are related to the evaluation of a certain
1927: characteristic class, or, in other words, to an index theorem for
1928: a Fredholm operator.
1929:
1930: More precisely, in the physical model one makes the simplifying
1931: assumption that the IQHE can be described by non--interacting
1932: particles. The Hamiltonian then describes the motion of a single
1933: electron subject to the magnetic field and an additional potential
1934: representing the lattice of ions in the conductor. In a perfect
1935: crystal and in the absence of a magnetic field, there is a group
1936: of translational symmetries. This corresponds to a group of
1937: unitary operators $U(a)$, $a\in G$, where $G$ is the locally
1938: compact group of symmetries. Turning on the magnetic field breaks
1939: this symmetry, in the sense that translates of the Hamiltonian
1940: $H_a = U(a) H U(a)^{-1}$ no longer commute with the Hamiltonian
1941: $H$. Since there is no preferred choice of one translate over the
1942: others, the algebra of observables must include all translates of
1943: the Hamiltonian, or better their resolvents, namely the bounded
1944: operators
1945: \begin{equation} \label{translR} R_a(z) = U(a) (zI - H)^{-1}
1946: U(a)^{-1}. \end{equation}
1947:
1948: For a particle of (effective) mass $m$ and charge $e$ confined to
1949: the plane, subject to a magnetic field of vector potential ${\bf
1950: A}$ and to a bounded potential $V$, the Hamiltonian is of the form
1951: \begin{equation} \label{magnLapl} H= \frac{1}{2m} \sum_{j=1,2} (p_j -e
1952: A_j)^2 + V = H_0 + V, \end{equation} where the unperturbed part
1953: $H_0$ is invariant under the magnetic translations, namely the
1954: unitary representation of the translation group $\R^2$ given by
1955: $$ U(a) \psi(x) = \exp\left( \frac{-ieB}{2\hbar} \omega(x,a) \right)
1956: \psi(x-a), $$ with $\omega$ the standard symplectic form in the
1957: plane. The hull (strong closure) of the translates (\ref{translR})
1958: yields a topological space, whose homeomorphism type is
1959: independent of the point $z$ in the resolvent of $H$. This
1960: provides a non--commutative version of the Brillouin zone.
1961:
1962: Recall that the Brillouin zones of a crystals are fundamental
1963: domains for the reciprocal lattice $\Gamma^\sharp$ obtained via
1964: the following inductive procedure. The {\em Bragg hyperplanes} of
1965: a crystal are the hyperplanes along which a pattern of diffraction
1966: of maximal intensity is observed when a beam of radiation (X-rays
1967: for instance) is shone at the crystal. The $N$-th Brillouin zone
1968: consists of all the points in (the dual) $\R^d$ such that the line
1969: from that point to the origin crosses exactly $(n-1)$ Bragg
1970: hyperplanes of the crystal.
1971:
1972: \begin{figure}
1973: \begin{center}
1974: \includegraphics[scale=0.5]{2dBZonesBW.ps}
1975: \end{center}
1976: \caption{Brillouin zones for a 2-dimensional crystal
1977: \label{FigBrill}}
1978: \end{figure}
1979:
1980: More precisely, in our case, if $e_1$ and $e_2$ are generators of the
1981: periodic lattice, we obtain a commutation relation
1982: $$ U(e_1) U(e_2) = e^{2\pi i \theta} U(e_2) U(e_1), $$
1983: where $\theta$ is the flux of the magnetic field through a
1984: fundamental domain for the lattice, in dimensionless units, hence
1985: the non-commutative Brillouin zone is described by a
1986: non-commutative torus.
1987:
1988: This can also be seen easily in a discrete model, where the
1989: Hamiltonian is given by an operator
1990: \begin{equation}\label{discr-Ham}
1991: \begin{array}{rll}
1992: (H_a \, f)(m,n) = & e^{-ia_1 n} f(m+1,n) + & e^{ia_2 n} f(m-1,n)
1993: \\[3mm] + &
1994: e^{-ia_2 m} f(m, n+1) + & e^{ia_2 m } f(m, n-1),
1995: \end{array}
1996: \end{equation}
1997: for $f\in L^2(\Z^2)$. This is a discrete version of the magnetic
1998: Laplacian. Notice then that \eqref{discr-Ham} can be written in
1999: the form
2000: $$ H_a = U+V + U^* + V^*, $$
2001: for
2002: $$ (U\, f)(m,n)= e^{-ia_2m} f(m,n+1) \ \ \ \ (V\, f)(m,n)=
2003: e^{-ia_1 n} f(m+1,n).
2004: $$
2005: These clearly satisfy the commutation relation \eqref{[U,V]} of
2006: $\bT^2_\theta$ with $\theta= a_2-a_1$.
2007:
2008:
2009: In the zero-temperature limit, the Hall conductance satisfies the
2010: Kubo formula
2011: \begin{equation}\label{kubo} \sigma_H = \frac{1}{2\pi i R_H}
2012: \tau(P_\mu [\delta_1 P_\mu, \delta_2 P_\mu]),
2013: \end{equation}
2014: where $P_\mu$ is a spectral projection of the Hamiltonian on
2015: energies smaller or equal to the Fermi level $E_\mu$, $\tau$ is
2016: the trace on $\cA_\theta$ given by
2017: \begin{equation} \label{trace-theta}
2018: \tau\left( \sum a_{n,m} U^n V^m \right) = a_{0,0}.
2019: \end{equation}
2020: and $\delta_1$, $\delta_2$ are as in
2021: \eqref{derivationsUV}. Here we assume that the Fermi level $\mu$
2022: is in a gap in the spectrum of the Hamiltonian. Then the spectral
2023: projections $P_\mu$ belong to the $C^*$-algebra of observables.
2024:
2025: The Kubo formula \eqref{kubo} can be derived from purely physical
2026: considerations, such as transport theory and the quantum adiabatic
2027: limit.
2028:
2029: The main result then is the fact that the integrality of the
2030: conductance observed in the Integer Quantum Hall Effect is explained
2031: topologically, that is, in terms of the integrality of the {\em
2032: cyclic cocycle} $\tau(a^0 (\delta_1 a^1 \delta_2 a^2 - \delta_2 a^1
2033: \delta_1 a^2))$ (\cf \cite{C3}).
2034:
2035: \medskip
2036:
2037: \begin{center}
2038: \begin{figure}
2039: \includegraphics[scale=0.6]{FQHE.eps}
2040: \caption{Observed fractions in the quantum Hall effect.
2041: \label{FigFQHE}}
2042: \end{figure}
2043: \end{center}
2044:
2045:
2046: The fractional QHE was discovered by Stormer and Tsui in 1982. The
2047: setup is as in the quantum Hall effect: in a high quality
2048: semi-conductor interface, which will be modelled by an infinite
2049: 2-dimensional surface, with low carrier concentration and
2050: extremely low temperatures $\sim 10 mK$, in the presence of a very
2051: strong magnetic field, the experiment shows that the graph of
2052: $\frac{h}{e^2}\sigma_H$ against the filling factor $\nu$ exhibits
2053: plateaux at certain fractional values.
2054:
2055: The independent electron approximation that, in the case of
2056: the integer quantum Hall effect, reduces the problem to a single
2057: electron wavefunction is no longer viable in this case and one has to
2058: incorporate the Coulomb interaction between the electrons in a
2059: many-electron theory. Nonetheless, it is possible to use a crude
2060: approximation, whereby one alters the underlying geometry to
2061: account for an average effect of the multi-electron interactions.
2062: One can obtain this way a model of the fractional quantum Hall effect
2063: via noncommutative geometry (\cf \cite{MM1}, \cite{MM2}, \cite{MM3}),
2064: where one uses hyperbolic geometry to simulate the interactions.
2065:
2066: The noncommutative geometry approach to the quantum Hall effect
2067: described above was extended to hyperbolic geometry in \cite{CHMM}.
2068: The analog of the operator \eqref{discr-Ham} is given by the Harper
2069: operator on the Cayley graph of a finitely generated discrete subgroup
2070: $\Gamma$ of $\PSL_2(\R)$. Given $\sigma: \Gamma \times
2071: \Gamma \to U(1)$ satisfying $\sigma(\gamma,1)=\sigma(1,\gamma)=1$
2072: and
2073: $$ \sigma(\gamma_1,\gamma_2)\sigma(\gamma_1\gamma_2,\gamma_3)=
2074: \sigma(\gamma_1,\gamma_2\gamma_3)\sigma(\gamma_2,\gamma_3), $$
2075: one considers the right $\sigma$-regular representation on
2076: $\ell^2(\Gamma)$ of the form
2077: \begin{equation}\label{rightreg}
2078: R^\sigma_\gamma
2079: \psi(\gamma') = \psi(\gamma' \gamma) \sigma(\gamma', \gamma)
2080: \end{equation}
2081: satisfying $$R^\sigma_\gamma
2082: R^\sigma_{\gamma'} = \sigma(\gamma,\gamma')
2083: R^\sigma_{\gamma\gamma'}.$$
2084: For $\{\gamma_i\}_{i=1}^r$ a symmetric set of generators of
2085: $\Gamma$, the Harper operator is of the form
2086: \begin{equation}\label{HarperGamma}
2087: \cR_\sigma = \sum_{i=1}^r R^\sigma_{\gamma_i},
2088: \end{equation}
2089: and the operator $r-\cR_\sigma$ is the discrete analog of the magnetic
2090: Laplacian (\cf \cite{Sun}).
2091:
2092: The idea is that, by effect of the strong
2093: interaction with the other electrons, a single electron ``sees''
2094: the surrounding geometry as hyperbolic, with lattice sites that
2095: appear (as a multiple image effect) as the points in a lattice
2096: $\Gamma\subset \PSL_2(\R)$. Thus, one considers the general form
2097: of such a lattice
2098: \begin{equation}\label{FuchsianG}
2099: \Gamma = \Gamma(g; \nu_1, \ldots, \nu_n),
2100: \end{equation}
2101: with generators $a_i, b_i, c_j$, with $i=1,\ldots,g$ and
2102: $j=1,\ldots,n$ and a presentation of the form
2103: \begin{equation}\label{presentation}
2104: \Gamma(g; \nu_1, \ldots, \nu_n)=\langle a_i, b_i, c_j\,\,
2105: \left|\,\, \prod_{i=1}^g [a_i,b_i]c_1\cdots c_n =1, \,\,\,
2106: c_j^{\nu_j} =1 \right. \rangle.
2107: \end{equation}
2108: The quotient of the action of $\Gamma$ by isometrieson $\H$,
2109: \begin{equation}\label{hyporb}
2110: \Sigma(g; \nu_1, \ldots, \nu_n):= \Gamma \backslash \H,
2111: \end{equation}
2112: is a hyperbolic orbifold.
2113:
2114: \begin{center}
2115: \begin{figure}
2116: \includegraphics[scale=0.6]{orbifold.eps}
2117: \caption{Hyperbolic orbifolds.
2118: \label{FigOrbi}}
2119: \end{figure}
2120: \end{center}
2121:
2122:
2123: Let $P_E$ denote denote the spectral projection associated to the
2124: Fermi level, \ie $P_E = \chi_{(-\infty, E]}(H)$. Then, in the zero
2125: temperature limit, the Hall conductance is given by
2126: \begin{equation}\label{sigmaE}
2127: \sigma_E = \tr_K (P_E, P_E, P_E),
2128: \end{equation}
2129: where $\tr_K$ denotes the conductance 2-cocycle. It is a
2130: cyclic 2-cocycle on the twisted group algebra $\C(\Gamma, \sigma)$
2131: of the form
2132: \begin{equation}\label{trK}
2133: \tr_K(f_0, f_1, f_2) =
2134: \sum_{j=1}^g \tr(f_0 (\delta_j(f_1)\delta_{j+g}(f_2) -
2135: \delta_{j+g}(f_1)\delta_j(f_2))),
2136: \end{equation}
2137: where the $\delta_j$ are derivations associated to the
2138: 1-cocycles $a_j$ associated to a symplectic basis
2139: $\{ a_j, b_j\}_{j=1,\ldots, g}$ of $H^1(\Gamma, \R)$ (\cf
2140: \cite{MM3}).
2141:
2142: Within this model, one obtains the fractional values of the Hall
2143: conductance as integer multiples of orbifold Euler characteristics
2144: \begin{equation}\label{orbEulch}
2145: \chi_{orb}(\Sigma(g;\nu_1,\ldots,\nu_n))=2-2g +\nu -n \in \Q.
2146: \end{equation}
2147: In fact, one shows (\cf \cite{MM2}, \cite{MM3}) that the conductance
2148: 2-cocycle is cohomologous to another cocycle, the area 2-cocycle, for
2149: which one can compute the values on $K$-theory (hence the value of
2150: \eqref{sigmaE}) by applying a twisted version of the Connes--Moscovici
2151: higher index theorem \cite{CoMoind}.
2152:
2153: \smallskip
2154:
2155: While in the case of the integer quantum Hall effect the
2156: noncommutative geometry model is completely satisfactory and explains
2157: all the physical properties of the system, in the fractional case
2158: the orbifold model can be considered
2159: as a first rough approximation to the quantum field theory that
2160: coverns the fractional quantum Hall effect. For instance, the geometry of
2161: 2-dimensional hyperbolic orbifolds is related to Chern--Simons theory
2162: through the moduli spaces of vortex equations. This remains an interesting
2163: open question.
2164:
2165: \bigskip
2166:
2167:
2168: \section{Tilings}\label{tiles}
2169:
2170: In general, by a {\em tiling} $\cT$ in $\R^d$ one means the following. One
2171: considers a finite collection $\{ \tau_1, \ldots, \tau_N \}$ of closed
2172: bounded subsets of $\R^d$ homeomorphic to the unit ball. These are
2173: called the {\em prototiles}. One usually assumes that the prototiles
2174: are polytopes in $\R^d$ with a single $d$-dimensional cell which is
2175: the interior of the prototile, but this assumption can be relaxed.
2176: A tiling $\cT$ of $\R^d$ is then a covering of $\R^d$ by sets with disjoint
2177: interior, each of which is a {\em tile}, that is, a translate of one
2178: of the prototiles.
2179:
2180: \begin{center}
2181: \begin{figure}
2182: \includegraphics[scale=0.8]{PersianArt.eps}
2183: \caption{Prototiles for tilings.
2184: \label{FigTiles}}
2185: \end{figure}
2186: \end{center}
2187:
2188:
2189: Given a tiling $\cT$ of $\R^d$ one can form its orbit closure
2190: under translations. The metric on tilings is defined by saying that
2191: two tilings are close if they almost agree on a large ball centered
2192: at the origin in $\R^d$ (for more details and equivalent definitions
2193: see \eg \cite{AndPut}, \cite{BellC}).
2194:
2195: Tilings can be periodic or aperiodic. There are many familiar examples
2196: of periodic tilings, while the best known examples of aperiodic
2197: tilings are the Penrose tilings \cite{Penrose}. Similar types of
2198: aperiodic tilings have been widely studied in the physics of
2199: quasicrystals (\cf \eg \cite{BaMo}, \cite{BellC}).
2200:
2201: It was understood very early on in the development of noncommutative
2202: geometry (\cf \cite{Co-fr} and pp.5--7, pp.88--93, and pp.175--178 of
2203: \cite{Co94}) that Penrose tilings provide an interesting class of
2204: noncommutative spaces.
2205:
2206: In fact, one can consider on the set $\Omega$ of tilings $\cT$ with
2207: given prototiles $\{ \tau_1, \ldots, \tau_N \}$ the equivalence relation
2208: given by the action of $\R^d$ by translations, \ie one identifies
2209: tilings that can be obtained from one another by translations. In the
2210: case of aperiodic tilings, this yields the type of quotient
2211: construction described in Section \ref{quotients}, which leads
2212: naturally to noncommutative spaces. An explicit description of this
2213: noncommutative space for the case of Penrose tilings can be found in
2214: \S II.3 of \cite{Co94}.
2215:
2216: \begin{center}
2217: \begin{figure}
2218: \includegraphics[scale=0.6]{zellijs.eps}
2219: \caption{Quasiperiodic tilings and zellijs.
2220: \label{FigZell}}
2221: \end{figure}
2222: \end{center}
2223:
2224: To simplify the picture slightly, we can consider the similar problem
2225: (dually) with arrangements of points in $\R^d$ instead of
2226: tilings. This is the formulation used in the theory of aperiodic
2227: solids and quasicrystals (\cf \cite{BellC}). Then, instead of tilings
2228: $\cT$, we consider discrete subsets of points $\cL\subset \R^d$. Such
2229: $\cL$ is a Delauney set if there are radii $r,R>0$ such that every
2230: open ball of radius $r$ meets $\cL$ in at most one point and every
2231: closed ball of radius $R$ meets $\cL$ in at least one point. One can
2232: describe $\cL$ by the counting measure
2233: $$ \mu_{\cL}(f) =\sum_{x\in\cL} f(x), $$
2234: and one can take the orbit closure $\Omega$ of the action of $\R^d$ by
2235: translations $$\mu_{\cL}\mapsto T_{-a}\mu_{\cL}=\mu_{\cL}\circ T_a, \
2236: \ \ \text{ for } a\in \R^d, $$
2237: in the space $\cM(\R^d)$ of Radon measures with the weak$^*$
2238: topology. The Hull of $\cL$ is the dynamical system $(\Omega,T)$,
2239: where $T$ denotes the action of $\R^d$ by translations.
2240:
2241: This dynamical system determines a corresponding noncommutative space,
2242: describing the quotient of $\Omega$ by translations, namely the
2243: crossed product $C^*$-algebra
2244: \begin{equation}\label{crossOmega}
2245: \cA=C(\Omega)\rtimes_T \R^d.
2246: \end{equation}
2247: In fact, one can also consider the groupoid with set of units the
2248: transversal
2249: \begin{equation}\label{transvOmega}
2250: X=\{ \omega\in \Omega: 0\in {\rm Support}(\omega)\},
2251: \end{equation}
2252: arrows of the form $(\omega,a)\in \Omega\times \R^d$, with
2253: source and range maps $s(\omega,a)=T_{-a}\omega$, $r(\omega,a)=\omega$
2254: and $(\omega,a)\circ (T_{-a}\omega,b)=(\omega, a+b)$ (\cf
2255: \cite{BellC}). This defines a locally compact groupoid $\cG(\cL,X)$.
2256: The $C^*$-algebras $C^*(\cG(\cL,X))$ and $C(\Omega)\rtimes_T \R^d$ are
2257: Morita equivalent.
2258:
2259: In the case where $\cL$ is a periodic arrangement of points with
2260: cocompact symmetry group $\Gamma\subset \R^d$, the space
2261: $\Omega$ is an ordinary commutative space, which is topologically
2262: a torus $\Omega=\R^d/\Gamma$. The $C^*$-algebra $\cA$ is in this case
2263: isomorphic to $C(\hat\Gamma)\otimes \cK$, where $\cK$ is the algebra
2264: of compact operators and $\hat\Gamma$ is the Pontrjagin dual of the
2265: abelian group $\Gamma\cong \Z^d$, isomorphic to $T^d$, obtained by taking
2266: the dual of $\R^d$ modulo the reciprocal lattice
2267: \begin{equation}\label{recLat}
2268: \Gamma^\sharp = \{ k \in \R^d : \langle k, \gamma\rangle\in 2
2269: \pi \Z, \forall \gamma \in \Gamma \}.
2270: \end{equation}
2271: Thus, in physical language, $\hat\Gamma$ is identified with the
2272: Brillouin zone $B=\R^d/\gamma^\sharp$ of the periodic
2273: crystal $\cL$ (\cf Section \ref{qhe}). In this periodic case, the
2274: transversal $X=\cL/\Gamma$ is a finite set of points. The groupoid
2275: $C^*$-algebra $C^*(\cG(\cL,X))$ is in this case isomorphic to
2276: $C(\hat\Gamma)\otimes M_k(\C)$, where $k$ is the cardinality of the
2277: transversal $X$. Thus, the periodic case falls back into the realm of
2278: commutative spaces, while the aperiodic patterns give rise to truly
2279: noncommutative spaces, which are highly nontrivial and interesting.
2280:
2281: \begin{center}
2282: \begin{figure}
2283: \includegraphics[scale=0.6]{muqarnas.eps}
2284: \caption{Quasiperiodic tilings and muqarnas.
2285: \label{FigMuq}}
2286: \end{figure}
2287: \end{center}
2288:
2289: One of the richest sources of interesting tilings are the
2290: {\em zellijs} and {\em muqarnas} widely used in ancient architecture.
2291: Also collectively defined as ``arabesques'', not only these patterns
2292: exhibit highly nontrivial geometries, but they reflect
2293: the intricate interplay between philosophy, mathematics,
2294: and aesthetics (\cf \cite{ArdBak}, \cite{Bulatov}).
2295: Some of the best studies on
2296: zellijs and muqarnas concentrate on 2-dimensional {\em periodic}
2297: patterns. For instance we find in \cite{ArdBak}, p.43:
2298: \begin{quote}
2299: ``As Nature is based on rhythm, so the arabesque is rhythmic
2300: in concept. It reflects movement marked by the regular recurrence
2301: of features, elements, phenomena; hence it has periodicity.''
2302: \end{quote}
2303: It seems from this viewpoint that only the theory of periodic
2304: tilings (\ie commutative geometry) should be relevant in this
2305: context. However, more recent
2306: studies (\cf \cite{Bulatov}, \cite{Castera}, \cite{Castera2},
2307: \cite{Nagy}) suggest that the design of zellijs and muqarnas was not
2308: limited to 2-dimensional crystallographic groups, but, especially
2309: during the Timurid period, it
2310: involved also {\em aperiodic} patterns with fivefold symmetry, analogous to
2311: those observed in quasi-crystals. This is no accident and was
2312: certainly the result of a highly developed geometric theory:
2313: already in the historic textbook of Abu'l-Wafa' al-Buzjani (940-998)
2314: on geometric constructions \cite{Wafa} there is explicit mention of
2315: meetings and discussions where mathematicians were directly involved
2316: alongside artisans in the design of arabesque patterns.
2317:
2318: The appearance of aperiodic
2319: tilings is documented in the anonymous Persian manuscript
2320: \cite{Anonym} ``On interlocking similar and congruent figures'', which
2321: dates back to the 11th-13th century. Some of these aperiodic aspects of
2322: zellijs and muqarnas were studied by Bulatov in the
2323: book \cite{Bulatov}, which also contains Vil'danova's Russian
2324: translation of the ancient Persian text.
2325:
2326:
2327: \bigskip
2328:
2329: \section{Noncommutative spaces from dynamical systems}\label{dynsys}
2330:
2331:
2332:
2333: We will look at some examples of noncommutative spaces associated to
2334: a discrete dynamical system $T$, for instance given by a self
2335: mapping of a Cantor set. Such noncommutative spaces
2336: have been extensively studied in a
2337: series of papers (\cf \cite{GPS} and \cite{skau} for a survey) where
2338: C. Skau and his coworkers have obtained remarkable results on the
2339: classification of minimal actions of $\Z$ on Cantor
2340: sets using the $K$-theory of the associated $C^*$-algebra.
2341:
2342: It was found recently (\cf \cite{CM1}, \cite{CM2}, \S 4 of
2343: \cite{Mar} and \S 8 of \cite{MaPa}) that the mapping torus of such
2344: systems can be used to model the ``dual graph'' of the fibers at the
2345: archimedean primes of arithmetic surfaces, in Arakelov geometry, in
2346: the particular case in which the dynamical system $T$ is a subshift
2347: of finite type encoding the action of a Schottky group
2348: $\Gamma\subset \SL_2(\C)$ on its limit set $\Lambda_\Gamma\subset
2349: \P^1(\C)$. In fact, the results of \cite{CM1} were motivated by
2350: earlier results of Manin \cite{Man-hyp} that provided a geometric
2351: model for such dual graphs in terms of hyperbolic geometry and
2352: Schottky uniformizations.
2353:
2354:
2355: \smallskip
2356:
2357: More generally, given an alphabet with letters $\{ \ell_1, \ldots, \ell_N \}$,
2358: the space $\cS^+_A$ of a subshift of finite type consists
2359: of all right-infinite {\em admissible} sequences
2360: \begin{equation}\label{admseqplus}
2361: w=a_0 a_1 a_2 \ldots a_n \ldots
2362: \end{equation}
2363: in the letters of the alphabet. Namely, $a_i\in \{ \ell_1, \ldots,
2364: \ell_N \}$ subject to an admissibility condition specified by an
2365: $N\times N$ matrix $A$ with entries in $\{0,1\}$. Two letters
2366: $\ell_i$ and $\ell_j$ in the alphabet can appear as consecutive digits
2367: $a_k$, $a_{k+1}$ in the word $w$ if and only if the entry $A_{ij}$ of
2368: the admissibility matrix $A$ is equal to $1$.
2369: One defines similarly the space $\cS_A$ as the set of doubly-infinite
2370: admissible sequences
2371: \begin{equation}\label{admseq}
2372: w=\ldots a_{-m}\ldots a_{-2} a_{-1} a_0 a_1 a_2 \ldots a_n \ldots
2373: \end{equation}
2374: The sets $\cS^+_A$ and $\cS_A$ have a natural choice of a topology.
2375: In fact, on $\cS_A$ we can put the topology generated by the
2376: sets $W^s(x,\ell) =\{ y\in {\mathcal S}_A | x_k = y_k, k\geq
2377: \ell \}$, and the $W^u(x,\ell)=\{ y\in {\mathcal S}_A | x_k = y_k, k\leq
2378: \ell \}$ for $x\in {\mathcal S}_A$ and $\ell \in \Z$. This induces a
2379: topology with analogous properties on ${\mathcal S}^+_A$ by realizing it as a
2380: subset of ${\mathcal S}_A$, for instance, by extending each sequence to
2381: the left as a constant sequence. One then considers on $\cS_A$ (or on
2382: $\cS_A^+$) the action of the two-sided (resp.~one-sided) shift
2383: $T$ defined by $(Tw)_k = a_{k+1}$, where the $a_k$ are the digits of
2384: the word $w$. Namely, the one-sided shift on $\cS_A^+$ is of the form
2385: \begin{equation}\label{shift+}
2386: T( a_0 a_1 a_2 \ldots a_\ell \ldots) = a_1 a_2 \ldots
2387: a_\ell \ldots
2388: \end{equation}
2389: while the two-sided shift on $\cS_A$ acts as
2390: \begin{equation}\label{shift}
2391: \begin{array}{rcccccccccl} T(& \ldots & a_{-m} & \ldots & a_{-1}& a_0
2392: & a_1 &
2393: \ldots & a_{\ell} & \ldots &) = \\ & \ldots & a_{-m+1} & \ldots & a_{0} &
2394: a_1 & a_2 & \ldots & a_{\ell+1} & \ldots & \end{array}
2395: \end{equation}
2396: Tipically spaces ${\mathcal S}^+_A$ and ${\mathcal S}_A$ are topologically
2397: Cantor sets. The one-sided shift $T$ of \eqref{shift+} is a
2398: continuous surjective map on ${\mathcal S}^+_A$, while the two-sided
2399: shift $T$ of \eqref{shift} is a homeomorphism of ${\mathcal S}_A$.
2400:
2401: For example, let $\Gamma$ be a free group in $g$ generators $\{
2402: \gamma_1,\ldots, \gamma_g \}$. Consider the alphabet
2403: $\{ \gamma_1,\ldots, \gamma_g, \gamma_1^{-1}, \ldots, \gamma_g^{-1}
2404: \}$. Then one can consider the right-infinite, or doubly-infinite
2405: words in these letters, without cancellations, that is, subject to the
2406: admissibility rule that $a_{k+1}\neq a_k^{-1}$. This defines a
2407: subshift of finite type where the matrix $A$ is the symmetric
2408: $2g\times 2g$ matrix with $A_{ij}=0$ for $|i-j|=g$ and $A_{ij}=1$
2409: otherwise. Suppose that $\Gamma$ is
2410: a Schottky group of genus $g$, \ie a finitely generated discrete
2411: subgroup $\Gamma\subset \SL_2(\C)$, isomorphic to a free group in $g$
2412: generators, where all nontrivial elements are hyperbolic.
2413: Then the points in $\cS_A^+$ parameterize points in the
2414: limit set $\Lambda_\Gamma\subset \P^1(\C)$ (the set of accumulation
2415: points of orbits of $\Gamma$).
2416: The points in $\cS_A$ parameterize geodesics in the
2417: three dimensional real hyperbolic space $\H^3$ with ends at points on
2418: the limit set $\Lambda_\Gamma$.
2419:
2420: The pair $({\mathcal S}_A,T)$ is a typical example of an interesting
2421: class of dynamical systems, namely it is a {\em Smale
2422: space}. This means that locally ${\mathcal S}_A$ can be decomposed
2423: as the product of expanding and contracting directions for $T$.
2424: Namely, the following
2425: properties are satisfied.
2426: \begin{itemize}
2427: \item For every point $x\in {\mathcal S}_A$ there exist subsets $W^s(x)$
2428: and $W^u(x)$ of ${\mathcal S}_A$, such that $W^s(x) \times W^u(x)$ is
2429: homeomorphic to a neighborhood of $x$.
2430: \item The map $T$ is contracting on $W^s(x)$ and expanding on
2431: $W^u(x)$, and $W^s(Tx)$ and $T(W^s(x))$ agree in some neighborhood of
2432: $x$, and so do $W^u(Tx)$ and $T(W^u(x))$.
2433: \end{itemize}
2434:
2435: A construction of Ruelle shows that one can
2436: associate different $C^*$--algebras to Smale spaces (\cf \cite{Rue2},
2437: \cite{Putn}, \cite{PutSpi}). For Smale spaces like $({\mathcal
2438: S}_A,T)$ there are four basic possibilities: the crossed product algebra
2439: $C({\mathcal S}_A)\rtimes_T \Z$ and the ${\rm C}^*$--algebras ${\rm
2440: C}^*({\mathcal G}^s)\rtimes_T \Z$, ${\rm C}^*({\mathcal G}^u)\rtimes_T
2441: \Z$, ${\rm C}^*({\mathcal G}^a)\rtimes_T \Z$ obtained by considering
2442: the action of the shift $T$ on the groupoid ${\rm C}^*$--algebra
2443: associated to the groupoids ${\mathcal G}^s$, ${\mathcal G}^u$,
2444: ${\mathcal G}^a$ of the stable, unstable, and asymptotic equivalence
2445: relations on $({\mathcal S}_A,T)$.
2446:
2447: The first choice, $C({\mathcal S}_A)\rtimes_T \Z$, is
2448: closely related to the continuous dynamical system given by the
2449: mapping torus of $T$, while a choice like ${\rm C}^*({\mathcal
2450: G}^u)\rtimes_T \Z$ is related to the ``bad quotient'' of $\cS_A^+$ by
2451: the action of $T$. In the example of the Schottky group this
2452: corresponds to the action of $\Gamma$ on its limit set.
2453:
2454: One can consider the suspension flow $\cS_T$ of a dynamical
2455: system $T$, that is, the mapping torus of the dynamical
2456: system $( {\mathcal S}_A, T)$, which is defined as
2457: \begin{equation} \label{suspensionT}
2458: {\mathcal S}_T := {\mathcal S}_A
2459: \times [0,1] / (x,0)\sim (Tx,1).
2460: \end{equation}
2461:
2462: \begin{figure}
2463: \begin{center}
2464: \includegraphics[scale=0.65]{MapTorus.eps}
2465: \end{center}
2466: \caption{Mapping Torus \label{FigMapTorus}}
2467: \end{figure}
2468:
2469: The first cohomology group of $\cS_T$ is
2470: the ``ordered cohomology'' of the dynamical system $T$, in the sense of
2471: \cite{BoHa} \cite{PaTu}.
2472: There is an identification of $H^1({\mathcal S}_T,\Z)$ with the
2473: $K_0$-group of the crossed product ${\rm C}^*$-algebra for the
2474: action of $T$ on ${\mathcal S}_A$,
2475: \begin{equation}
2476: H^1({\mathcal S}_T,\Z) \cong K_0({\rm C}({\mathcal
2477: S}_A) \rtimes_T \Z). \label{H1K0}
2478: \end{equation}
2479: This can be seen from the Pimsner--Voiculescu exact sequence (\cf
2480: \cite{PV}) for the $K$-theory of a crossed product by $\Z$, which in
2481: this case reduces to
2482: \begin{equation}\label{PV} 0 \to K_1({\rm C}({\mathcal S})\rtimes_T
2483: \Z) \to {\rm C}({\mathcal S},\Z) \stackrel{I-T_*}{\to} {\rm
2484: C}({\mathcal S},\Z) \to K_0({\rm C}({\mathcal S})\rtimes_T \Z) \to
2485: 0, \end{equation}
2486: It can also be seen in terms of the Thom isomorphism
2487: of \cite{Co-Thom}, \cite{Co1}.
2488:
2489: In fact, as we discussed in Section \ref{road}, one of the
2490: fundamental construction of noncommutative geometry (\cf
2491: \cite{Co1}) is that of {\em homotopy
2492: quotients}. These are commutative spaces, which provide, up to
2493: homotopy, geometric models for the corresponding noncommutative
2494: spaces. The noncommutative spaces themselves, as we are going to show
2495: in our case, appear as quotient spaces of foliations on the homotopy
2496: quotients with contractible leaves.
2497:
2498: For the noncommutative space ${\mathcal S}_A/\Z$, with $\Z$ acting as
2499: powers of the invertible two-sided shift, the homotopy quotient
2500: is precisely the mapping torus \eqref{suspensionT},
2501: \begin{equation}\label{htpyquot2}
2502: {\mathcal S}_T = {\mathcal S}\times_\Z \R.
2503: \end{equation}
2504: The noncommutative space ${\mathcal S}/\Z$ can be
2505: identified with the quotient space of the natural foliation on
2506: \eqref{htpyquot2} whose generic leaf is contractible (a copy of
2507: $\R$).
2508:
2509: Another noncommutative space associated to a subshift
2510: of finite type $T$ (which, up to Morita equivalence, correspond to
2511: another choice of the $C^*$-algebra of a Smale space, as mentioned
2512: above) is the Cuntz--Krieger algebra $\cO_A$,
2513: where $A$ is the admissibility matrix of the subshift finite type
2514: (\cf \cite{Cu} \cite{CuKrie}).
2515:
2516: A partial isometry is a linear operator $S$ satisfying the relation
2517: $S= S S^* S$. The Cuntz--Krieger algebra ${\mathcal O}_A$ is defined
2518: as the universal $C^*$--algebra generated by partial isometries
2519: $S_1, \ldots, S_{N}$, satisfying the relations
2520: \begin{equation} \label{CK1rel} \sum_j S_j S_j^* =I \end{equation}
2521: \begin{equation} \label{CK2rel} S_i^* S_i =\sum_j A_{ij} \, S_j
2522: S_j^*. \end{equation}
2523:
2524:
2525: In the case of a Schottky group $\Gamma\subset \PSL_2(\C)$
2526: of genus $g$, the Cuntz--Krieger
2527: algebra ${\mathcal O}_A$ can be described in terms
2528: of the action of the free group $\Gamma$ on its limit set
2529: $\Lambda_\Gamma\subset \P^1(\C)$ (\cf \cite{Rob}, \cite{Spi}), so that we can regard
2530: ${\mathcal O}_A$ as a noncommutative space replacing the classical
2531: quotient $\Lambda_\Gamma / \Gamma$,
2532: \begin{equation} \label{treefree}
2533: {\mathcal O}_A \cong
2534: C(\Lambda_\Gamma) \rtimes \Gamma.
2535: \end{equation}
2536: The quotient space
2537: \begin{equation}\label{htpyquot1}
2538: \Lambda_\Gamma \times_\Gamma \H^3 = \Lambda_\Gamma \times_\Gamma
2539: \underline{E}\Gamma,
2540: \end{equation}
2541: is precisely the homotopy quotient of $\Lambda_\Gamma$ with respect to
2542: the action of $\Gamma$, with $\underline{E}\Gamma =\H^3$ and the
2543: classifying space $\underline{B}\Gamma = \H^3/\Gamma$. Here
2544: $\H^3/\Gamma$ is a hyperbolic 3-manifold of infinite volume, which
2545: is topologically a handlebody of genus $g$.
2546: In this case also we find
2547: that the noncommutative space $\Lambda_\Gamma /\Gamma$ is the quotient
2548: space of a foliation on the homotopy quotient \eqref{htpyquot1} with
2549: contractible leaves $\H^3$.
2550:
2551: \bigskip
2552:
2553:
2554: \section{Noncommutative spaces from string theory}\label{string}
2555:
2556: The main aspects of string and D-brane theory that
2557: involve noncommutative geometry are the bound states of
2558: configurations of parallel D-branes \cite{Wit1}, the matrix models
2559: for M-theory \cite{BFSS} and the strong coupling limit of
2560: string theory (\cf \eg \cite{Ard1}, \cite{Ard2}). It also plays an
2561: important role in the M-theory compactifications \cite{CDS}.
2562: We shall not discuss all these aspects in detail here. Since the
2563: focus of this review is on examples we only mention a
2564: couple of examples of noncommutative spaces arising from
2565: string and D-brane theory.
2566:
2567: \medskip
2568:
2569: The noncommutative tori and the
2570: components of the Yang-Mills connections appear in the
2571: classification of the BPS states in M-theory \cite{CDS}.
2572:
2573: Recall first that Yang--Mills theory on noncommutative tori
2574: can be formulated (\cf \cite{CoRi}) using suitable notions of
2575: connections and curvature for noncommutative spaces. In fact,
2576: the analogs of connection
2577: and curvature of vector bundles are straightforward to obtain
2578: (\cite{C3}): a connection is just given by the associated
2579: covariant differentiation $\nabla$ on the space of smooth
2580: sections. Thus here it is given by a pair of linear operators,
2581: \begin{equation}
2582: \nabla_j : \cS (\R) \rightarrow \cS (\R)
2583: \label{nabla}
2584: \end{equation}
2585: such that
2586: \begin{equation}
2587: \nabla_j (\xi b) = (\nabla_j \xi)b + \xi \delta_j (b) \quad
2588: \forall \xi \in \cS \ , b \in \cA_\theta \, .
2589: \label{connection}
2590: \end{equation}
2591: One checks that, as in the usual case, the trace of the curvature
2592: $$\Omega = \nabla_1 \nabla_2 - \nabla_2 \nabla_1, $$
2593: is independent of the choice of the connection.
2594:
2595: We can make the following choice for the connection:
2596: \begin{equation}
2597: (\nabla_1 \xi) (s) = - \frac{2\pi i s}{\theta} \, \xi (s) \
2598: \,\,\,\,\,\, (\nabla_2 \xi)(s) = \xi' (s) \, . \label{connection2}
2599: \end{equation}
2600: Notice that, up to the correct powers of
2601: $2\pi i$, the total curvature of $\cS$ is an {\em integer}.
2602: In fact, the curvature $\Omega$ is constant, equal to $\frac{1}{\theta}$,
2603: so that the irrational number $\theta$ disappears in the total
2604: curvature, $\theta \times \frac{1}{\theta}$.
2605: This integrality phenomenon, where the pairing of dimension and
2606: curvature (both of which are non--integral) yields an integer:
2607: $$ \dim \times \Omega \sim \theta \times \frac{1}{\theta} =
2608: \text{integer}, $$ is the basis for the development of a theory of
2609: characteristic classes for non--commutative spaces. In the general
2610: case, this requires the development of more sophisticated tools,
2611: since analogs of the derivations $\delta_i$ used in the
2612: case of the noncommutative tori are not there in general.
2613: The general theory is obtained through cyclic
2614: homology, as developed in \cite{ConnesCH}.
2615:
2616: \smallskip
2617:
2618: Consider then the modules ${\mathcal H}^\theta_{p,q}$ described in
2619: Section \ref{nctori}. It is possible to define an $\cA_\theta$ valued
2620: inner product $\langle \cdot, \cdot \rangle_{\cA}$ on ${\mathcal
2621: H}^\theta_{p,q}$, as in \cite{Rieffel}, which is used to show that
2622: ${\mathcal H}^\theta_{p,q}$ is a projective module.
2623: Connections are required to be compatible with the metric,
2624: \begin{equation}\label{metr-conn}
2625: \delta_j \langle\xi,\eta \rangle_{\cA} =\langle \nabla_j \xi,\eta
2626: \rangle_{\cA} +
2627: \langle \xi, \nabla_j \eta \rangle_{\cA}.
2628: \end{equation}
2629: It is proved in \cite{C3} that such connections always exist. The
2630: curvature $\Omega$ has values in $\cE=\End_{\cA}({\mathcal H})$.
2631: An $\cE$--valued inner product on ${\mathcal H}$ is given by
2632: $$ \langle \xi,\eta \rangle_{\cE} \zeta = \xi \langle \eta, \zeta
2633: \rangle_{\cA}, $$ and a canonical faithful trace $\tau_E$ is
2634: defined as
2635: $$ \tau_E(\langle \xi,\eta \rangle_{\cE}) = \tau(\langle \eta,\xi
2636: \rangle_{\cA}), $$ where $\tau$ is the trace on the algebra
2637: $\cA_\theta$, given by \eqref{trace-theta}.
2638:
2639: The Yang--Mills action is defined (\cf \cite{CoRi}) as
2640: \begin{equation} \label{YM-action}
2641: \tau(\langle \Omega, \Omega\rangle_{\cE}).
2642: \end{equation}
2643: One seeks for minima of the Yang--Mills action among metric
2644: compatible connections \eqref{metr-conn}. The main result is
2645: that this recovers the classical moduli spaces
2646: of Yang--Mills connections on the ordinary torus (\cite{CoRi}):
2647:
2648: \begin{thm} \label{NCT-YM}
2649: For a choice of a pair of integers $(p,q)$ with $p + q\theta \geq
2650: 0$, the moduli space of Yang--Mills connections on the
2651: $\cA_\theta$ module ${\mathcal H}^\theta_{pq}$ is a classical
2652: space given by the symmetric product
2653: $$ s^N(T^2)= (T^2)^N / \Sigma_N, $$
2654: where $\Sigma_N$ is the group of permutations in $N$--elements,
2655: for $N=\gcd (p,q)$.
2656: \end{thm}
2657:
2658: \smallskip
2659:
2660: In the matrix formulation of M-theory the basic equations to
2661: obtain periodicity of two of the basic coordinates $X_i$ turn out
2662: to be the
2663: \begin{equation}
2664: U_i X_j U_i^{-1} =X_j + a \delta_i^j , \ \ \ i= 1,2, \label{Mtheory}
2665: \end{equation}
2666: where the $U_i$ are unitary gauge transformations.
2667: The multiplicative commutator $U_1 U_2 U_1^{-1}U_2^{-1}$ is then
2668: central and in the irreducible case its scalar value $ \lambda
2669: =\exp 2\pi i \theta $ brings in the algebra of coordinates on the
2670: noncommutative torus. The $X_j$ are then the components of the
2671: Yang-Mills connections. The same
2672: picture emerged from the other information one has about M-theory
2673: concerning its relation with 11 dimensional supergravity and that
2674: string theory dualities can then be interpreted using Morita
2675: equivalence, relating the values of $\theta$ on an orbit of $SL_2
2676: (\Z)$.
2677:
2678: \smallskip
2679:
2680: Nekrasov and Schwarz \cite{NS} showed that Yang-Mills gauge
2681: theory on noncommutative $\R^4$ gives a conceptual understanding of the
2682: nonzero B-field desingularization of the moduli space of
2683: instantons obtained by perturbing the ADHM equations.
2684: In \cite{Witten}, Seiberg and Witten exhibited the unexpected
2685: relation between the standard gauge theory and the noncommutative
2686: one, and clarified the limit in which the entire string dynamics
2687: is described by a gauge theory on a noncommutative space.
2688: Techniques from noncommutative differential and Riemannian geometry,
2689: in the sense discussed in Section \ref{road} were applied to
2690: string theory, for instance in \cite{Ard1}. The role of noncommutative
2691: geometry in the context of $T$-duality was considered in very
2692: interesting recent work of Mathai and collaborators, \cite{BMa1},
2693: \cite{BMa2}, \cite{RoMa}.
2694:
2695: \medskip
2696:
2697: Recently, in the context of the holographic description of
2698: type IIB string theory on the plane-wave background,
2699: Shahin M.M.~Sheikh-Jabbari obtained (\cf \cite{Shahin}) an interesting
2700: class of noncommutative spaces from the quantization of Nambu $d$-brackets.
2701: The classical Nambu brackets
2702: \begin{equation}\label{NambuC}
2703: \{ f_1,\ldots, f_k \}= \epsilon^{i_1\cdots i_k} \frac{\partial
2704: f_1}{\partial x^{i_1}}\cdots \frac{\partial
2705: f_k}{\partial x^{i_k}}
2706: \end{equation}
2707: of $k$ real valued functions of variables $(x^1,\ldots, x^k)$
2708: is quantized in the even case to the expression in $2k$
2709: operators
2710: \begin{equation}\label{NambuQev}
2711: \frac{1}{i^k} [ F_1,\ldots, F_{2k}]=\frac{1}{i^k (2k)!}
2712: \epsilon^{i_1\cdots i_{2k}} F_{i_1}\cdots F_{i_{2k}}.
2713: \end{equation}
2714: This generalizes the Poisson bracket quantization
2715: $\{f_1,f_2\} \mapsto \frac{-i}{\hbar} [F_1,F_2]$.
2716: The odd case is more subtle and it involves an additional operator
2717: related to the chirality operator $\gamma_5$. One sets
2718: \begin{equation}\label{NambuQodd}
2719: \frac{1}{i^k} [ F_1,\ldots, F_{2k-1}, \gamma]=\frac{1}{i^k (2k)!}
2720: \epsilon^{i_1\cdots i_{2k}} F_{i_1}\cdots F_{i_{2k-1}} \gamma,
2721: \end{equation}
2722: where $\gamma$ is the chirality operator in $2k$-dimensions.
2723: For example, for $k=2$ one gets
2724: $$ \begin{array}{rl}
2725: [F_1,F_1,F_3,\gamma]=& \frac{1}{24}([F_1,F_2][F_3,\gamma]
2726: +[F_3,\gamma][F_1,F_2] \\[2mm] & -([F_1,F_3][F_2,\gamma]+[F_2,\gamma][F_1,F_3])
2727: \\[2mm] & +[F_2,F_3][F_1,\gamma] +[F_1,\gamma][F_2,F_3]).
2728: \end{array} $$
2729: If one describes the ordinary $d$-dimensional sphere of radius $R$
2730: by the equation
2731: \begin{equation}\label{Sd1}
2732: \sum_{i=1}^{d+1} (x^i)^2=R^2,
2733: \end{equation}
2734: the coordinates satisfy
2735: \begin{equation}\label{Sd2}
2736: \{ x^{i_1},\ldots,x^{i_d}\}=R^{d-1} \epsilon^{i_1\cdots i_{d+1}}
2737: x^{i_{d+1}}.
2738: \end{equation}
2739: The equation \eqref{Sd1} and \eqref{Sd2} are then replaced by their
2740: quantized version, using the quantization of the Nambu bracket and
2741: the introduction of a quantization parameter $\ell$. This defines
2742: algebras generated by unitaries $X^i$ subject to the relations given
2743: by the quantization of \eqref{Sd1} and \eqref{Sd2}. Matrix
2744: representations of these algebra correspond to certain fuzzy spheres.
2745: It would be interesting to study the general structure of these
2746: noncommutative spaces from the point of view of the various steps
2747: introduced in Section \ref{road}, \cf also the discussion in Section
2748: \ref{deform}.
2749:
2750:
2751: \bigskip
2752:
2753:
2754: \section{Groupoids and the index theorem }
2755: \label{cotangent}
2756:
2757: Since the construction of the $C^*$-algebra of foliations based on
2758: the holonomy groupoid (section \ref{foliations}), groupoids have
2759: played a major role in noncommutative geometry. In fact the original
2760: construction of matrix mechanics by Heisenberg (section
2761: \ref{heisenberg}) is exactly that of the convolution algebra of the
2762: groupoid of transitions imposed by experimental results. The
2763: convolution algebra of groupoids can be defined in the context of
2764: von-Neumann algebras and of $C^*$-algebras (\cf \cite{Co78} and
2765: \cite{Ren}). It is particularly simple and canonical in the context
2766: of smooth groupoids (\cf \cite{Co94} section II.5). One virtue of
2767: the general construction is that it provides a geometric mental
2768: picture of complicated analytical constructions. The prototype
2769: example is given by the tangent groupoid of a manifold (\cf
2770: \cite{Co94} section II.5). It is obtained by blowing up the diagonal
2771: in the square $V\times V$ of the manifold and as a set is given by
2772: $$
2773: G_V=\,V\times V\times ]0,1]\cup TV
2774: $$
2775: where $TV$ is the (total space of the) tangent bundle of $V$. A
2776: tangent vector $X\in T_x(V)$ appears as the limit of nearby triples
2777: $(x_1, x_2, \varepsilon)$ provided in any chart the ratios
2778: $(x_1-x_2)/\varepsilon$ converge to $X$. When $\varepsilon \to 0$
2779: the Heisenberg (matrix) law of composition :
2780: $$
2781: (x_1, x_2, \varepsilon)\circ (x_2, x_3, \varepsilon)=(x_1, x_3,
2782: \varepsilon)
2783: $$
2784: converges to the addition of tangent vectors, so that $G_V$ becomes
2785: a smooth groupoid. The functoriality of the construction $G\to
2786: C^*(G)$ of the convolution algebra for smooth groupoids $G$ is then
2787: enough to define the Atiyah-Singer analytic index of
2788: pseudo-differential operators. It is simply given by the map in
2789: K-theory for the exact sequence of $C^*$-algebras associated to the
2790: geometric sequence
2791: $$
2792: V\times V\times ]0,1] \to G_V \supset TV
2793: $$
2794: where $TV$ is viewed as a closed subgroupoid of $G_V$. The
2795: corresponding exact sequence of $C^*$-algebras can be written as
2796: $$
2797: 0\to C_0(]0,1])\times \mathcal K\to C^*(G_V)\to C_0(T^*V)\to 0
2798: $$
2799: and is a geometric form of the extension of pseudo-differential
2800: operators. By construction the algebra $C_0(]0,1])$ is {\em
2801: contractible} and the same holds for the tensor product
2802: $C_0(]0,1])\times \mathcal K$ by the algebra $\mathcal K$ of compact
2803: operators. This shows that the restriction map $C^*(G_V)\to
2804: C_0(T^*V)$ is an isomorphism in K-theory :
2805: \begin{equation}\label{tstariso}
2806: K_0(C_0(T^*V))\sim K_0(C^*(G_V))
2807: \end{equation}
2808: Since the K-theory of
2809: $\mathcal K$ is $\Z$ for $K_0$, one gets the analytic index by the
2810: evaluation map
2811: $$
2812: C^*(G_V)\to \mathcal K\,,\quad K_0(C^*(G_V))\to K_0(\mathcal K)=\Z
2813: $$
2814: composed with the isomorphism \eqref{tstariso}. Using the Thom
2815: isomorphism yields a geometric proof (\cf \cite{Co94}) of the
2816: Atiyah-Singer index theorem, where all the analysis has been taken
2817: care of once and for all by the functor $G\to C^*(G)$.
2818:
2819: This paradigm for a geometric set-up of the index theorem has been
2820: successfully extended to manifolds with singularities (\cf
2821: \cite{Month} \cite{Month1} and references there) and to manifolds
2822: with boundary \cite{ANS}.
2823:
2824: \bigskip
2825:
2826:
2827:
2828:
2829: \section{Riemannian manifolds, conical singularities}\label{singconic}
2830:
2831: A main property of the homotopy type of a compact
2832: oriented manifold is that it satisfies Poincar\'e duality not
2833: just in ordinary homology but also in $K$-homology.
2834: In fact, while
2835: Poincar\'e duality in ordinary homology is not sufficient to describe
2836: homotopy type of manifolds (\cf \cite{Mi-S}), Sullivan proved (\cf
2837: \cite{Sull}) that for simply connected PL
2838: manifolds of dimension at least $5$, ignoring 2-torsion,
2839: the same property in $KO$-homology suffices and the
2840: Chern character of the $KO$-homology
2841: fundamental class carries all the rational information on the
2842: Pontrjagin classes.
2843:
2844: For an ordinary manifold the choice of the fundamental cycle in
2845: $K$-homology is a refinement of the choice of orientation of the
2846: manifold and, in its simplest form, it is a choice of
2847: Spin-structure. Of course the role of a spin structure is to allow
2848: for the construction of the corresponding Dirac operator which
2849: gives a corresponding Fredholm representation of the algebra of
2850: smooth functions. The choice of a square root involved in the
2851: Dirac operator $D$ corresponds to a choice of $K$-orientation.
2852:
2853: $K$-homology admits a
2854: fairly simple definition in terms of Hilbert spaces and Fredholm
2855: representations of algebras. In fact,
2856: we have the following notion of Fredholm module (\cf \cite{Co94}):
2857:
2858: \begin{defn}\label{Fred-mod-act}
2859: Let $\cA$ be an algebra, an {\em odd} Fredholm module over $\cA$
2860: is given by:
2861: \begin{enumerate}
2862: \item a representation of $\cA$ in a Hilbert space $\cH$.
2863: \item an operator $F = F^*$, $F^2 = 1$, on $\cH$ such that
2864: $$
2865: [F , a] \ \hbox{is a compact operator for any} \ a \in \cA \, .
2866: $$
2867: \end{enumerate}
2868: An {\it even} Fredholm module is given by an odd Fredholm module
2869: $(\cH , F)$ together with a $\Z / 2$ grading
2870: $\gamma$, $\gamma = \gamma^*$,
2871: $\gamma^2 = 1$ of the Hilbert space $\cH$ satisfying:
2872: \begin{enumerate}
2873: \item $\gamma a = a \gamma$, for all $a \in \cA$
2874: \item $\gamma F = -F \gamma$.
2875: \end{enumerate}
2876: \end{defn}
2877:
2878: This definition is derived from
2879: Atiyah's definition \cite{AT} of abstract elliptic operators,
2880: and agrees with Kasparov's definition \cite{Kasp} for the cycles
2881: in $K$-homology, $KK(A,\C)$, when $A$ is a $C^*$-algebra.
2882:
2883: The notion of Fredholm module can be illustrated by the following
2884: examples (\cf \cite{Co94}).
2885:
2886: \begin{ex}\label{Fred-manifold} {\em If $X$ is a manifold, an
2887: elliptic operator on $X$ can be twisted with vector bundles, so as
2888: to give rise to an index map ${\rm Ind}: K_0(C(X)) \to \Z$. If $P$
2889: is an elliptic operator (the symbol is invertible) and a
2890: pseudodifferential operator of order zero, $P: L^2(X,E_+)\to
2891: L^2(X,E_-)$, then there exists a parametrix $Q$ for $P$. This is
2892: also an operator of order zero, and a quasi-inverse for $P$, in
2893: the sense that it is an inverse at the symbol level, namely $PQ-I$
2894: and $QP-I$ are compact operators. Consider then the operator
2895: $$ F=\left( \begin{array}{cc} 0 & Q \\ P & 0 \end{array} \right) $$
2896: on $\cH=L^2(X,E_+)\oplus L^2(X,E_-)$. The algebra $C(X)$ acts on
2897: $\cH$ and $[F,f]$ is a compact operator for all $f\in C(X)$. Since
2898: $F^2-I$ is compact, it is possible to add to $\cH$ a finite
2899: dimensional space to obtain $F^2=I$. Notice that the functions of
2900: $C(X)$ act differently on this modified space. In particular the
2901: function $f\equiv 1$ no longer acts as the identity: one recovers
2902: the index of $P$ this way.}
2903: \end{ex}
2904:
2905: \begin{ex}\label{Fred-freeG} {\em Let $\Gamma=\Z * \Z$ a free
2906: group, and let $\cA =\C \Gamma$. Let $\cT$ be the tree of $\Gamma$
2907: with $\cT^0$ the set of vertices and $\cT^1$ the set of edges. Fix
2908: an origin $x_0$ in $\cT^0$. For any vertex $v\in \cT^0$ there
2909: exists a unique path connecting it to the origin $x_0$. This
2910: defines a bijection $\phi: \cT^0 \backslash \{ x_0 \} \to \cT^1$
2911: that assigns $v\mapsto \phi(v)$ with $\phi(v)$ this unique edge.
2912: Let $U_\phi$ be the unitary operator implementing $\phi$, and
2913: consider the operator
2914: $$ F = \left( \begin{array}{cc} 0 & U_\phi \\ U_\phi^* & 0
2915: \end{array} \right) $$
2916: acting on $\cH = \ell^2(\cT^0) \oplus \ell^2(\cT^1) \oplus \C$. By
2917: construction $\Gamma$ acts naturally on $\cT_j$ which gives a
2918: corresponding action of $\cA$ in $\cH$. The pair $(\cH,F)$ is a
2919: Fredholm module over $\cA$.}
2920: \end{ex}
2921:
2922:
2923: \begin{ex}\label{Hilb-transform} {\em On $S^1\simeq
2924: \bP^1(\R)$, consider the algebra of functions ${\rm
2925: C}(\bP^1(\R))$, acting on the Hilbert space $L^2(\R)$, as
2926: multiplication operators $(f\, \xi)(s) = f(s) \xi(s)$. Let $F$ be
2927: the Hilbert transform
2928: $$ (F\, \xi)(s)=\frac{1}{\pi i} \int \frac{f(t)}{s-t} \, dt. $$
2929: This multiplies by $+1$ the positive Fourier modes and by $-1$ the
2930: negative Fourier modes. A function $f\in {\rm C}(\bP^1(\R))$ has
2931: the property that $[F,f]$ is of finite rank if and only if $f$ is
2932: a rational function $f(s) =P(s)/Q(s)$. This is Kronecker's
2933: characterization of rational functions.}
2934: \end{ex}
2935:
2936: \medskip
2937:
2938: Besides the $K$-homology class, specified by a Fredholm module,
2939: one also wants to generalize to the noncommutative setting the
2940: infinitesimal line element $ds$ of a Riemannian manifold. In ordinary
2941: Riemannian geometry one deals rather with the $ds^2$ given by
2942: the usual local expression $g_{\mu \nu} \, dx^{\mu} \, dx^{\nu}$.
2943: However, in order to extend the notion of metric space to the
2944: noncommutative setting it is more natural to deal with $ds$, for
2945: which the ansatz is
2946: \begin{equation}
2947: ds = \times\!\!\!\!-\!\!\!-\!\!\!-\!\!\!\!\!\times \, ,
2948: \label{dsx-x}
2949: \end{equation}
2950: where the right hand side has the meaning usually attributed to it in
2951: physics, namely the Fermion propagator
2952: \begin{equation}
2953: \times\!\!\!\!-\!\!\!-\!\!\!-\!\!\!\!\!\times = D^{-1} ,
2954: \label{Fermprop}
2955: \end{equation}
2956: where $D$ is the Dirac operator. In other words, the presence of a
2957: spin (or spin$^c$) structure makes it possible to extract the square
2958: root of $ds^2$, using the Dirac operator as a
2959: differential square root of a Laplacian.
2960:
2961: This prescription recovers the usual geodesic distance on a Riemannian
2962: manifold, by the following result (\cf
2963: \cite{Co-fr}).
2964:
2965: \begin{lem}\label{geod-distance}
2966: On a Riemannian spin manifold the geodesic distance $d(x,y)$
2967: between two points is computed by the formula
2968: \begin{equation}
2969: d(x,y) = \sup \{ \vert f(x) - f(y) \vert \, ; \ f \in \cA
2970: \, , \ \Vert [D,f] \Vert \leq 1 \} \label{geoddist}
2971: \end{equation}
2972: where $D$ is the Dirac operator, $D=ds^{-1}$, and $\cA$ is the
2973: algebra of smooth functions.
2974: \end{lem}
2975:
2976: This essentially follows from the fact that the quantity $\Vert
2977: [D,f] \Vert$ can be identified with the Lipschitz norm of the
2978: function $f$,
2979: $$ \Vert [D,f] \Vert = {\rm ess} \sup_{x\in M} \| (\nabla f)_x \|
2980: = \sup_{x\neq y}\frac{ | f(x)-f(y) | }{d(x,y)}. $$
2981:
2982: Notice that, if $ds$ has the dimension of a length $L$, then $D$
2983: has dimension $L^{-1}$ and the above expression for $d(x,y)$ also
2984: has the dimension of a length. On a Riemannian spin manifold $X$,
2985: the condition $\| [D,f] \|\leq 1$, for $D$ the Dirac operator, is
2986: equivalent to the condition that $f$ is a Lipschitz function with
2987: Lipschitz constant $c\leq 1$.
2988:
2989: The advantage of the definition \eqref{dsx-x}, \eqref{Fermprop} of the
2990: line element is that it is of a {\em spectral} and operator theoretic
2991: nature, hence it extends to the noncommutative setting.
2992:
2993: The structure that combines the $K$-homology fundamental cycle with
2994: the spectral definition of the line element $ds$ is the notion of {\em
2995: spectral triple} $(\cA,\cH,D)$ (\cf \cite{ConnesS3}, \cite{CoMo}).
2996:
2997: \begin{defn}\label{def-triple}
2998: A (compact) noncommutative geometry is a triple
2999: \begin{equation}
3000: (\cA ,\cH ,D)
3001: \label{triple}
3002: \end{equation}
3003: where $\cA$ is a unital algebra
3004: represented concretely as an
3005: algebra of bounded operators on the Hilbert space $\cH$. The unbounded
3006: operator $D$ is the inverse of the line element
3007: \begin{equation}
3008: ds=1/D.\label{dsD}
3009: \end{equation}
3010: Such a triple $(\cA ,\cH ,D)$ is requires to satisfy the properties:
3011: \begin{enumerate}
3012: \item $[D,a]$ is bounded for any $a \in A^\infty$, a dense subalgebra of
3013: the $C^*$-algebra $A=\bar\cA$.
3014: \item $D=D^*$ and $(D+\lambda)^{-1}$ is a compact operator, for
3015: all $\lambda \not\in \R$.
3016: \end{enumerate}
3017: We say that a spectral triple $(\cA ,\cH ,D)$ is even if the
3018: Hilbert space $\cH$ has a $\Z/2$-grading by an operator $\gamma$
3019: satisfying
3020: \begin{equation}\label{gammaZ2}
3021: \gamma =\gamma^*, \ \ \ \gamma^2 =1, \ \ \ \gamma\, D =- D \,
3022: \gamma, \ \ \ \gamma \, a = a\, \gamma \ \ \forall a \in A.
3023: \end{equation}
3024: \end{defn}
3025:
3026: This definition is entirely spectral. The elements of the algebra (in
3027: general noncommutative) are operators and the line element is also an
3028: operator. The polar decomposition $D=|D|F$ recovers the Fredholm
3029: module $F$ defining the fundamental class in $K$-homology.
3030: The formula for the geodesic distance extends to this
3031: context as follows.
3032:
3033: \begin{defn}\label{distanceNC}
3034: Let $\varphi_i: A \to \C$, for $i=1,2$, be states on $A$, \ie
3035: normalized positive linear functionals on $A$ with $\varphi_i (1) =
3036: 1$ and $\varphi_i (a^* a) \geq 0$ for all $a \in A$. Then the
3037: distance between them is given by the formula
3038: \begin{equation}
3039: d(\varphi_1 ,\varphi_2) = \sup \{ \vert \varphi_1 (a) -
3040: \varphi_2 (a) \vert \ ;
3041: \ a\in A \ , \ \Vert [D,a]\Vert \leq 1 \} \, .
3042: \label{geodNC}
3043: \end{equation}
3044: \end{defn}
3045:
3046: \smallskip
3047:
3048: A spectral triple $(\cA,\cH,D)$ is of {\em metric dimension} $p$,
3049: or $p$-summable, if $|D|^{-1}$ is an infinitesimal of order $1/p$
3050: (\ie $|D|^{-p}$ is an infinitesimal of order one). Here $p<
3051: \infty$ is a positive real number. A spectral triple $(\cA,\cH,D)$ is
3052: $\theta$-summable if $\Tr(e^{-t D^2}) <\infty$ for all $t>0$.
3053: The latter case corresponds to an infinite dimensional geometry.
3054:
3055: Spectral triples also provide a more refined notion of dimension
3056: besides the metric dimension (summability). It is given by the
3057: {\em dimension spectrum}, which is not a number but a subset of
3058: the complex plane.
3059:
3060: More precisely, let $(\cA,\cH,D)$ be
3061: a spectral triple satisfying
3062: the regularity hypothesis
3063: \begin{equation}
3064: a \ \hbox{and } \ [D,a] \ \in \ \cap_k {\rm Dom}( \delta^k ) , \ \forall \,
3065: a\in A^\infty ,
3066: \label{regS3}
3067: \end{equation}
3068: where $\delta$ is the derivation $\delta(T) = [\vert D \vert,T]$, for any
3069: operator $T$.
3070: Let $\cB$ denote the algebra generated by $\delta^k (a)$ and $\delta^k
3071: ([D,a])$.
3072: The dimension spectrum of the triple $(\cA,\cH,D)$
3073: is the subset $\Sigma \subset \C$ consisting of all the singularities of
3074: the analytic functions $\zeta_b (z)$ obtained by continuation of
3075: \begin{equation}
3076: \zeta_b (z) = \Tr (b \vert D \vert^{-z}), \ \ \ \Re(z)
3077: > p \ , \ \ \ b\in \cB \, . \label{zeta-b}
3078: \end{equation}
3079:
3080:
3081: \begin{ex} {\em
3082: Let $M$ be a smooth compact Riemannian spin manifold, and
3083: $(\cA,\cH,D)$ is the corresponding spectral triple given by the
3084: algebra of smooth functions, the space of spinors, and the Dirac
3085: operator. Then the metric dimension agrees with the
3086: usual dimension $n$ of $M$. The dimension spectrum of $M$ is the
3087: set $\{ 0,1,\ldots ,n\}$, where $n=\dim M$, and it is simple.
3088: (Multiplicities appear for singular manifolds.)}
3089: \end{ex}
3090:
3091: It is interesting in the case of an ordinary Riemannian manifold $M$
3092: to see the meaning of the points in the dimension spectrum that are
3093: smaller than $n=\dim M$. These are dimensions in which the space
3094: ``manifests itself nontrivially'' with some interesting geometry.
3095:
3096: For instance, at the point $n=\dim M$ of the dimension spectrum one
3097: can recover the
3098: volume form of the Riemannian metric by the equality (valid up to
3099: a normalization constant \cf \cite{Co94})
3100: \begin{equation}
3101: {\int \!\!\!\!\!\! -}f \,\vert ds\vert^n = \, \int_{M_n} f \,
3102: \sqrt{g} \ d^n x \, , \label{nvolume}
3103: \end{equation}
3104: where the integral $\cutint T$ is given (\cf \cite{Co94}) by the
3105: Dixmier trace (\cf \cite{Dix}) generalizing the Wodzicki residue of
3106: pseudodifferential operators (\cf \cite{Wo}).
3107:
3108: One can also consider integration $\cutint ds^k$ in any other
3109: dimension in the
3110: dimension spectrum, with $ds=D^{-1}$ the line element.
3111: In the case of a Riemannian manifold one finds
3112: other important curvature expressions. For instance, if $M$ is a
3113: manifold of dimension $\dim M=4$, when one considers integration in
3114: dimension $2$ one finds the Einstein--Hilbert action. In fact,
3115: a direct computation
3116: yields the following result (\cf \cite{Kast} \cite{K-W}):
3117:
3118: \begin{prop}
3119: Let $dv=\sqrt{g} \ d^4 x$ denote the volume form, $ds=D^{-1}$ the
3120: length element, \ie the inverse of the Dirac operator, and $r$ the
3121: scalar curvature. We obtain:
3122: \begin{equation}
3123: {\int \!\!\!\!\!\! -} \, ds^2 = \frac{-1}{48\pi^2} \, \int_{M_4} r
3124: \, \, dv \, . \label{EHaction}
3125: \end{equation}
3126: \end{prop}
3127:
3128: In general, one obtains the scalar curvature of an $n$-dimensional manifold
3129: from the integral $\cutint ds^{n-2}$.
3130:
3131: \medskip
3132:
3133: Many interesting examples of spectral triples just satisfy the
3134: conditions stated in Definition \ref{def-triple}. However, there are
3135: significant case where more refined properties of manifolds carry over
3136: to the noncommutative case, such as the presence of a real structure
3137: (which makes it possible to distinguish between $K$-homology and
3138: $KO$-homology) and the ``order one condition'' for the Dirac
3139: operator. These properties are described as follows (\cf \cite{Co3}
3140: and \cite{Coo2}).
3141:
3142: \begin{defn}\label{realstr}
3143: A real structure on an $n$-dimensional spectral triple
3144: $(\cA,\cH,D)$ is an antilinear isometry $J: \cH \to \cH$, with the
3145: property that
3146: \begin{equation}\label{per8}
3147: J^2 = \varepsilon, \ \ \ \ JD = \varepsilon' DJ, \ \ \text{and} \ \
3148: J\gamma = \varepsilon'' \gamma
3149: J \, \text{(even case)}.
3150: \end{equation}
3151: The numbers $\varepsilon ,\varepsilon' ,\varepsilon'' \in \{ -1,1\}$
3152: are a function of $n \mod 8$ given by
3153:
3154: \begin{center}
3155: \begin{tabular}
3156: {|c| r r r r r r r r|} \hline {\bf n }&0 &1 &2 &3 &4 &5 &6 &7 \\
3157: \hline \hline
3158: $\varepsilon$ &1 & 1&-1&-1&-1&-1& 1&1 \\
3159: $\varepsilon'$ &1 &-1&1 &1 &1 &-1& 1&1 \\
3160: $\varepsilon''$&1 &{}&-1&{}&1 &{}&-1&{} \\ \hline
3161: \end{tabular}
3162: \end{center}
3163:
3164:
3165: Moreover, the action of $\cA$ satisfies the commutation rule
3166: \begin{equation}\label{comm-rule}
3167: [a,b^0] = 0 \quad \forall \, a,b \in \cA,
3168: \end{equation}
3169: where
3170: \begin{equation}\label{b0}
3171: b^0 = J b^* J^{-1} \qquad \forall b \in \cA,
3172: \end{equation}
3173: and the operator $D$ satisfies
3174: \begin{equation}\label{order1}
3175: [[D,a],b^0] = 0 \qquad \forall \, a,b \in \cA \, .
3176: \end{equation}
3177: \end{defn}
3178:
3179: The anti-linear isometry $J$ is given, in ordinary Riemannian
3180: geometry, by the charge conjugation operator acting on spinors. In
3181: the noncommutative case, this is replaced by the Tomita
3182: antilinear conjugation operator (\cf \cite{Tak}).
3183:
3184: \smallskip
3185:
3186: In \cite{Co3} and \cite{FGV} Theorem 11.2, necessary and sufficient
3187: conditions are given that a spectral triple $(\cA,\cH,D)$ (with real
3188: structure $J$) should fulfill in order to come from an ordinary
3189: compact Riemannian spin manifold:
3190:
3191: \begin{enumerate}
3192: \item $ds=D^{-1}$ is an infinitesimal of order $1/n$.
3193: \item There is a real structure in the sense of Definition
3194: \ref{realstr}.
3195: \item The commutation relation \eqref{order1} holds (this is
3196: $[[D,a],b]=0$, for all $a,b\in \cA$, when $\cA$ is commutative).
3197: \item The regularity hypothesis of \eqref{regS3} holds:
3198: $a$ and $[D,a]$ are in $\cap_k {\rm Dom}( \delta^k)$ for all $a\in \cA_\infty$.
3199: \item There exists a Hochschild cycle $c\in Z_n(\cA_\infty,\cA_\infty)$ such
3200: that its representation $\pi(c)$ on $\cH$ induced by
3201: $$ \pi(a^0\otimes\cdots\otimes a^n)=a^0 [D,a^1]\cdots [D,a^n] $$
3202: satisfies $\pi(c)=\gamma$, for $\gamma$ as in \eqref{gammaZ2}, in
3203: the even case, and $\pi(c)=1$ in the odd case.
3204: \item The space $\cH^\infty= \cap_k {\rm Dom}( D^k)$ is a finite
3205: projective $\cA$-module, endowed with a $\cA$-valued inner product
3206: $\langle \xi, \eta \rangle_{\cA}$ defined by
3207: $$ \langle a \xi, \eta \rangle = \cutint a\, \langle \xi, \eta
3208: \rangle_{\cA}\, ds^n. $$
3209: \item The intersection form
3210: \begin{equation}
3211: K_* (\cA) \times K_* (\cA) \to \Z
3212: \label{intform}
3213: \end{equation}
3214: obtained from the Fredholm index of $D$ with
3215: coefficients in $K_* (\cA \otimes \cA^0)$
3216: is invertible.
3217: \end{enumerate}
3218:
3219: When $\cA$ is commutative, the above conditions characterize a
3220: smooth Riemannian manifold $M$, with $\cA_\infty=\cC^\infty(M)$ (we
3221: refer to \cite{FGV} for the precise statement). However, the
3222: conditions can be stated without any commutativity assumption on
3223: $\cA$. They are satisfied, for instance, by the isospectral
3224: deformations of \cite{CLa}, which we discuss in Section
3225: \ref{isodef}. Another very significant noncommutative example is the
3226: standard model of elementary particles (\cf \cite{Co3}), which we
3227: discuss in Section \ref{stmod}.
3228:
3229:
3230: \bigskip
3231:
3232: Another example of spectral triple associated to a classical
3233: space, which is not classically a smooth manifold, is the case of
3234: manifolds with singularities. In particular, one can consider the
3235: case of an isolated conical singularity. This case was studied by
3236: Lescure \cite{Lescure}.
3237:
3238: \smallskip
3239:
3240:
3241: Let $X$ be a manifold with an isolated conical singularity. The
3242: cone point $c\in X$ has the property that there is a neighborhood
3243: $U$ of $c$ in $X$, such that $U\smallsetminus \{ c \}$ is of the
3244: form $(0, 1] \times N$, with $N$ a smooth compact manifold, and
3245: metric $g|_U = dr^2 + r^2 g_N$, where $g_N$ is the metric on $N$.
3246:
3247: \smallskip
3248:
3249: A natural class of differential operators on manifolds with
3250: isolated conical singularities is given by the elliptic operators
3251: of Fuchs type, acting on sections of a bundle $E$. These are
3252: operators whose restriction over $(0, 1] \times N$ takes the form
3253: $$ r^{-\nu} \sum_{k=0}^d a_k(r) (-r \partial_r)^k, $$
3254: for $\nu \in \R$ and $a_k \in C^\infty([0,1], {\rm Diff}^{d-k}(N,
3255: E|_N))$, which are elliptic with symbol $\sigma_M(D)=\sum_{k=0}^d
3256: a_k(0) z^k$ that is an elliptic family parameterized by $Im(z)$. In
3257: particular, operators of Dirac type are elliptic of Fuchs type. For
3258: such an operator $D$, which is of first order and symmetric, results
3259: of Chou \cite{chou}, Br\"uning, Seeley \cite{BruSee} and Lesch
3260: \cite{Lesch} show that its self-adjoint extension has discrete
3261: spectrum, with $(n+1)$-summable resolvent, for $\dim X= n$.
3262:
3263: \smallskip
3264:
3265: The algebra that is used in the construction of the spectral
3266: triple is $\cA= C^\infty_c(X) \oplus \C$, the algebra of functions
3267: that are smooth on $X\smallsetminus \{ c \}$ and constant near the
3268: singularity. The Hilbert space on which $D$ acts is chosen from a
3269: family of weighted Sobolev spaces. Roughly, one defines weighted
3270: Sobolev spaces that look like the standard Sobolev space on the
3271: smooth part and on the cone are defined locally by norms
3272: $$
3273: \| f \|_{s,\gamma}^2 = \int_{\R^*_+ \times \R^{m-1}} \left(1 + (\log
3274: t)^2 +\xi^2\right)^s \left| \widehat{(r^{-\gamma +1/2} f)}
3275: (t,\xi)\right|^2 \frac{dt}{t} d\xi,
3276: $$
3277: where $\hat f$ denotes Fourier transform on the group $\R_+^*
3278: \times \R^{m-1}$.
3279:
3280: \smallskip
3281:
3282: Then one obtains the following result (\cf \cite{Lescure})
3283:
3284: \begin{thm}
3285: The data $(\cA, \cH, D)$ given above define a spectral triple. In
3286: particular, the zeta functions $\Tr(a |D|^{-z})$ admit analytic
3287: continuation to $\C \smallsetminus \Sigma$, where the dimension
3288: spectrum is of the form
3289: $$ \Sigma=\{ \dim X - k, k\in \N \}, $$
3290: with multiplicities $\leq 2$.
3291: \end{thm}
3292:
3293: \smallskip
3294:
3295: The analysis of the zeta functions uses the heat kernel
3296: $$ \Tr (| D |^{-z}) = \frac{1}{\Gamma(z/2)} \int_0^\infty t^{z/2
3297: -1} \Tr (e^{-t D^2}) dt, $$ for which one can rely on the results of
3298: \cite{chou} and \cite{Lesch}. The case of $\Tr(a |D|^{-z})$ of the
3299: form $\Tr(Q |D|^{-z})$ with $Q\in \Psi^\ell_c(E)$, is treated by
3300: splitting $Q|D|^{-z}$ as a sum of a contribution from the smooth
3301: part and one from the singularity.
3302:
3303: \smallskip
3304:
3305: The Chern character for this spectral triple gives a map
3306: $$ Ch: K_*(X) \to H_*(X,\C), $$
3307: where we have $K^*(X)\cong K_*(\cA)$ and $H_*(X,\C) \cong
3308: PHC^*(\cA)$, the periodic cyclic homology of the algebra $\cA$.
3309:
3310: The cocycles $\varphi_n$ in the $(b,B)$-bicomplex for the algebra
3311: $\cA$ have also been computed explicitly and are of the form
3312: $$ \varphi_n (a_0,\ldots,a_n) =\nu_n \int_X a_0 da_1 \wedge \cdots
3313: \wedge da_n \wedge \hat A(X) \wedge Ch(E),
3314: $$
3315: for $n \geq 1$, while for $n=0$, $\lambda\in \C$,
3316: $\varphi_0(a+\lambda)=\int_X a \hat A(X)\wedge Ch(E) + \lambda
3317: \Ind(D_+)$.
3318:
3319: \bigskip
3320:
3321: \section{Cantor sets and fractals}\label{fractals}
3322:
3323: An important class of $C^*$-algebras are those obtained as
3324: direct limits of a sequence of finite dimensional
3325: subalgebras and embeddings. These are called {\em approximately finite
3326: dimensional}, or simply AF-algebras.
3327:
3328: An AF algebra $\cA$ is determined by a diagram of finite
3329: dimensional algebras and inclusions, its Bratteli diagram
3330: \cite{Bratteli}, and from the diagram itself it is possible to
3331: read a lot of the structure of the algebra, for instance its ideal
3332: structure. Some simple examples of algebras that belong to this
3333: class are:
3334:
3335: \begin{ex}\label{Cantorset} {\it
3336: An example of a commutative AF is the algebra of complex valued
3337: continuous functions on a Cantor set, where a Bratteli diagram is
3338: determined by a decreasing family of disjoint intervals covering
3339: the Cantor set. } \end{ex}
3340:
3341: A non--commutative example of AF algebra is given by the algebra
3342: of the canonical anticommutation relations of quantum mechanics.
3343:
3344: \begin{ex} {\it Consider a real Hilbert space $\cE$ and a
3345: linear map $\cE \to \cB(\cH)$, $f \mapsto T_f$, to bounded operators
3346: in a Hilbert space $\cH$, satisfying
3347: $$ T_f T_g +T_g T_f =0 $$
3348: $$ T_f^* T_g+ T_g T_f^* = \langle g,f \rangle I, $$
3349: and the algebra $\cA$ generated by all the operators $T_f$
3350: satisfying these relations. } \label{exCAR}
3351: \end{ex}
3352:
3353: A survey with many examples of AF algebras and their properties is
3354: given for instance in \cite{Davidson}.
3355:
3356: Let $\cA$ be a commutative AF ${\rm C}^*$-algebra. A
3357: commutative AF algebra $\cA$ is spanned by its projections,
3358: since finite dimensional commutative algebras are generated by
3359: orthogonal projections. This condition is equivalent to the
3360: spectrum $\Lambda=\Sp(\cA)$ of the algebra being a totally
3361: disconnected compact Hausdorff space, typically a Cantor set.
3362: Realizing such Cantor set as the intersection of a decreasing
3363: family of disjoint intervals covering $\Lambda$ also provides a
3364: Bratteli diagram for the AF algebra $\cA=C(\Lambda)$.
3365:
3366: As described in \cite{Co94},
3367: in order to construct the Hilbert space $\cH$ for a Cantor set
3368: $\Lambda\subset \R$, let $J_k$ be the collection of bounded open
3369: intervals in $\R\setminus \Lambda$. We denote by $L=\{ \ell_k
3370: \}_{k\geq 1}$ the countable collection of the lengths of the
3371: intervals $J_k$. We can assume that the lengths are ordered
3372: \begin{equation}\label{ord-ell}
3373: \ell_1 \geq \ell_2 \geq \ell_3 \geq \cdots \geq \ell_k \cdots
3374: >0.
3375: \end{equation}
3376: We also denote by $E=\{ x_{k,\pm} \}$ the set of the endpoints of
3377: the intervals $J_k$, with $x_{k,+}>x_{k,-}$. Consider the Hilbert
3378: space
3379: \begin{equation}\label{H-endpts}
3380: \cH := \ell^2 (E)
3381: \end{equation}
3382:
3383: Since the endpoints of the $J_k$ are points of $\Lambda$, there is
3384: an action of ${\rm C}(\Lambda)$ on $\cH$ given by
3385: \begin{equation}\label{act-CS}
3386: f\cdot \xi (x) = f(x) \xi(x), \ \ \ \forall f\in {\rm
3387: C}(\Lambda),\, \, \forall \xi \in \cH,\, \, \forall x \in E.
3388: \end{equation}
3389:
3390: A sign operator $F$ can be obtained (\cf \cite{Co94})
3391: by choosing the closed subspace
3392: $\hat\cH\subset \cH$ given by
3393: \begin{equation}\label{PH}
3394: \hat\cH=\{ \xi \in \cH : \xi(x_{k,-})=\xi(x_{k,+}), \, \, \forall
3395: k \}.
3396: \end{equation}
3397: Then $F$ has eigenspaces $\hat\cH$ with eigenvalue $+1$ and
3398: $\hat\cH^\perp$ with eigenvalue $-1$, so that, when restricted to
3399: the subspace $\cH_k$ of coordinates $\xi(x_{k,+})$ and
3400: $\xi(x_{k,-})$, the sign $F$ is given by
3401: $$ F|_{\cH_k} = \left(\begin{array}{cc} 0&1\\1&0
3402: \end{array}\right). $$
3403:
3404: Finally, a Dirac operator $D=|D| F$ is obtained as
3405: \begin{equation}\label{fractalDirac}
3406: D|_{\cH_k} \left(\begin{array}{c} \xi(x_{k,+})\\ \xi(x_{k,-})
3407: \end{array}\right) = \ell_k^{-1} \cdot \, \left(\begin{array}{c}
3408: \xi(x_{k,-})\\ \xi(x_{k,+}) \end{array}\right).
3409: \end{equation}
3410:
3411: We then obtain the following result.
3412:
3413: \begin{prop}
3414: Let $\Lambda\subset \R$ be a Cantor set. Let $\cA_\infty\subset
3415: C(\Lambda)$ be the dense subalgebra of locally constant functions on
3416: the Cantor set. Then the data $(\cA,\cH,D)$ form a spectral triple,
3417: with $\cH$ as in
3418: (\ref{H-endpts}), the action (\ref{act-CS}), and $D$ as in
3419: (\ref{fractalDirac}). The zeta function satisfies
3420: $$ \Tr( |D|^{-s} ) = 2 \zeta_L (s), $$
3421: where $\zeta_L(s)$ is the geometric zeta function of
3422: $L=\{ \ell_k \}_{k\geq 1}$, defined as
3423: \begin{equation}\label{zeta-fractal}
3424: \zeta_L(s):= \sum_k \ell_k^s.
3425: \end{equation}
3426: \label{propZCS}
3427: \end{prop}
3428:
3429: \smallskip
3430:
3431: These zeta functions are related to the theory of Dirichlet series
3432: and to other arithmetic zeta functions, and also to Ruelle's
3433: dynamical zeta functions (\cf \cite{LapFra}).
3434:
3435: \smallskip
3436:
3437: For example, for the classical middle-third Cantor set, we have
3438: set of lengths $\ell_k = 3^{-k}$ and multiplicities $m_k=2^{k-1}$,
3439: for $k\geq 1$, so that we obtain
3440: \begin{equation}\label{zeta-CS}
3441: \Tr(|D|^{-s})= 2\zeta_L(s)=\sum_{k\geq 1} 2^k 3^{-sk} =
3442: \frac{2\cdot 3^{-s}}{1-2\cdot 3^{-s}}.
3443: \end{equation}
3444: This shows that the dimension spectrum of the spectral triple of a
3445: Cantor set has points off the real line. In fact, the set of poles
3446: of (\ref{zeta-CS}) is
3447: \begin{equation}\label{poles-CS}
3448: \left\{ \frac{\log 2}{\log 3} + \frac{2\pi i n}{\log 3}
3449: \right\}_{n\in \Z}.
3450: \end{equation}
3451:
3452:
3453: In this case the dimension spectrum lies on a vertical line and it
3454: intersects the real axis in the point $D=\frac{\log 2}{\log 3}$
3455: which is the Hausdoff dimension of the ternary Cantor set. The
3456: same is true for other Cantor sets, as long as the self-similarity
3457: is given by a unique contraction (in the ternary case the original
3458: interval is replaced by two intervals of lengths scaled by 1/3).
3459:
3460: If one considers slightly more complicated fractals in $\R$, where
3461: the self-similarity requires more than one scaling map, the
3462: dimension spectrum may be correspondingly more complicated. This
3463: can be seen in the case of the Fibonacci Cantor set, for instance (\cf
3464: \cite{LapFra}).
3465:
3466: The Fibonacci Cantor set $\Lambda$ is obtained from the interval
3467: $I=[0,4]$ by successively removing $F_{n+1}$ open intervals
3468: $J_{n,j}$ of lengths $\ell_n =1/2^n$
3469: according to the rule of Figure \ref{FigFC}.
3470: We can associate to this Cantor set the
3471: commutative AF algebra $\cA=C(\Lambda)$.
3472:
3473: \begin{center}
3474: \begin{figure}
3475: \includegraphics[scale=0.35]{FibCantor.eps}
3476: \caption{The Fibonacci Cantor set.
3477: \label{FigFC}}
3478: \end{figure}
3479: \end{center}
3480:
3481: To obtain the Hilbert space we consider again the set $E$ of
3482: endpoints $x_{n,j,\pm}$ of the intervals $J_{n,j}$ and we take
3483: $\cH=\ell^2(E)$. We define the Dirac operator as in the previous
3484: case, and we again consider the dense involutive subalgebra
3485: $\cA_\infty$ of locally constant functions.
3486:
3487: The data $(\cA,\cH,D)$ give a spectral triple. The Dirac
3488: operator is related to the geometric zeta function of the
3489: Fibonacci Cantor set by
3490: $$ Tr(|D|^{-s})= 2 \zeta_F(s) =\frac{2}{1-2^{-s}-4^{-s}}, $$
3491: where the geometric zeta function is $\zeta_F(s)=\sum_n F_{n+1}
3492: 2^{-ns}$, with $F_n$ the Fibonacci numbers.
3493:
3494: A simple argument shows that the dimension
3495: spectrum is given by the set
3496: $$ \Sigma= \left\{
3497: \frac{\log\phi}{\log 2} + \frac{2\pi i n}{\log 2} \right\}_{n\in
3498: \Z} \cup \left\{ -\frac{\log\phi}{\log 2} + \frac{2\pi i
3499: (n+1/2)}{\log 2} \right\}_{n\in \Z}, $$ where
3500: $\phi=\frac{1+\sqrt{5}}{2}$ is the golden ratio.
3501:
3502:
3503: \medskip
3504:
3505: Recent results on the noncommutative geometry of fractals and Cantor
3506: sets and spectral triple constructions for AF algebras can be found
3507: in \cite{AntChris}, \cite{GuIso1}, \cite{GuIso2}. The construction
3508: in \cite{AntChris} is in fact a spectral triple for the dual group
3509: of the Cantor set seen as the product of countably many copies of
3510: the group $\Z/2$. The recent work
3511: \cite{IvChris} shows that it is easy to describe a compact metric
3512: space exactly (\ie recovering the metric) via a spectral triple,
3513: which is a sum of two-dimensional modules, but spectral triples
3514: carry much more information than just the one regarding the metric.
3515:
3516:
3517: \bigskip
3518:
3519: \section{Spaces of dimension $z$ and DimReg}\label{DimReg}
3520:
3521: In perturbative quantum field theory, one computes expectation values
3522: of observables via a formal series, where the terms are parameterized
3523: by Feynman graphs and reduce to ordinary finite dimensional integrals
3524: in momentum space of expressions assigned to the graphs by the
3525: Feynman rules. These expressions typically produce divergent
3526: integrals. For example, in the example of the scalar $\phi^3$ theory
3527: in dimension $D=4$ or $D=4+2\N$, one encounters a divergence already
3528: in the simplest one loop diagram, with corresponding integral (in
3529: Euclidean signature)
3530: \begin{equation}\label{Fint}
3531: \hbox{\psfig{figure=feydiag.eps}\quad =} \int \frac{1}{k^2+m^2}\,
3532: \frac{1}{((p+k)^2+m^2)} \, d^D k.
3533: \end{equation}
3534:
3535: One needs therefore a regularization procedure for these divergent
3536: integrals. The regularization most commonly adopted in quantum field
3537: theory computation is ``Dimensional Regularization (DimReg) and Minimal
3538: Subtraction (MS)''. The method was introduced in the '70s in
3539: \cite{BoGia} and \cite{tHV} and it has the advantage of preserving
3540: basic symmetries.
3541:
3542: The regularization procedure of DimReg is essentially based on the use
3543: of the formula
3544: \begin{equation}\label{gaussienne}
3545: \int\,e^{-\lambda\,k^2}\,d^{d}
3546: k\,=\,{\pi^{d/2}}\,{\lambda^{-d/2}}\,,
3547: \end{equation}
3548: to {\em define} the meaning of the integral in
3549: $d=(D-z)$-dimensions, for $z\in \C$ in a neighborhood of zero.
3550: For instance, in the case of \eqref{Fint}, the procedure of
3551: dimensional regularization yields the result
3552: $$ \pi^{(D-z)/2} \Gamma\left(\frac{4-D+z}{2}\right) \int_0^1 \left( (
3553: x-x^2)p^2 +m^2\right)^{\frac{D-z-4}{2}} \, dx .$$
3554:
3555: \medskip
3556:
3557: In the recent survey \cite{ManDim}, Yuri Manin refers to DimReg as
3558: ``dimensions in search of a space''\footnote{Nicely reminiscent of
3559: Pirandello's play ``Six characters in search of an author''.}.
3560: Indeed, in the usual approach in perturbative quantum field theory, the
3561: dimensional regularization procedure is just regarded as a formal rule
3562: of analytic continuation of formal (divergent) expressions in integral
3563: dimensions $D$ to complex values of the variable $D$.
3564:
3565: However, using noncommutative geometry, it is possible to construct
3566: actual spaces (in the sense of noncommutative Riemannian geometry)
3567: $X_z$ whose dimension (in the sense of dimension spectrum) is a
3568: point $z\in \C$ (\cf \cite{CoMar-an}).
3569:
3570: It is well known in the physics literature that there are problems
3571: related to using dimensionl regularization in chiral theory, because
3572: of how to give a consistent prescription on how to extend the
3573: $\gamma_5$ (the product of the matrices $\gamma^i$ when $D=4$) to
3574: noninteger dimension $D-z$. It turns out that a prescription
3575: known as Breitenlohner-Maison (\cite{BM}, \cite{Collins}) admits an
3576: interpretation in terms of the cup product of spectral triples, where
3577: one takes the product of the spectral triple associated to the ordinary
3578: geometry in the integer dimension $D$ by
3579: a spectral triple $X_z$ whose dimension spectrum is reduced
3580: to the complex number $z$ (\cf \cite{CoMar-an}).
3581:
3582: We illustrate here the construction for the case where $z\in
3583: \R^*_+$. The more general case of $z\in \C$ is more delicate.
3584:
3585: One needs to work in a slightly modified setting for spectral triples,
3586: which is given by the type II spectral triples (\cf \cite{BF},
3587: \cite{CPRS}, \cite{CPRS1}). In this setting the usual type I trace
3588: of operators in $\cL(\cH)$ is replaced by the
3589: trace on a type II$_{\infty}$ von-Neumann algebra.
3590:
3591: One considers a self-adjoint operator $Y$, affiliated to
3592: a type II$_{\infty}$ factor $N$, with spectral measure given by
3593: \begin{equation}\label{Sp-measure}
3594: \Tr_N (\chi_E(Y))=\,\frac{1}{2}\,
3595: \int_E\,dy
3596: \end{equation}
3597: for any interval $E\subset \R$, with characteristic function
3598: $\chi_E$.
3599:
3600: If $Y=F\, |Y|$ is the polar decomposition of $Y$, one sets
3601: \begin{equation}\label{newdz}
3602: D_z=\, \rho(z) \,F\, |Y|^{1/z}
3603: \end{equation}
3604: with the complex
3605: power $|Y|^{1/z}$ defined by the
3606: functional calculus.
3607: The normalization constant $ \rho(z)$
3608: is chosen to be
3609: \begin{equation}\label{rhoz}
3610: \rho(z)=\,\pi^{-\frac{1}{2}}\,
3611: \left(\Gamma(\frac{z}{2}+1)\right)^{\frac{1}{z}}
3612: \end{equation}
3613: so that one obtains
3614: \begin{equation}\label{basic}
3615: \Tr \left(e^{-\lambda D^2}\right)=\,
3616: \,{\pi^{z/2}}\,{\lambda^{-z/2}} \ \ \ \forall \lambda \in \R^*_+\,.
3617: \end{equation}
3618:
3619: This gives a geometric meaning to the basic formula \eqref{gaussienne}
3620: of DimReg. The algebra $\cA$ of the spectral triple $X_z$ can be made
3621: to contain any operator $a$ such that $[D_z,a]$ is bounded
3622: and both $a$
3623: and $[D_z,a]$ are smooth for the ``geodesic flow"
3624: \begin{equation}\label{geod}
3625: T\mapsto \,e^{it|D_z|}\,T\,e^{-it|D_z|}.
3626: \end{equation}
3627: The dimension spectrum of $X_z$ is reduced to the
3628: single point $z$, since
3629: \begin{equation}\label{trace'}
3630: {\rm Trace}'_N(
3631: (D^2_z)^{-s/2})=\,\rho^{-s}\, \int_1^\infty\,u^{-s/z}\,du=
3632: \,\rho^{-s}\,\frac{z}{s-z}
3633: \end{equation}
3634: has a single (simple) pole
3635: at $s=z$ and is absolutely convergent in the half space
3636: Re$(s/z)>1$. Here ${\rm Trace}'_N$ denotes the trace with an
3637: infrared cutoff (\ie integrating outside $|y|<1$).
3638:
3639: \bigskip
3640:
3641: \section{Local algebras in supersymmetric QFT}\label{susy}
3642:
3643:
3644: It is quite striking that the general framework of noncommutative
3645: geometry is suitable not only for handling finite dimensional
3646: spaces (commutative or not, of non-integer dimension etc.) but is
3647: also compatible with infinite dimensional spaces. We already saw in
3648: Section \ref{discgr} that discrete groups of exponential growth
3649: naturally give rise to noncommutative spaces which are described by
3650: a $\theta$-summable spectral triple, but not by a finitely summable
3651: spectral triple. This is characteristic of an infinite dimensional
3652: space and in that case, as we saw for discrete groups, cyclic
3653: cohomology needs to be extended to {\em entire} cyclic cohomology. A
3654: very similar kind of noncommutative spaces arises from Quantum Field
3655: Theory in the supersymmetric context \cite{Co94} Section
3656: IV.9.$\beta$. We briefly recall this below and then explain
3657: open questions also in the context of supersymmetric theories.
3658:
3659: The simplest example to understand the framework is that of the free
3660: Wess-Zumino model in two dimensions, a supersymmetric free field
3661: theory in a two dimensional space-time where space is compact
3662: (\cite{Co94}). Thus space is a circle $S^1$ and space-time is a
3663: cylinder $C=\,S^1\times \R$ endowed with the Lorentzian metric. The
3664: fields are given by a complex scalar bosonic field $\phi$ of mass
3665: $m$ and a spinor field $\psi$ of the same mass. The Lagrangian of
3666: the theory is of the form $\cL=\,\cL_b +\cL_f$ where,
3667: $$
3668: \cL_b=\,\frac{1}{2}(|\partial_0 \phi|^2- |\partial_1 \phi|^2 - m^2
3669: |\phi|^2)
3670: $$
3671: and for the fermions,
3672: $$
3673: \cL_f=\,i\,\bar \psi\,\gamma^\mu\,\partial_\mu\,\psi-\,m\,\bar
3674: \psi\,\psi
3675: $$
3676: where the spinor field is given by a column matrix, with $\bar
3677: \psi=\,\gamma^0\,\psi^*$ and the $\gamma^\mu$ are two by two Pauli
3678: matrices, anticommuting, self-adjoint and of square $1$.
3679:
3680: The Hilbert space of the quantum theory is the tensor product
3681: $\cH=\,\cH_b \otimes \cH_f$ of the bosonic one $\cH_b$ by the
3682: fermionic one $\cH_f$. The quantum field $\phi(x)$ and its conjugate
3683: momentum $\pi(x)$ are operator valued distributions in $\cH_b$ and
3684: the bosonic Hamiltonian is of the form
3685: $$
3686: H_b=\,\int_{S^1} \,:\,|\pi(x)|^2 + |\partial_1 \phi(x)|^2 + m^2
3687: |\phi(x)|^2 \,:\, dx
3688: $$
3689: where the Wick ordering takes care of an irrelevant additive
3690: constant. The fermionic Hilbert space $\cH_f$ is given by the Dirac
3691: sea representation which simply corresponds to a suitable spin
3692: representation of the infinite dimensional Clifford algebra
3693: containing the fermionic quantum fields $\psi_j(x)$. The fermionic
3694: Hamiltonian is then the positive operator in $\cH_f$ given by
3695: $$
3696: H_f=\,\int_{S^1}\,:\,\bar
3697: \psi\,\gamma^1\,i\,\partial\,\psi-\,m\,\bar \psi\,\psi \,:\,
3698: $$
3699: The full Hamiltonian of the non-interacting theory is acting in the
3700: Hilbert space $\cH=\,\cH_b \otimes \cH_f$ and is the positive
3701: operator
3702: $$
3703: H=\,H_b\otimes 1 +\,1\otimes H_f \,.
3704: $$
3705: This is were supersymmetry enters the scene in finding a
3706: self-adjoint square root of $H$ in the same way as the Dirac
3707: operator is a square root of the Laplacian in the case of finite
3708: dimensional manifolds. This square root, called the {\em
3709: supercharge} operator, is given by
3710: $$
3711: Q=\,\frac{1}{\sqrt{2}}\,\int_{S^1}(
3712: \psi_1(x)(\pi(x)-\partial\phi^*(x)-i m \phi(x))+
3713: \psi_2(x)(\pi^*(x)-\partial\phi(x)-i m \phi^*(x)) + {\rm h.c.})dx
3714: $$
3715: where the symbol $+{\rm h.c.}$ means that one adds the hermitian
3716: conjugate.
3717:
3718: The basic relation with spectral triples is then given by the
3719: following result (\cite{Co94} Section IV).
3720:
3721: \begin{thm} For any local region $\mathcal O\subset C$ let $\cA(\mathcal
3722: O)$ be the algebra of functions of quantum fields with support in
3723: $\mathcal O$ acting in the Hilbert space $\cH$. Then the triple
3724: $$
3725: (\cA(\mathcal O),\cH,Q)
3726: $$
3727: is an even $\theta$-summable spectral triple, with $\Z_2$-grading
3728: given by the operator $\gamma=\,(-1)^{N_f}$ counting the parity of
3729: the fermion number operator $N_f$.
3730: \end{thm}
3731:
3732:
3733: To be more specific the algebra $\cA(\mathcal O)$ is generated by
3734: the imaginary exponentials $e^{i(\phi(f)+\phi(f)^*)}$ and
3735: $e^{i(\pi(f)+\pi(f)^*)}$ for $f\in C_c^{\infty}(\mathcal O)$. As
3736: shown in \cite{Co94} Section IV.9.$\beta$, and exactly as in the
3737: case of discrete groups with exponential growth, one needs the
3738: entire cyclic cohomology rather than its finite dimensional version
3739: in order to obtain the Chern character of $\theta$-summable spectral
3740: triples. Indeed the index map is non polynomial in the above example
3741: of the Wess-Zumino model in two dimensions and the $K$-theory of the
3742: above local algebras is highly non-trivial. In fact it is in that
3743: framework that the JLO-cocycle was discovered by Jaffe-Lesniewski
3744: and Osterwalder \cite{JLO}.
3745:
3746:
3747: It is an open problem to extend the above result to interacting
3748: theories in higher dimension and give a full computation of the
3749: $K$-theory of the local algebras as well as of the Chern character
3750: in entire cyclic cohomology. The results of Jaffe and his
3751: collaborators on constructive quantum field theory yield many
3752: interacting non-trivial examples of supersymmetric two dimensional
3753: models. Moreover the recent breakthrough of Puschnigg in the case of
3754: lattices of semi-simple Lie groups of rank one opens the way to the
3755: computation of the Chern character in entire cyclic cohomology.
3756:
3757: \bigskip
3758:
3759: \section{Spacetime and the standard model of elementary
3760: particles}\label{stmod}
3761:
3762: The standard model of elementary particle physics provides a
3763: surprising example of a spectral triple in the noncommutative
3764: setting, which in addition to the conditions of Definition
3765: \ref{def-triple} also has a real structure satisfying all the
3766: additional conditions of Definition \ref{realstr}.
3767:
3768: The noncommutative geometry of the standard model developed in
3769: \cite{Co3} (\cf also \cite{C-C}, \cite{C-C2}, \cite{CL},
3770: \cite{Kast2}) gives a concise conceptual way to describe, through
3771: a simple mathematical structure, the full complexity of the input
3772: from physics. As we recall here, the model also allows for
3773: predictions.
3774:
3775: The physics of the standard model can be described by a
3776: Lagrangian. We consider here the standard model minimally coupled
3777: to gravity, so that the Lagrangian we shall be concerned with is
3778: the sum
3779: \begin{equation}\label{Lagr-EHSM}
3780: \cL = \cL_{EH} + \cL_{SM}
3781: \end{equation}
3782: of the Einstein--Hilbert Lagrangian $\cL_{EH}$ and the standard
3783: model Lagrangian $\cL_{SM}$.
3784:
3785: The standard model Lagrangian $\cL_{SM}$ has a very complicated
3786: expression, which, if written in full, might take a full page (\cf
3787: \eg \cite{Velt2}). It comprises five types of terms,
3788: \begin{equation}\label{Lagr-SM}
3789: \cL_{SM} = \cL_G + \cL_{GH} + \cL_H + \cL_{Gf} + \cL_{Hf},
3790: \end{equation}
3791: where the various terms involve:
3792: \begin{itemize}
3793: \item spin $1$ bosons $G$: the eight gluons, $\gamma$, $W^\pm$, $Z$;
3794: \item spin $0$ bosons $H$ such as the Higgs fields;
3795: \item spin $1/2$ fermions $f$: quarks and leptons.
3796: \end{itemize}
3797: The term $\cL_G$ is the pure gauge boson part, $\cL_{GH}$ for the
3798: minimal coupling with the Higgs fields, and $\cL_H$ gives the
3799: quartic Higgs self interaction. In addition to the coupling
3800: constants for the gauge fields, the fermion kinetic term
3801: $\cL_{Gf}$ contains the hypercharges $Y_L$, $Y_R$. These numbers,
3802: which are constant over generations, are assigned
3803: phenomenologically, so as to obtain the correct values of the
3804: electromagnetic charges. The term $\cL_{Hf}$ contains the Yukawa
3805: coupling of the Higgs fields with fermions. A more detailed and
3806: explicit description of the various terms of \eqref{Lagr-SM} is
3807: given in \cite{Co94} \S VI.5.$\beta$. See also \cite{Velt2}.
3808:
3809: \begin{figure}
3810: \begin{center}
3811: \includegraphics[scale=0.65]{standardmodel.eps}
3812: \end{center}
3813: \caption{Elementary particles \label{FigSM}}
3814: \end{figure}
3815:
3816:
3817: The symmetry group of the Einstein--Hilbert Lagrangian $\cL_{EH}$
3818: by itself would be, by the equivalence principle, the
3819: diffeomorphism group ${\rm Diff}(X)$ of the space-time manifold.
3820: In the standard model Lagrangian $\cL_{SM}$, on the other hand,
3821: the gauge theory has another huge symmetry group which is the
3822: group of local gauge transformations. According to our current
3823: understanding of elementary particle physics, this is given by
3824: \begin{equation}\label{gauge-gr}
3825: G_{SM}(X)= {\rm C}^\infty(X,U(1)\times SU(2)\times SU(3)).
3826: \end{equation}
3827: (At least in the case of a trivial principal bundle, \eg when the
3828: spacetime manifold $X$ is contractible.)
3829:
3830: Thus, when one considers the Lagrangian $\cL$ of
3831: \eqref{Lagr-EHSM}, the full symmetry group $G$ will be a
3832: semidirect product
3833: \begin{equation}\label{symm-gr-SM}
3834: G(X) = G_{SM}(X) \rtimes {\rm Diff}(X).
3835: \end{equation}
3836: In fact, a diffeomorphism of the manifold relabels the gauge
3837: parameters.
3838:
3839: To achieve a {\em geometrization} of the standard model, one would
3840: like to be able to exhibit a space $X$ for which
3841: \begin{equation}\label{diffG}
3842: G(X)={\rm Diff}(X).
3843: \end{equation}
3844: If such a space existed, then we would be able to say that the whole
3845: theory is pure gravity on $X$. However, it is impossible to find
3846: such a space $X$ among ordinary manifolds. In fact, a result of W.
3847: Thurston, D. Epstein and J. Mather (\cf \cite{mather}) shows that
3848: the connected component of the identity in the diffeomorphism group
3849: of a (connected) manifold is a simple group (see \cite{mather} for
3850: the precise statement). A simple group cannot have a nontrivial
3851: normal subgroup, so it cannot have the structure of semi-direct
3852: product like $G(X)$ in \eqref{symm-gr-SM}.
3853:
3854: However, it is possible to obtain a space with the desired
3855: properties among {\em noncommutative spaces}. What plays the role of
3856: the connected component of the identity in the diffeomorphism group
3857: ${\rm Diff}(X)$ in the noncommutative setting is the group
3858: $\Aut^+(\cA)$ of automorphisms of the (noncommutative) algebra that
3859: preserve the fundamental class in $K$-homology \ie that can be
3860: implemented by a unitary compatible with the grading and real
3861: structure.
3862:
3863: When the algebra $\cA$ is not commutative, among its automorphisms
3864: there are, in particular, inner ones. They associate to an element
3865: $x$ of the algebra the element $uxu^{-1}$, for some $u\in \cA$. Of
3866: course $uxu^{-1}$ is not, in general, equal to $x$ because the
3867: algebra is not commutative. The inner automorphisms form a normal
3868: subgroup of the group of automorphisms. Thus, we see that the
3869: group $\Aut^+ (\cA)$ has in general the same type of structure as
3870: our desired group of symmetries $G(X)$, namely, it has a normal
3871: subgroup of inner automorphisms and it has a quotient. It is
3872: amusing how the physical and the mathematical vocabularies agree
3873: here: in physics one talks about internal symmetries and in
3874: mathematics one talks about inner automorphisms (one might as well
3875: call them internal automorphisms).
3876:
3877: There is a very simple non commutative algebra $\cA$ whose group
3878: of inner automorphisms corresponds to the group of gauge
3879: transformations $G_{SM}(X)$, and such that the quotient $\Aut^+
3880: (\cA) / {\rm Inn} (\cA)$ corresponds exactly to diffeomorphisms
3881: (\cf \cite{Schu}). The noncommutative space is a product $X\times
3882: F$ of an ordinary spacetime manifold $X$ by a ``finite
3883: noncommutative space'' $F$. The noncommutative algebra $\cA_F$ is
3884: a direct sum of the algebras $\bC$, $\bH$ (here denoting the
3885: quaternions), and $M_3 (\bC)$ (the algebra of $3\times 3$ complex
3886: matrices).
3887:
3888: The algebra $\cA_F$ corresponds to a {\it finite} space where the
3889: standard model fermions and the Yukawa parameters (masses of
3890: fermions and mixing matrix of Kobayashi Maskawa) determine the
3891: spectral geometry in the following manner. The Hilbert space
3892: $\cH_F$ is finite-dimensional and admits the set of elementary
3893: fermions as a basis. This comprises the generations of quarks
3894: (down--up, strange--charmed, bottom--top),
3895: \begin{equation}
3896: \begin{array}{ccccccccc}
3897: u_L & u_R & d_L & d_R& & \bar{u}_L & \bar{u}_R & \bar{d}_L & \bar{d}_R \\[2mm]
3898: c_L & c_R & s_L & s_R& & \bar{c}_L & \bar{c}_R & \bar{s}_L & \bar{s}_R \\[2mm]
3899: t_L & t_R & b_L & b_R& & \bar{t}_L & \bar{t}_R & \bar{b}_L &
3900: \bar{b}_R,
3901: \end{array}
3902: \label{quarks}
3903: \end{equation}
3904: with the additional color index $(y,r,b)$, and the generations of
3905: leptons (electron, muon, tau, and corresponding neutrinos)
3906: \begin{equation}
3907: \begin{array}{ccccccc}
3908: e_L& e_R& \nu^e_L& & \bar e_L& \bar e_R& \bar \nu^e_L
3909: \\[2mm]
3910: \mu_L& \mu_R& \nu^\mu_L& & \bar \mu_L& \bar \mu_R& \bar \nu^\mu_L
3911: \\[2mm]
3912: \tau_L& \tau_R& \nu^\tau_L& & \bar \tau_L& \bar \tau_R &
3913: \bar\nu^\tau_L
3914: \end{array}
3915: \label{leptons}
3916: \end{equation}
3917: (We discuss here only the minimal standard model with no right
3918: handed neutrinos.)
3919:
3920: The $\Z/2$ grading $\gamma_F$ on the Hilbert space $\cH_F$ has
3921: sign $+1$ on left handed particles (\eg the $u_L$, $d_L$, etc.)
3922: and sign $-1$ on the right handed particles. The involution $J_F$
3923: giving the real structure is the charge conjugation, namely, if we
3924: write $\cH_F = \cE \oplus \bar \cE$, then $J_F$ acts on the
3925: fermion basis as $J_F(f,\bar h) = (h,\bar f)$. This satisfies
3926: $J_F^2=1$ and $J_F\gamma_F=\gamma_F J_F$, as should be for
3927: dimension $n=0$.
3928:
3929: The algebra $\cA_F$ admits a natural representation in $\cH_F$
3930: (see \cite{Coo2}). An element $(z,q,m)\in \C \oplus \H \oplus
3931: M_3(\C)$ acts as
3932: $$ (z,q,m) \cdot \left(\begin{array}{c} u_R \\ d_R \end{array}\right)
3933: = \left(\begin{array}{c} z\, u_R \\ \bar z\, d_R
3934: \end{array}\right) \ \ \ \ (z,q,m) \cdot e_R = \bar z\, e_R $$
3935: $$ (z,q,m) \cdot \left(\begin{array}{c} u_L \\ d_L \end{array}\right)
3936: = q \left(\begin{array}{c} u_L \\ d_L \end{array}\right) \ \ \ \
3937: (z,q,m) \cdot \left(\begin{array}{c} \nu^e_L \\ e_L
3938: \end{array}\right) = q \left(\begin{array}{c} \nu^e_L \\ e_L
3939: \end{array}\right), $$
3940: $$ (z,q,m) \cdot \left(\begin{array}{c} \bar e_L \\ \bar e_R
3941: \end{array}\right) = \left(\begin{array}{c} z\, \bar e_L \\ z\, \bar e_R
3942: \end{array}\right) $$
3943: $$ (z,q,m) \cdot \bar u_R = m\, \bar u_R \ \ \ \ (z,q,m) \cdot \bar
3944: d_R =m \, \bar d_R $$ and similarly for the other generations.
3945: Here $q\in \H$ acts as multiplication by the matrix
3946: $$ q= \left(\begin{array}{cc} \alpha & \beta \\
3947: -\bar \beta & \bar \alpha \end{array}\right), $$ where $q=\alpha+
3948: \beta\, j$, with $\alpha,\beta\in \C$. The matrix $m\in M_3(\C)$
3949: acts on the color indices $(y,r,b)$.
3950:
3951: The data $(\cA_F, \cH_F)$ can be completed to a spectral triple
3952: $(\cA_F,\cH_F,D_F)$ where the Dirac operator (in this finite
3953: dimensional case a matrix) is given by
3954: \begin{equation}\label{YukawaD}
3955: D_F = \left( \begin{array}{cc} Y & 0 \\ 0 & \bar Y
3956: \end{array}\right)
3957: \end{equation}
3958: on $\cH_F=\cE\oplus \bar \cE$, where $Y$ is the Yukawa coupling
3959: matrix, which combines the masses of the elementary fermions
3960: together with the Cabibbo--Kobayashi--Maskawa (CKM) quark mixing matrix.
3961:
3962: The fermionic fields acquire mass through the spontaneous symmetry
3963: breaking produced by the Higgs fields. The Yukawa coupling matrix
3964: takes the form $Y=Y_q \otimes 1 \oplus Y_f$, where the matrix
3965: $Y_f$ is of the form
3966: $$ \left(\begin{array}{ccc} 0&0& M_e\\
3967: 0&0&0\\ M_e^* &0&0 \end{array}\right), $$ in the basis
3968: $(e_R,\nu_L,e_L)$ and successive generations, while $Y_q$ is of
3969: the form
3970: $$ \left(\begin{array}{cccc} 0&0& M_u & 0\\
3971: 0&0&0& M_d \\M_u^* &0&0&0 \\ 0& M_d^* & 0&0
3972: \end{array}\right), $$
3973: in the basis given by $(u_R,d_R,u_L,d_L)$ and successive
3974: generations. In the case of the lepton masses, up to rotating the
3975: fields to mass eigenstates, one obtains a mass term for each
3976: fermion, and the off diagonal terms in $M_e$ can be reabsorbed in
3977: the definition of the fields. In the quark case, the situation is
3978: more complicated and the Yukawa coupling matrix can be reduced to
3979: the mass eigenvalues and the CKM quark mixing. By rotating the
3980: fields, it is possible to eliminate the off diagonal terms in
3981: $M_u$. Then $M_d$ satisfies $V M_d V^*=M_u$, where $V$ is the CKM
3982: quark mixing, given by a $3\times 3$ unitary matrix
3983: $$ V= \left( \begin{array}{ccc} V_{ud} & V_{us} & V_{ub} \\
3984: V_{cd} & V_{cs} & V_{cb} \\
3985: V_{td} & V_{ts} & V_{tb} \end{array} \right) $$ acting on the
3986: charge $-e/3$ quarks (down, strange, bottom). The entries of this
3987: matrix can be expressed in terms of three angles
3988: $\theta_{12},\theta_{23},\theta_{13}$ and a phase, and can be
3989: determined experimentally from weak decays and deep inelastic
3990: neutrino scatterings.
3991:
3992: The detailed structure of the Yukawa coupling matrix $Y$ (in
3993: particular the fact that color is not broken) allows one to check
3994: that the finite geometry $(\cA_F,\cH_F,D_F)$ satisfies all the
3995: axioms of Definition \ref{realstr} for a noncommutative spectral
3996: manifold. The key point is that elements $a \in \cA_F$ and
3997: $[D_f,a]$ commute with $J_F\, \cA_F\, J_F$. These operators
3998: preserve the subspace $\cE \subset \cH_F$. On this subspace, for
3999: $b=(z,q,m)$, the action of $J_F \, b^* \, J_F$ is by
4000: multiplication by $z$ or by the transpose $m^t$. It is then not
4001: hard to check explicitly the commutation with $a$ or $[D,a]$ (\cf
4002: \cite{Co94} \S VI.5.$\delta$). By exchanging the roles of $a$ and
4003: $b$, one sees analogously that $a$ commutes with $J_F b J_F$ and
4004: $[D,J_F b J_F]$ on $\bar \cE$, hence the desired commutation
4005: relations hold on all of $\cH_F$.
4006:
4007: We can then consider the product $X\times F$, where $X$ is an
4008: ordinary $4$--dimensional Riemannian spin manifold and $F$ is the
4009: finite geometry described above. This product geometry is a
4010: spectral triple $(\cA,\cH,D)$ obtained as the cup product of a
4011: triple $(\cC^\infty(X),L^2(X,S), D_1)$, where $D_1$ is the Dirac
4012: operator on $X$ acting on square integrable spinors in $L^2(X,S)$,
4013: with the spectral triple $(\cA_F,\cH_F,D_F)$ described above.
4014: Namely, the resulting (smooth) algebra and Hilbert space are of
4015: the form
4016: \begin{equation}\label{algXF}
4017: \cA_\infty=\cC^\infty(X,\cA_F) \ \ \ \ \cH= L^2(X,S)\otimes
4018: \cH_F,
4019: \end{equation}
4020: and the Dirac operator is given by
4021: \begin{equation}\label{diracXF}
4022: D = D_1 \otimes 1 + \gamma \otimes D_F,
4023: \end{equation}
4024: where $\gamma$ is the usual $\Z/2$ grading on the spinor bundle
4025: $S$. The induced $\Z/2$ grading on $\cH$ is the tensor product
4026: $\gamma\otimes \gamma_F$, and the real structure is given by
4027: $J=C\otimes J_F$, where $C$ is the charge conjugation operator on
4028: spinors.
4029:
4030: Notice that, so far, we have only used the information on the
4031: fermions of the standard model. We'll see now that the bosons,
4032: with the correct quantum numbers, are {\em deduced} as inner
4033: fluctuations of the metric of the spectral triple $(\cA,\cH,D)$.
4034:
4035: It is a general fact that, for noncommutative geometries
4036: $(\cA,\cH,D)$, one can consider inner fluctuations of the metric
4037: of the form
4038: $$
4039: D\mapsto D+A+JAJ^{-1}
4040: $$
4041: where $A$ is of the form
4042: \begin{equation}\label{Aab}
4043: A= \sum a_i \, [D, a_i' ] \ \ \ \ a_i,a_i' \in \cA.
4044: \end{equation}
4045:
4046: In the case of the standard model, a direct computation of the
4047: inner fluctuations gives the standard model gauge bosons $\gamma ,
4048: W^{\pm} ,Z$, the eight gluons and the Higgs fields $\varphi$ with
4049: accurate quantum numbers (\cf \cite{Co3}). In fact, a field $A$ of
4050: the form \eqref{Aab} can be separated in a ``discrete part''
4051: $A^{(0,1)} = \sum a_i \, [\gamma\otimes D_F, a_i' ]$ and a
4052: continuous part $A^{(1,0)} = \sum a_i \, [D_1\otimes 1, a_i' ]$,
4053: with $a_i=(z_i,q_i,m_i)$ and $a_i'=(z_i',q_i',m_i')$,
4054: $q_i=\alpha_i + \beta_i j$ and $q_i'=\alpha_i' + \beta_i' j$. The
4055: discrete part gives a quaternion valued function
4056: $$ q(x) = \sum z_i \left( (\alpha_i'-z_i') + z_i \beta_i'\, j \right)
4057: = \varphi_1 + \varphi_2 \, j $$ which provides the Higgs doublet.
4058: The continuous part gives three types of fields:
4059: \begin{itemize}
4060: \item A $U(1)$ gauge field $U =\sum z_i \, d z_i'$
4061: \item An $SU(2)$ gauge field $Q= \sum q_i \, d q_i'$
4062: \item A $U(3)$ gauge field $M=\sum m_i \, d m_i'$,
4063: which can be reduced to an $SU(3)$ gauge field $M'$ by subtracting
4064: the scalar part of the overall gauge field which eliminates
4065: inessential fluctuations that do not change the metric.
4066: \end{itemize}
4067: The resulting internal fluctuation of the metric $A+J A J^{-1}$ is
4068: then of the form (\cf \cite{Co3})
4069: $$ \left(\begin{array}{ccc} -2U & 0 & 0 \\
4070: 0 & Q_{11}-U & Q_{12} \\ 0& Q_{21} & Q_{22}-U
4071: \end{array} \right) $$ on the basis of leptons $(e_R,\nu_L,e_L)$
4072: and successive generations, and
4073: $$ \left(\begin{array}{cccc} \frac{4}{3} U + M' & 0 & 0 & 0 \\[2mm]
4074: 0 & \frac{-2}{3} U + M' & 0 & 0 \\[2mm]
4075: 0 & 0 & Q_{11} + \frac{1}{3} U + M' & Q_{12} \\[2mm]
4076: 0 & 0 & Q_{21} & Q_{22} + \frac{1}{3} U + M'
4077: \end{array}\right), $$
4078: on the basis of quarks given by $(u_R,d_R,u_L,d_L)$ and successive
4079: generations. A striking feature that these internal fluctuations
4080: exhibit is the fact that the expressions above recover all the
4081: exact values of the hypercharges $Y_L$, $Y_R$ that appear in the
4082: fermion kinetic term of the standard model Lagrangian.
4083:
4084: Finally, one can also recover the bosonic part of the standard model
4085: Lagragian from a very general principle, the {\em spectral action
4086: principle} of Chamseddine--Connes (\cf \cite{C-C}, \cite{C-C2},
4087: \cite{C-C3}). The result is that the Hilbert--Einstein action
4088: functional for the Riemannian metric, the Yang--Mills action for the
4089: vector potentials, and the self interaction and the minimal coupling
4090: for the Higgs fields all appear with the correct signs in the
4091: asymptotic expansion for large $\Lambda$ of the number $N(\Lambda)$
4092: of eigenvalues of $D$ which are $\leq \Lambda$ (\cf \cite{C-C}),
4093: \begin{equation}
4094: N(\Lambda) = \# \ \hbox{eigenvalues of $D$ in} \
4095: [-\Lambda,\Lambda].
4096: \end{equation}
4097:
4098: The spectral action principle, applied to a spectral triple
4099: $(\cA,\cH,D)$, can be stated as saying that the physical action
4100: depends only on $\Sp(D)\subset \R$. This spectral datum corresponds
4101: to the data $(\cH,D)$ of the spectral triple, independent of the
4102: action of $\cA$. Different $\cA$ that correspond to the same
4103: spectral data can be thought of as the noncommutative analog of
4104: isospectral Riemannian manifolds (\cf the discussion of isospectral
4105: deformations in Section \ref{isodef}). A natural expression for an
4106: action that depends only on $\Sp(D)$ and is {\em additive} for
4107: direct sums of spaces is of the form
4108: \begin{equation}\label{Sp-action}
4109: \Tr\, \chi (\frac{D}{\Lambda} )+ \langle \psi, D \, \psi \rangle,
4110: \end{equation}
4111: where $\chi$ is a positive even function and $\Lambda$ is a scale.
4112:
4113:
4114: In the case of the standard model, this formula \eqref{Sp-action} is
4115: applied to the full ``metric" including the internal fluctuations
4116: and gives the full standard model action minimally coupled with
4117: gravity. The Fermionic part of the action \eqref{Sp-action} gives
4118: (\cf \cite{C-C}, \cite{C-C2})
4119: \begin{equation}\label{fermionic}
4120: \langle \psi, D \, \psi \rangle= \int_X \left( \cL_{Gf} + \cL_{Hf}
4121: \right) \sqrt{|g|} d^4 x .
4122: \end{equation}
4123:
4124: The bosonic part of the action \eqref{Sp-action} evaluated via heat
4125: kernel invariants gives the standard model Lagrangian minimally
4126: coupled with gravity. Namely, one writes the function $\chi
4127: (\frac{D}{\Lambda})$ as the superposition of exponentials. One then
4128: computes the trace by a semiclassical approximation from local
4129: expressions involving the familiar heat equation expansion. This
4130: delivers all the correct terms in the action (\cf \cite{C-C2} for an
4131: explict calculation of all the terms involved).
4132:
4133: Notice that here one treats the spacetime manifold $X$ in the
4134: Euclidean signature. The formalism of spectral triple can be
4135: extended in various ways to the Lorentzian signature (\cf \eg
4136: \cite{EHaw}). Perhaps the most convenient choice is to drop the
4137: self-adjointness condition for $D$ while still requiring $D^2$ to
4138: be self-adjoint.
4139:
4140: \bigskip
4141:
4142:
4143: \section{Isospectral deformations}\label{isodef}
4144:
4145: A very rich class of examples of noncommutative manifolds is
4146: obtained by considering isospectral deformations of a classical
4147: Riemannian manifold. These examples satisfy all the axioms of
4148: ordinary Riemannian geometry (\cf \cite{Co3}) except
4149: commutativity. They are obtained by the following result
4150: (Connes--Landi \cite{CLa}):
4151:
4152: \begin{thm}\label{Theorem6}
4153: Let $M$ be a compact Riemannian spin manifold. Then if the
4154: isometry group of $M$ has rank $r \geq 2$, $M$ admits a
4155: non-trivial one parameter isospectral deformation to
4156: noncommutative geometries $M_{\theta}$.
4157: \end{thm}
4158:
4159: \smallskip
4160:
4161: The main idea of the construction is to deform the standard
4162: spectral triple describing the Riemannian geometry along a two
4163: torus embedded in the isometry group, to a family of spectral
4164: triples describing non-commutative geometries.
4165:
4166: \smallskip
4167:
4168: More precisely, under the assumption on the rank of the group of
4169: isometries of the compact spin manifold $X$, there exists a
4170: two-torus
4171: $$ T^2 \subset {\rm Isom}(X), $$
4172: where we identify $T^2 = \R^2 / (2\pi \Z)^2$. Let $U(s)$ be the
4173: unitary operators in this subgroup of isometries, for
4174: $s=(s_1,s_2)\in T^2$, acting on the Hilbert space $\cH = L^2(X,S)$
4175: of the spectral triple $$(C^\infty(X),L^2(X,S),D,J).$$
4176: Equivalently, we write $U(s)= \exp (i(s_1 P_1 + s_2 P_2))$, where
4177: $P_i$ are the corresponding Lie algebra generators, with ${\rm
4178: Spec}(P_i)\subset \Z$, satisfying $[D,P_i]=0$ and $P_i J = - J
4179: P_i$, so that $[U(s),D]=[U(s),J]=0$.
4180:
4181: \smallskip
4182:
4183: The action $\alpha_s(T)=U(s) T U(s)^{-1}$ has the following
4184: property. Any operator $T$ such that the map $s \mapsto
4185: \alpha_s(T)$ is smooth can be uniquely written as a norm
4186: convergent series
4187: \begin{equation}\label{T-bidegree}
4188: T = \sum_{n_1, n_2 \in \Z} \hat T_{n_1,n_2}
4189: \end{equation}
4190: where each term $\hat T_{n_1,n_2}$ is an operator of bi-degree
4191: $(n_1,n_2)$, that is,
4192: $$ \alpha_s(\hat T_{n_1,n_2})= \exp(i(s_1 n_1 + s_2 n_2)) \hat
4193: T_{n_1,n_2}, $$ for each $s=(s_1,s_2)\in T^2$, and the sequence of
4194: norms $\| \hat T_{n_1,n_2} \|$ is of rapid decay.
4195:
4196: \smallskip
4197:
4198: This property makes it possible to define left and right twists
4199: for such operators $T$, defined as
4200: \begin{equation}\label{T-left-twist}
4201: \ell (T):= \sum_{n_1,n_2} \hat T_{n_1,n_2} \exp\left( 2 \pi i
4202: \theta n_2 P_1 \right)
4203: \end{equation}
4204: and
4205: \begin{equation}\label{T-right-twist}
4206: r (T):=\sum_{n_1,n_2} \hat T_{n_1,n_2} \exp\left( 2 \pi i \theta
4207: n_1 P_2 \right).
4208: \end{equation}
4209: Both series still converge in norm, since the $P_i$ are
4210: self-adjoint operators.
4211:
4212: \smallskip
4213:
4214: It is then possible to introduce a (left) deformed product
4215: \begin{equation}\label{left-twist-prod}
4216: x * y = \exp ( 2 \pi i \theta n_1 ' n_2 ) xy,
4217: \end{equation}
4218: for $x$ a homogeneous operator of bi-degree $(n_1, n_2)$ and $y$ a
4219: homogeneous operator of bi-degree $(n_1', n_2')$. A (right)
4220: deformed product is similarly defined by setting $x *_r y = \exp (
4221: 2 \pi i \theta n_1 n_2' ) xy$. These deformed products satisfy
4222: $\ell(x)\ell(y)= x* y$ and $r(x)r(y)=x *_r y$.
4223:
4224: \smallskip
4225:
4226: The deformed spectral triples are then obtained by maintaining the
4227: same Hilbert space $\cH = L^2(X,S)$ and Dirac operator $D$, while
4228: modifying the algebra $C^\infty(X)$ to the non-commutative algebra
4229: $\cA_\theta :=\ell(C^\infty(X))$ and the involution $J$ that
4230: defines the real structure to $J_\theta := \exp(2\pi i \theta
4231: P_1P_2) J$.
4232:
4233:
4234:
4235: \bigskip
4236:
4237: \section{Algebraic deformations}\label{deform}
4238:
4239: There is a very general context in which one constructs noncommutative
4240: spaces via deformations of commutative algebras. Unlike the
4241: isospectral deformations discussed in Section \ref{isodef}, here one
4242: proceeds mostly at a formal algebraic level, without involving the
4243: operator algebra structure and without invoking the presence of a
4244: Riemannian structure.
4245:
4246: The idea of deformation quantization originates in the idea that
4247: classical mechanics has as setting a smooth manifold (phase space)
4248: with a symplectic structure, which defines a Poisson bracket
4249: $\{,\}$. The system is quantized by deforming the pointwise
4250: product in the algebra $\cA=\cC^\infty(M)$ (or in a suitable
4251: subalgebra) to a family $*_\hbar$ of products satisfying $f
4252: *_\hbar g \to fg$ as $\hbar\to 0$, which are associative but no
4253: longer necessarily commutative. These are also required to satisfy
4254: $$ \frac{f *_\hbar g - g*_\hbar f}{i\hbar} \to \{ f, g \}, $$
4255: as $\hbar\to 0$, namely, the ordinary product is deformed in the
4256: direction of the Poisson bracket. On the algebra $\cC^\infty(M)$ a
4257: Poisson bracket is specified by assigning a section $\Lambda$ of
4258: $\Lambda^2(TM)$ with the property that
4259: $$ \{ f, g \} = \langle \Lambda, df\wedge dg \rangle $$
4260: satisfies the Jacobi identity. Typically, this produces a {\it
4261: formal deformation}: a formal power series in $\hbar$. Namely, the
4262: deformed product can be written in terms of a sequence of
4263: bi--differential operators $B_k$ satisfying
4264: \begin{equation}
4265: f*g = fg + \hbar B_1(f,g) + \hbar^2 B_2(f,g) + \cdots
4266: \label{Poisson-deform}
4267: \end{equation}
4268:
4269: \smallskip
4270:
4271: Under this perspective, there is a good understanding of formal
4272: deformations. For instance, Kontsevich \cite{KontIHES97} proved
4273: that formal deformations always exist, by providing an explicit
4274: combinatorial formula that generates all the $\{ B_2, B_3, \ldots
4275: \}$ in the expansion from the $B_1$, hence in terms of the Poisson
4276: structure $\Lambda$. The formal solution (\ref{Poisson-deform})
4277: can then be written as
4278: $$ \sum_{n=0}^\infty \hbar^n \sum_{\Gamma\in G[n]} \omega_\Gamma
4279: B_{\Gamma,\Lambda}(f,g), $$ where $G[n]$ is a set of $(n(n+1))^n$
4280: labeled graphs with $n+2$ vertices and $n$ edges, $\omega_\Gamma$
4281: is a coeffcient obtained by integrating a differential form
4282: (depending on the graph $\Gamma$) on the configuration space of
4283: $n$ distinct points in the upper half plane, and
4284: $B_{\Gamma,\Lambda}$ is a bi--differential operator whose
4285: coefficients are derivatives of $\Lambda$ of orders specified by
4286: the combinatorial information of the graph $\Gamma$.
4287:
4288: \smallskip
4289:
4290: A setting of deformation quantization which is compatible with
4291: $C^*$--algebras was developed by Rieffel in \cite{Rieffel89}. We
4292: recall briefly Rieffel's setting. For simplicity, we restrict to
4293: the simpler case of a compact manifold.
4294:
4295: \begin{defn}\label{Rieffel-def}
4296: A strict (Rieffel) deformation quantization of $\cA=\cC^\infty(M)$
4297: is obtained by assigning an associative product $*_\hbar$, an
4298: involution (depending on $\hbar$) and a $C^*$--norm $\| \cdot
4299: \|_{\hbar}$ on $\cA$, for $\hbar \in I$ (some interval containing
4300: zero), such that:
4301:
4302: (i) For $\hbar=0$ these give the $C^*$--algebra $C(M)$,
4303:
4304: (ii) For all $f,g\in \cA$, as $\hbar \to 0$,
4305: $$ \left\| \frac{f *_\hbar g - g*_\hbar f}{i\hbar} - \{ f, g \}
4306: \right\|_{\hbar} \to 0. $$
4307:
4308: One denotes by $\cA_\hbar$ the $C^*$--algebra obtained by
4309: completing $\cA$ in the norm $\| \cdot \|_{\hbar}$.
4310: \end{defn}
4311:
4312: The functions of $\hbar$ are all supposed to be analytic, so that
4313: formal power series expansions make sense.
4314:
4315: \begin{rem}{\em The notion of a strict deformation quantization should be
4316: regarded as a notion of integrability for a formal solutions. }
4317: \end{rem}
4318:
4319: \smallskip
4320:
4321: Rieffel also provides a setting for compatible actions by a Lie
4322: group of symmetries, and proves that non--commutative tori (also
4323: of higher rank) are strict deformation quantizations of ordinary
4324: tori, that are compatible with the action of the ordinary torus as
4325: group of symmetry. Typically, for a given Poisson structure,
4326: strict deformation quantizations are not unique. This happens
4327: already in the case of tori.
4328:
4329: \smallskip
4330:
4331: In the same paper \cite{Rieffel89}, Rieffel uses a basic result of
4332: Wassermann \cite{Wassermann88} to produce an example where formal
4333: solutions are not integrable. The example is provided by the
4334: two--sphere $S^2$. There is on $S^2$ a symplectic structure, and a
4335: corresponding Poisson structure $\Lambda$ which is invariant under
4336: $SO(3)$. Rieffel proves the following striking result (Theorem 7.1
4337: of \cite{Rieffel89}):
4338:
4339: \begin{thm} There are no $SO(3)$--invariant strict deformations of the
4340: ordinary product on $\cC^\infty(S^2)$ in the direction of the
4341: $SO(3)$--invariant Poisson structure.
4342: \end{thm}
4343:
4344: In fact, the proof of this result shows more, namely that no
4345: $SO(3)$--invariant deformation of the ordinary product in $C(S^2)$
4346: can produce a non--commutative $C^*$--algebra. This rigidity
4347: result reflects a strong rigidity result for $SU(2)$ proved by
4348: Wassermann \cite{Wassermann88}, namely the only ergodic actions of
4349: $SU(2)$ are on von Neumann algebras of type I. The interest of
4350: this result lies in the fact that there are formal deformations of
4351: the Poisson structure that are $SO(3)$--invariant (see \eg
4352: \cite{GuVai}, \cite{BayFron}), but these only exist as a formal
4353: power series in the sense of (\ref{Poisson-deform}) and, by the
4354: results of Wassermann and Rieffel are not integrable.
4355:
4356: \smallskip
4357:
4358: Summarizing, we have the following type of phenomenon: on the one
4359: hand we have formal solutions, formal deformation quantizations
4360: about which a lot is known, but for which, in general, there may
4361: not be an integrability result. More precisely, when we try to
4362: pass from formal to actual solutions, there are cases where
4363: existence fails (the sphere), and others (tori) where uniqueness
4364: fails. The picture that emerges is remarkably similar to the case
4365: of formal and actual solutions of ordinary differential equations.
4366:
4367: \smallskip
4368:
4369: It is very instructive to build an analogy between the {\it problem
4370: of ambiguity} for formal solutions of ODE's and the present
4371: situation of formal non--commutative spaces and actual
4372: non--commutative spaces. The main conclusion to be drawn from this
4373: analogy is that there ought to be a {\it theory of ambiguity} which
4374: formulates precisely the relation between the formal
4375: non--commutative geometry and its integrated ($C^*$--algebraic)
4376: version.
4377:
4378: \smallskip
4379:
4380: To illustrate this concept, we take a closer look at the analogous
4381: story in the theory of ODE's. A good reference for a modern
4382: viewpoint is \cite{Ramis}. A formal solution of a differential
4383: equation is a power series expansion: for instance
4384: $\sum_{n=0}^\infty (-1)^n n! x^{n+1}$ is a formal solution of the
4385: Euler equation $x^2 y' +y=x$. Convergent series give rise to actual
4386: solutions, and more involved summation processes such as Borel
4387: summation can be used to transform a given formal solution of an
4388: analytic ODE into an actual solution on a sufficiently narrow sector
4389: in $\C$ of sufficiently small radius, but such solution is in
4390: general {\it not unique}. It is known from several classical methods
4391: that some divergent series can be ``summed'' modulo a function with
4392: exponential decrease of a certain order. This property (Gevrey
4393: summability) is also satisfied by formal solutions of analytic
4394: ODE's, and, stated in a more geometric fashion, it is essentially a
4395: cohomological condition. It also shows that, whereas on {\it small
4396: sectors} one has existence of actual solutions but not uniqueness,
4397: on {\it large sectors} one gains uniqueness, at the cost of possibly
4398: loosing existence. A complete answer to summability of formal
4399: solutions can then be given in terms of a more refined
4400: multi--summability (combining Gevrey series and functions of
4401: different order) and the Newton polygon of the equation.
4402:
4403: \smallskip
4404:
4405: The general flavor of this theory is surprisingly similar to the
4406: problem of formal solutions in non--commutative geometry. It is to
4407: be expected that an ambiguity theorem exists, which accounts for
4408: the cases of lack of uniqueness, or lack of existence, of actual
4409: solutions illustrated by the results of Rieffel.
4410:
4411:
4412:
4413:
4414: \medskip
4415:
4416: Already in dealing with our first truly non--trivial example of
4417: noncommutative spaces, the noncommutative tori, we encountered
4418: subtleties related to the difference between the quotient and the
4419: deformation approach to the construction of non--commutative
4420: spaces.
4421:
4422: \smallskip
4423:
4424: In fact, the non--commutative tori we described in Section
4425: \ref{nctori} admit a description as algebras
4426: obtained as deformations of the ordinary product of functions, by
4427: setting
4428: \begin{equation}\label{NCTdeform}
4429: (f*g)(x,y):= \left( e^{2\pi i\theta \frac{\partial}{\partial x}
4430: \frac{\partial}{\partial y'} } f(x,y) g(x',y') \right)_{x=x',y=y'} =
4431: \sum \frac{(i2\pi \theta)^n}{n!} D_1^n f D_2^n g.
4432: \end{equation}
4433:
4434: \smallskip
4435:
4436: Notice however that while $U\,\frac{\partial}{\partial U}$ and
4437: $V\,\frac{\partial}{\partial V}$ are derivations
4438: for the algebra of the non--commutative torus, this is {\it not}
4439: the case for
4440: $\frac{\partial}{\partial U}$ and $\frac{\partial}{\partial V}$. The
4441: same holds for the quantum plane (\cf \cite{[M]}) whose algebra of
4442: coordinates admits two generators $u,v$ with relation
4443: $$
4444: u\,v=\,q\,v\,u\,.
4445: $$
4446: These generators can be rotated ($u\mapsto \lambda
4447: \,u$, $v\mapsto \mu \,v$) without affecting the presentation but
4448: translations of the generators are not automorphisms of the algebra.
4449: In other words, one can view the
4450: non--commutative torus as a deformation of an ordinary torus, which
4451: in turn is a quotient of the classical plane $\R^2$ by a lattice of
4452: translations, but the action of translations does not extend to the
4453: quantum plane. This is an instance of the fact that the general
4454: operations of quotient and deformation, in constructing
4455: non--commutative spaces, do not satisfy any simple compatibility
4456: rules and need to be manipulated with care.
4457:
4458: Moreover, phenomena like the Morita equivalence between, for
4459: instance $\theta$ and $1/{\theta}$, are not detectable in a purely
4460: deformation theoretic perturbative expansion like the one given by
4461: the Moyal product \eqref{NCTdeform}. They are non-perturbative and
4462: cannot be seen at the perturbative level of the star product.
4463:
4464: \medskip
4465:
4466: In this respect, a very interesting recent result is that of Gayral,
4467: Gracia-Bondia, Iochum, Sch\"ucker, and Varilly, \cite{GGISV}, where
4468: they consider a version of the structure of spectral triple for
4469: non-compact spaces. In that case, for instance, one no longer can
4470: expect the Dirac operator to have compact resolvent and one can only
4471: expect a local version to hold, \eg $a (D-i)^{-1}$ is compact for
4472: $a\in \cA$. Other properties of Definitions \ref{def-triple} and
4473: \ref{realstr} are easily adapted to a ``local version'' but become
4474: more difficult to check than in the compact case. They show that the
4475: Moyal product deformation of $\R^{2n}$ fits in the framework of
4476: spectral triples and provides an example of such non-unital spectral
4477: triples. Thus, it appears that the structure of noncommutative
4478: Riemannian geometry provided by spectral triples should adapt nicely
4479: to some classes of algebraic deformations.
4480:
4481: \smallskip
4482:
4483: It appears at first that spectral triples may not be the right type
4484: of structure to deal with noncommutative spaces associated to
4485: algebraic deformations, because it corresponds to a form of
4486: Riemannian geometry, while many such spaces originate from K\"ahler
4487: geometry. However, the K\"ahler structure can often be also encoded
4488: in the setting of spectral triple, for example by considering also a
4489: second Dirac operator, as in \cite{BKLR} or through the presence of
4490: a Lefschetz operator as in \cite{CM1}.
4491:
4492: \medskip
4493:
4494: Noncommutative spaces obtained as deformations of commutative
4495: algebras fit in the context of a well developed algebraic theory of
4496: noncommutative spaces (\cf \eg \cite{Kontsevich93}
4497: \cite{Kontsevich00} \cite{Manin01RM} \cite{Manin91NC} \cite{[M]}
4498: \cite{Rosenberg98} \cite{Rosenberg95}, \cite{Soibelman01}). This
4499: theory touches on a variety of subjects like quantum groups and the
4500: deformation approach to non--commutative spaces and is interestingly
4501: connected to the theory of mirror symmetry. However, it is often not
4502: clear how to integrate this approach with the functional analytic
4503: theory of non--commutative geometry briefly summarized in section
4504: \ref{road}. Only recently, several results confirmed the existence
4505: of a rich interplay between the algebraic and functional analytic
4506: aspects of noncommutative geometry, especially through the work of
4507: Connes and Dubois-Violette (\cf \cite{CDV1}, \cite{CDV2}, \cite{CDV3})
4508: and of Polishchuk (\cf \cite{Poli}). Also, the work of Chakraborty and Pal
4509: \cite{ChakraPal} and Connes \cite{Co-Qgr} and more recently of van
4510: Suijlekom, Dabrowski, Landi, Sitarz, and Varilly \cite{SDLSV},
4511: \cite{SDLSV2} showed that quantum groups fit very nicely within the
4512: framework of noncommutative geometry described by spectral triples,
4513: contrary to what was previously belived. Ultimately, successfully
4514: importing tools from the theory of operator algebras into the realm
4515: of algebraic geometry might well land within the framework of what
4516: Manin refers to as a ``second quantization of algebraic geometry''.
4517:
4518:
4519: \bigskip
4520:
4521: \section{Quantum groups}\label{qgroups}
4522:
4523:
4524: For a long time it was widely believed that quantum groups could not
4525: fit into the setting of noncommutative manifolds defined in terms of
4526: spectral geometry. On the contrary, recent work of Chakraborti and Pal
4527: showed in \cite{ChakraPal} that the quantum group
4528: $SU_q(2)$, for $0\leq q < 1$, admits a spectral triple with Dirac
4529: operator that is equivariant with respect to its own (co)action.
4530:
4531: \smallskip
4532:
4533: The algebra $\cA$ of functions on the quantum group $SU_q(2)$ is generated
4534: by two elements $\alpha$ and $\beta$ with the relations
4535: \begin{equation}\label{Qgr-alg}
4536: \begin{array}{c}
4537: \alpha^*\alpha +\beta^*\beta =1, \ \ \ \ \alpha\alpha^*+ q^2 \beta\beta^*
4538: =1, \\[2mm] \alpha\beta = q\beta\alpha, \ \ \ \alpha\beta^* =q
4539: \beta^*\alpha, \ \ \ \beta^*\beta =\beta\beta^*. \end{array}
4540: \end{equation}
4541:
4542: \smallskip
4543:
4544: By the representation theory of the quantum group $SU_q(2)$ (\cf
4545: \cite{KlimSchmu}) there exists a Hilbert space $\cH$ with orthonormal
4546: basis $e^{(n)}_{ij}$, $n\in \frac{1}{2} \N$, $i,j\in \{ -n,\ldots, n \}$, and
4547: a unitary representation
4548: \begin{equation}\label{Qgr-rep}
4549: \begin{array}{ll}
4550: \alpha\, e^{(n)}_{ij} = & a_+(n,i,j)\, e^{(n+1/2)}_{i-1/2, j-1/2} +
4551: a_-(n,i,j)\, e^{(n-1/2)}_{i-1/2, j-1/2} \\[4mm]
4552: \beta\, e^{(n)}_{ij} = & b_+(n,i,j)\, e^{(n+1/2)}_{i+1/2, j-1/2} +
4553: b_-(n,i,j)\, e^{(n-1/2)}_{i+1/2, j-1/2},
4554: \end{array}
4555: \end{equation}
4556: with coefficients
4557: $$ \begin{array}{ll}
4558: a_+(n,i,j) = & q^{2n+i+j+1}\, Q(2n-2j+2,2n-2i+2,4n+2,4n+4) \\[2mm]
4559: a_-(n,i,j) = & Q(2n+2j,2n+2i,4n,4n+2) \\[2mm]
4560: b_+(n,i,j) = & -q^{n+j} \, Q(2n-2j+2,2n+2i+2,4n+2,4n+4) \\[2mm]
4561: b_-(n,i,j) = & q^{n+i} \, Q(2n+2j,2n-2i,4n,4n+2),
4562: \end{array} $$
4563: where we use the notation
4564: $$ Q(n,m,k,r)=\frac{(1-q^{ n })^{1/2}(1-q^{ m })^{1/2}}{(1-q^{ k
4565: })^{1/2}(1-q^{ r })^{1/2}}. $$
4566:
4567: \smallskip
4568:
4569: Consider then, as in \cite{Co-Qgr}, the operator
4570: \begin{equation}\label{Qgr-Dirac}
4571: D\, e^{(n)}_{ij} = \left\{ \begin{array}{rr} -2n & n\neq i \\
4572: 2n & n=i. \end{array} \right.
4573: \end{equation}
4574: More generally, one can consider operators of the form
4575: $D\, e^{(n)}_{ij} = d(n,i)\, e^{(n)}_{ij}$, as in \cite{ChakraPal},
4576: with $d(n,i)$ satisfying the conditions
4577: $d(n+1/2,i+1/2)-d(n,i) =O(1)$ and $d(n+1/2,i-1/2)-d(n,i) = O(n+i+1)$.
4578: Then one has the following result (Chakraborti--Pal
4579: \cite{ChakraPal}):
4580:
4581: \begin{thm}\label{Qgr-Sp3}
4582: The data $(\cA,\cH,D)$ ad above define an $SU_q(2)$ equivariant odd
4583: 3--summable spectral triple.
4584: \end{thm}
4585:
4586: The equivariance condition means that there is an action on $\cH$ of
4587: the enveloping algebra $\cU=U_q(\SL(2))$, which commutes with the
4588: Dirac operator $D$. This is generated by operators
4589: $$ \begin{array}{rl} k\, e^{(n)}_{ij} =& q^j \, e^{(n)}_{ij} \\[3mm]
4590: e\, e^{(n)}_{ij} =& q^{-n+1/2} (1-q^{2(n+j+1)})^{1/2}
4591: (1-q^{2(n-j)})^{1/2} (1-q^2)^{-1} \, e^{(n)}_{ij+1}, \end{array} $$
4592: satisfying the relations
4593: $$ ke=qek, \,\,\,\, kf=q^{-1} fk, \,\,\,\,
4594: [e,f]=\frac{k^2-k^{-2}}{q-q^{-1}},
4595: $$
4596: with $f=e^*$, and with coproduct
4597: $$ \Delta(k)=k\otimes k, \,\,\,\, \Delta(e)=k^{-1}\otimes e + e\otimes
4598: k, \,\,\,\, \Delta(f) =k^{-1}\otimes f + f\otimes k. $$
4599:
4600: \smallskip
4601:
4602: It is interesting that, while the classical $SU(2)$ is of
4603: (topological and metric) dimension three, the topological dimension of
4604: the algebra $\cA$ of $SU_q(2)$ drops to one (\cf \cite{ChakraPal}),
4605: but the metric dimension of the spectral triple remains equal to
4606: three as in the classical case.
4607:
4608: \smallskip
4609:
4610: Chakraborti and Pal showed in \cite{ChakraPal} that the Chern
4611: character of the spectral triple is nontrivial. Moreover, Connes in
4612: \cite{Co-Qgr} gave an explicit formula for its local index cocycle,
4613: where a delicate calculation provides the cochain whose coboundary is
4614: the difference between the Chern character and the local version in
4615: terms of remainders in the rational approximation to the logarithmic
4616: derivative of the Dedekind eta function.
4617:
4618: \smallskip
4619:
4620: The local index formula is obtained by constructing a symbol map
4621: $$ \rho : \cB \to C^\infty(S^*_q), $$
4622: where the algebra $C^\infty(S^*_q)$ gives a noncommutative version of
4623: the cosphere bundle, with a restriction map
4624: $r:C^\infty(S^*_q) \to C^\infty(D_{q+}^2 \times D_{q-}^2)$ to the
4625: algebra of two noncommutative disks. Here
4626: $\cB$ is the algebra generated by the elements $\delta^k(a)$, $a\in
4627: \cA$, with $\delta(a) =[|D|,a]$. On the cosphere bundle there is a
4628: geodesic flow, induced by the group of automorphisms $a\mapsto e^{it
4629: |D|} \, a \, e^{-it |D|}$. Then $\rho(b)^0$ denotes the component of
4630: degree zero with respect to the grading induced by this flow.
4631:
4632: \smallskip
4633:
4634: The algebra $C^\infty(D^2_q)$ is an extension
4635: $$ 0 \to \cS \to C^\infty(D^2_q) \stackrel{\sigma}{\to} C^\infty(S^1)
4636: \to 0, $$
4637: where the ideal $\cS$ is the algebra of rapidly decaying
4638: matrices. There are linear
4639: functionals $\tau_0$ and $\tau_1$ on $C^\infty(D^2_q)$,
4640: $$ \tau_1(a) = \frac{1}{2\pi} \int_0^{2\pi} \sigma(a) d\theta, $$
4641: $$ \tau_0(a) = \lim_{N\to \infty} \sum_{k=0}^N \langle a\, \epsilon_k,
4642: \epsilon_k \rangle - \tau_1(a) N, $$
4643: where $\tau_0$ is defined in terms of the representation of
4644: $C^\infty(D^2_q)$ on the Hilbert space $\ell^2(\N)$ with o.n.~basis $\{
4645: \epsilon_k \}$.
4646:
4647: \smallskip
4648:
4649: Recall that (\cf \cite{Co94}) a cycle $(\Omega,d,\int)$ is a triple with
4650: where $(\Omega , d )$ is a graded differential algebra,
4651: and $\int : \Omega^n \to \C$ is a closed graded trace on $\Omega$.
4652: A cycle over an algebra $\cA$ is given by a cycle $(\Omega , d , \int
4653: )$ together with a homomorphism $\rho :
4654: \cA \to \Omega^0$.
4655:
4656:
4657: In the case of the algebra $\cA$ of $SU_q(2)$, a cycle $(\Omega,d,\int)$
4658: is obtained in \cite{Co-Qgr} by considering $\Omega^1
4659: = \cA \oplus \Omega^{(2)}(S^1)$, with $\Omega^{(2)}(S^1)$ the space of
4660: weight two differential forms $f(\theta) d\theta^2$, with the
4661: $\cA$--bimodule structure
4662: $$ a\, (\xi,f)=(a\xi,\sigma(a) f) \ \ \ \ (\xi,f)\,a = (\xi a,
4663: -i\sigma(\xi) \sigma(a)' + f \sigma(a)), $$
4664: with differential
4665: $$ da = \partial a+ \frac{1}{2} \sigma(a)'' d\theta^2, $$
4666: with $\partial$ the derivation $\partial = \partial_\beta -
4667: \partial_\alpha$, and
4668: $$ \int \, (\xi,f) = \tau(\xi) + \frac{1}{2\pi i} \int f\, d\theta, $$
4669: where $\tau(a) = \tau_0(r_-(a^{(0)}))$, with $a^{(0)}$ the component
4670: of degree zero for $\partial$ and $r_-$ the restriction to
4671: $C^\infty(D^2_{q-})$. This definition of the cycle corrects for the
4672: fact that $\tau$ itself (as well as $\tau_0$) fails to be a trace.
4673:
4674: \smallskip
4675:
4676: The following result then holds (Connes \cite{Co-Qgr}):
4677:
4678: \begin{thm}\label{loc-Qgr}
4679: \begin{enumerate}
4680: \item The spectral triple $(\cA,\cH,D)$ of Theorem \ref{Qgr-Sp3} has
4681: dimension spectrum $\Sigma=\{ 1,2,3 \}$.
4682: \item The residue formula for pseudodifferential operators $a\in
4683: \cB$ in terms of their symbol is given by
4684: $$ \begin{array}{ll} \cutint\,\, a \, |D|^{-3} = & (\tau_1\otimes \tau_1)
4685: (r\rho(a)^0) \\[2mm]
4686: \cutint\,\, a \, |D|^{-2} = & (\tau_1\otimes\tau_0+\tau_0\otimes
4687: \tau_1)(r\rho(a)^0) \\[2mm]
4688: \cutint\,\, a \, |D|^{-1} = & (\tau_0\otimes\tau_0)(r\rho(a)^0)
4689: \end{array} $$
4690: \item The character $\chi(a_0,a_1)= \int a_0 \, da_1$ of the cycle
4691: $(\Omega,d,\int)$ is equal to the cocycle
4692: $$ \psi_1(a_0,a_1)= 2 \cutint a_0 \delta (a_1) P |D|^{-1} - \cutint a_0
4693: \delta^2 (a_1) P |D|^{-1}, $$
4694: with $P=(1+F)/2$. The local index formula is given by
4695: $$ \varphi_{odd} = \psi_1 + (b+B) \varphi_{even}, $$
4696: where $\varphi$ is the local index cocycle.
4697: \item The character $\Tr(a_0 [F,a_1])$ differs from the local form
4698: $\psi_1$ by the coboundary $b\psi_0$, with $\psi_0(a) = 2 \Tr (a P
4699: |D|^{-s})_{s=0}$. This cochain is determined by the values
4700: $\psi_0((\beta^*\beta)^n)$, which are of the form
4701: $$ \psi_0((\beta^*\beta)^n) = q^{-2n}(q^2 R_n(q^2) - G(q^2)), $$
4702: where $G$ is the logarithmic derivative of the Dedekind eta function
4703: \begin{equation}\label{DedEta}
4704: \eta(q^2)=q^{1/12} \prod_{k=1}^\infty (1-q^{2k}),
4705: \end{equation}
4706: and the $R_n$ are rational functions with poles only at roots of
4707: unity.
4708: \end{enumerate}
4709: \end{thm}
4710:
4711: More recently, another important breakthrough in the relation between
4712: quantum groups and the formalism of spectral triples was obtained by
4713: Walter van Suijlekom, Ludwik Dabrowski, Giovanni
4714: Landi, Andrzej Sitarz, Joseph C. Varilly, in \cite{SDLSV} and
4715: \cite{SDLSV2}.
4716:
4717: They construct a $3^+$ summable spectral triple $(\cA,\cH,D)$,
4718: where $\cA$ is, as before, the algebra of coordinates of the
4719: quantum group $SU_q(2)$. The geometry in this case is an isospectral
4720: deformation of the classical case, in the sense that the Dirac
4721: operator is the same as the usual Dirac operator for the round metric
4722: on the ordinary 3-sphere $S^3$. Moreover, the spectral triple
4723: $(\cA,\cH,D)$ is especially nice, in as it is equivariant with
4724: respect to both left and right action of the Hopf algebra
4725: $\cU_q(su_q(2))$.
4726:
4727: The classical Dirac operator for the round metric on $S^3$ has
4728: spectrum $\Sigma=\Sigma_+ \cup \Sigma_-$ with $\Sigma_+=\{ (2j+3/2):
4729: j=0,1/2,1,3/2,\ldots \}$ with multiplicities $(2j+1)(2j+2)$ and
4730: $\Sigma_-=\{ -(2j+1/2): j=1/2,1,3/2,\ldots\}$ with multiplicities
4731: $2j(2j+1)$. The Hilbert space is obtained by taking $V\otimes \C^2$,
4732: where $V$ is the left regular representation of $\cA$. It is very
4733: important here to take $V\otimes \C^2$ instead of $\C^2\otimes V$.
4734: Not only the latter violates the equivariance condition, but it was
4735: shown by Ghoswami that it produces unbounded commutators $[D,a]$,
4736: hence one does not obtain a spectral triple in that way.
4737:
4738: The spectral triple contructed in \cite{SDLSV} and \cite{SDLSV2} has
4739: a real structure $J$ and the Dirac operator satisfies a weak form
4740: of the ``order one condition'' (\cf Section \ref{singconic} above).
4741: The local index formula of \cite{Co-Qgr} (\cf Theorem \ref{loc-Qgr}
4742: above) extends to the spectral triple of \cite{SDLSV}, as proved in
4743: \cite{SDLSV2} and the structures of the cotangent space and the
4744: geodesic flow are essentially the same.
4745:
4746:
4747:
4748:
4749:
4750: \bigskip
4751: \bigskip
4752:
4753: \section{Spherical manifolds}\label{ncspheres}
4754:
4755: The noncommutative spheres $S^3_\varphi\subset{\mathbb R}^4_\varphi$
4756: are obtained as solutions of a very simple problem, namely the vanishing
4757: of the first component of the Chern character of a unitary $U\in
4758: M_2(\cA)$ where $\cA$ is the algebra of functions on the sphere and
4759: the Chern character is taken in the cyclic homology (b,B) bicomplex.
4760: The origin of this problem is to quantize the volume form of a three
4761: manifold (\cf \cite{CDV1}). The solutions are parameterized by three
4762: angles $\varphi_k$, $k\in \{1,2,3\}$ and the corresponding algebras
4763: are obtained by imposing the ``unit sphere relation"
4764: \begin{equation}
4765: \sum \,x_\mu^2=\,1
4766: \end{equation}
4767: to the four generators $x_0, x_1, x_2, x_3$ of the quadratic algebra
4768: $C_{\mathrm{alg}}({\mathbb R}^4_\varphi)$
4769: with the six relations
4770:
4771: \smallskip
4772: \begin{equation}
4773: \label{pres1} \sin (\varphi_k) \, [x_0 , x_k]_+ = i\; \cos
4774: (\varphi_{\ell} - \varphi_m) \, [x_{\ell} , x_m]
4775: \end{equation}
4776: \begin{equation}
4777: \label{pres2} \cos (\varphi_k) \, [x_0 , x_k] = i \;\sin
4778: (\varphi_{\ell} - \varphi_m) \, [x_{\ell} , x_m]_+ \, ,
4779: \end{equation}
4780:
4781: \medskip
4782:
4783: where $ [a , b]_+=\,a\,b+\,b\,a$ is the anticommutator and by
4784: convention the indices $k,l,m\in \{1,2,3\}$ always appear in cyclic
4785: order.
4786:
4787: The analysis of these algebras is a special case of the general
4788: theory of central quadratic forms for quadratic algebras developed
4789: in \cite{CDV2}, \cite{CDV3} and which we briefly recall below.
4790:
4791: Let $\cala=A(V,R)=T(V)/(R)$ be a quadratic algebra where $V$ is the
4792: linear span of the generators and $(R)\subset T(V)$ the ideal
4793: generated by the relations. The geometric data $\{E\,,\,
4794: \sigma\,,\,\call\}$ is given by an algebraic variety $E$, a
4795: correspondence $\sigma$ on $E$ and a line bundle $\call$ over $E$.
4796: These data are defined so as to yield an homomorphism $h$ from
4797: $\cala$ to a crossed product algebra constructed from sections of
4798: powers of the line bundle $\call$ on the graphs of the iterations of
4799: the correspondence $\sigma$. This crossed product only involves the
4800: positive powers of the correspondence $\sigma$ and thus remains
4801: ``triangular" and far remote from the ``semi-simple" set-up of
4802: $C^*$-algebras.
4803:
4804: This morphism $h$ can be considerably refined using the notion of
4805: positive central quadratic form.
4806:
4807:
4808: \begin{defn} \label{cent}
4809: Let $Q \in S^2(V)$ be a symmetric bilinear form on $V^\ast$ and $C$
4810: a component of $E \times E$. We say that $Q$ is
4811: \underline{central} on $C$ iff for all ($Z,\,Z'$) in $C$ and
4812: $\omega\in R$ one has,
4813: \begin{equation} \label{defcentral}
4814: \omega(Z,Z')\, Q(\sigma(Z'),\sigma^{-1}(Z))+Q(Z,Z')\,
4815: \omega(\sigma(Z'),\sigma^{-1}(Z)) =0
4816: \end{equation}
4817: \end{defn}
4818:
4819:
4820: This makes it possible to construct purely algebraically a crossed product
4821: algebra and an homomorphism from $\cala=A(V,R)$ to this crossed
4822: product \cite{CDV2}, \cite{CDV3}.
4823: The relation with $C^*$-algebras arises from
4824: positive central quadratic forms which make sense on involutive
4825: quadratic algebras.
4826:
4827: Let $\cala=A(V,R)$ be an {\sl involutive} quadratic algebra \ie an
4828: algebra over $\C$ which is a
4829: $\ast$-algebra with involution $x\mapsto x^\ast$ preserving the subspace
4830: $V$ of the generators.
4831: The real structure of $V$ is given by the antilinear involution
4832: $v\mapsto j( v)$ restriction of $x\mapsto x^\ast$. As
4833: $(xy)^\ast=y^\ast x^\ast$ for $x,y\in \cala$, the space $R$ of
4834: relations fulfills
4835: \begin{equation}
4836: (j \otimes j)( R)=t(R) \label{eq5.3}
4837: \end{equation}
4838: in $V\otimes V$ where $t:V\otimes V\rightarrow V\otimes V$ is the
4839: transposition $v\otimes w \mapsto t(v\otimes w)=w\otimes v$. This
4840: implies that the characteristic variety is stable under the
4841: involution $j$ and one has
4842: $$
4843: \sigma( j(Z)) =\, j( \sigma^{-1}(Z))
4844: $$
4845:
4846:
4847:
4848: \smallskip
4849:
4850: Let then $C$ be an invariant component of $E \times E$, we say that
4851: $C$ is $j$-{\em real} when it is globally invariant under the
4852: involution
4853: \begin{equation}\label{tildej}
4854: \tilde j(Z,\,Z'):=( j( Z'),\, j( Z))
4855: \end{equation}
4856: Let then $Q$ be a central qudratic form on $C$, we say that $Q$ is
4857: positive on $C$ iff
4858:
4859: $$
4860: Q(Z,j(Z))> 0 \qqq Z\in K\,.
4861: $$
4862:
4863:
4864: \smallskip One can then endow the line bundle $\call$ dual of the tautological bundle
4865: on $P(V^*)$ with the hermitian metric given by
4866: \begin{equation}\label{herm}
4867: \langle f\,L,\,g\, L'\rangle_Q(Z) =\,f(Z)\,\overline{g(Z)} \,\frac{
4868: L(Z)\,\overline{ L'(Z)}}{Q(Z,\,j(Z))} \qquad L, L' \in V,\quad Z \in
4869: K\, \,.
4870: \end{equation}
4871: ($\forall f,g \in C(K)$)
4872:
4873: \medskip One then defines a generalized crossed product $C^*$-algebra
4874: $C(K) \times_{\sigma,\,\call}
4875: \mathbb {Z} $ following M. Pimsner \cite{pims:1997}. Given a compact
4876: space $K$, an homeomorphism $\sigma$ of $K$ and a hermitian line
4877: bundle $\call$ on $K$ we define the $C^\ast$-algebra $C(K)
4878: \times_{\sigma,\,\call} \mathbb {Z} $ as the twisted cross-product
4879: of $C(K)$ by the Hilbert $C^*$-bimodule associated to $\call$ and
4880: $\sigma$ (\cite{aba-eil-exel:1998}, \cite{pims:1997}).
4881:
4882: We let for each $n \geq 0$, $\call^{\sigma^n}$ be the hermitian line
4883: bundle pullback of $\call$ by $\sigma^n$ and (cf.
4884: \cite{art-tat-vdb:1990}, \cite{smi-sta:1992})
4885: \begin{equation}
4886: \call_n := \call \otimes \call^{\sigma} \otimes \cdots \otimes
4887: \call^{\sigma^{n-1}} \label{gene2}
4888: \end{equation}
4889: We first define a $\ast$-algebra as the linear span of the monomials
4890: \begin{equation}
4891: \xi \, W^n\, , \quad W^{\ast n} \, \eta^\ast \,,\quad \xi\,,\eta \in
4892: C(K,\call_n) \label{gene}
4893: \end{equation}
4894: with product given as in (\cite{art-tat-vdb:1990},
4895: \cite{smi-sta:1992}) for $(\xi_1 \, W^{n_1})\,(\xi_2 \, W^{n_2})$ so
4896: that
4897: \begin{equation}
4898: (\xi_1 \, W^{n_1})\,(\xi_2 \, W^{n_2}):= (\xi_1 \otimes
4899: (\xi_2\circ{\sigma^{n_1}}) )\, W^{n_1+n_2} \label{gene3}
4900: \end{equation}
4901: We use the hermitian structure of $\call_n $ to give meaning to the
4902: products $\eta^\ast \,\xi$ and $\xi \;\eta^\ast$ for $\xi\,,\eta \in
4903: C(K,\call_n)$. The product then extends uniquely to an associative
4904: product of $\ast$-algebra fulfilling the following additional rules
4905: \begin{equation}
4906: (W^{\ast k} \, \eta^\ast)\,( \xi \, W^k):= \, (\eta^\ast\, \xi)\circ
4907: \sigma^{-k}\,,\qquad ( \xi \, W^k)\,(W^{\ast k} \, \eta^\ast)\,:= \,
4908: \xi \;\eta^\ast \label{gene1}
4909: \end{equation}
4910:
4911:
4912: The $C^\ast$-norm of $C(K) \times_{\sigma,\,\call} \mathbb {Z} $
4913: is defined as for ordinary cross-products and due to the amenability of the group $\mathbb {Z} $
4914: there is no distinction between the reduced and maximal norms. The
4915: latter is obtained as the supremum of the norms in involutive
4916: representations in Hilbert space. The natural positive conditional
4917: expectation on the subalgebra $C(K)$ shows that the $C^\ast$-norm
4918: restricts to the usual sup norm on $C(K)$.
4919:
4920:
4921:
4922: \medskip
4923:
4924:
4925: \begin{thm}\label{C*}
4926: Let $K \subset E$ be a compact
4927: $\sigma$-invariant subset and $Q$ be central and strictly positive
4928: on $\{(Z,\,\bar Z);\, Z\in K\}$. Let $\call$ be the restriction to
4929: $K$ of the dual of the tautological line bundle on $P(V^\ast)$
4930: endowed with the hermitian metric $\langle\;,\; \rangle_Q$.
4931:
4932:
4933: (i) The equality $\sqrt{2}\,\theta(Y):= Y\, W + W^\ast\,\bar Y^\ast$
4934: yields a $\ast$-homomorphism $$\theta:\cala=A(V,R) \to C(K)
4935: \times_{\sigma,\,\call} \mathbb {Z} $$
4936:
4937:
4938: (ii) For any $Y \in V$
4939: the $C^\ast$-norm of $\theta(Y)$ fulfills
4940: $$\sup_K \|Y\|\leq \sqrt{2}\| \,\theta(Y)\|
4941: \leq 2\sup_K \|Y\| $$
4942:
4943:
4944: (iii) If $\sigma^4 \neq \bbbone$, then $\theta(Q)= 1$ where
4945: $Q$ is viewed as an element of $T(V)/(R)$.
4946: \end{thm}
4947:
4948: \bigskip
4949:
4950:
4951: In the above case of the sphere $S^3_\varphi$ one lets $Q$ be the
4952: quadratic form
4953: \begin{equation}
4954: Q(x,\,x'):=\sum x_\mu\,x'_\mu
4955: \label{quad}
4956: \end{equation}
4957:
4958: In the generic case one has :
4959:
4960: \begin{prop} 1) The characteristic variety is the union of 4 points with an elliptic curve $F_\varphi$.
4961:
4962: 2) The quadratic form $Q$ is central and positive on $F_\varphi
4963: \times F_\varphi$.
4964: \end{prop}
4965:
4966: \smallskip
4967: In suitable coordinates the equations defining the elliptic curve
4968: $F_\varphi$ are
4969: \begin{equation}
4970: \frac{Z_0^2-Z_1^2}{
4971: s_1}=\frac{Z_0^2-Z_2^2}{s_2}=\frac{Z_0^2-Z_3^2}{s_3} \label{charZ}
4972: \end{equation}
4973: where $s_k := 1 + t_\ell\, t_m \, , \: t_k:= {\rm tan}\,\varphi_k$.
4974:
4975: The positivity of $Q$ is automatic since in the coordinates $x$ the
4976: involution $j_\varphi$
4977: of the $\ast$-algebra $C_{\mathrm{alg}}({\mathbb R}^4_\varphi)$
4978: is just $j_\varphi(Z)= \bar Z$, so that $Q(X,\,j_\varphi(X))>0$ for
4979: $X \neq 0$.
4980:
4981:
4982: \begin{cor} \label{II}
4983: Let $K \subset F_\varphi $ be a compact $\sigma$-invariant subset.
4984: The homomorphism $\theta$ of Theorem \ref{C*} is a unital
4985: $\ast$-homomorphism from $ C_{\mathrm{alg}}(S^3_\varphi)$
4986: to the cross-product $ C^{\infty}(K) \times_{\sigma,\,\call} \mathbb
4987: {Z} $.
4988: \end{cor}
4989:
4990:
4991:
4992:
4993: It follows that one obtains a
4994: non-trivial $C^\ast$-algebra $C^\ast(S^3_\varphi)$ as the completion
4995: of $ C_{\mathrm{alg}}(S^3_\varphi)$ for the semi-norm,
4996: \begin{equation}
4997: \| P \|:= \sup \| \,\pi(P) \|
4998: \label{norm}
4999: \end{equation}
5000: where $\pi$ varies through all unitary representations of $
5001: C_{\mathrm{alg}}(S^3_\varphi)$. It was clear from the start that
5002: \eqref{norm} defines a finite $C^\ast$-semi-norm on $
5003: C_{\mathrm{alg}}(S^3_\varphi)$ since the equation of the sphere
5004: $\sum x_\mu^2=1$ together with the self-adjointness
5005: $x_\mu=\,x_\mu^*$ show that in any unitary representation one has
5006: $$
5007: \|\, \pi(x_\mu) \|\leq 1\qqq \mu\,.
5008: $$
5009: What the above corollary gives is a lower bound for the
5010: $C^\ast$-norm such as that given by statement (ii) of Theorem
5011: \ref{C*} on the linear subspace $V$ of generators.
5012:
5013:
5014: \smallskip The correspondence
5015: $\sigma$ on $F_\varphi $,
5016: is for generic $\varphi$ a translation of module $\eta$ of the elliptic curve
5017: $F_\varphi$ and one distinguishes two cases : the {\em even} case when it preserves
5018: the two real components of the curve $F_\varphi \cap
5019: P_3(\mathbb {R})$ and the odd case when it permutes them.
5020:
5021: \begin{prop} \label{siginvar} Let $\varphi $ be generic and even.
5022:
5023: (i) The cross-product $ C(F_\varphi) \times_{\sigma,\,\call} \mathbb
5024: {Z} $ is isomorphic to the mapping torus of the automorphism $\beta
5025: $ of the noncommutative torus ${\mathbb T}_{\eta}^2 = C_\varphi
5026: \times_\sigma \mathbb {Z} $ acting on the generators by the matrix
5027: $\left[
5028: \begin{array}{cc}
5029: 1& 4\\
5030: 0& 1
5031: \end{array}
5032: \right] $.
5033:
5034:
5035: (ii) The crossed product $F_\varphi \times_{\sigma,\,\call} \mathbb
5036: {Z} $ is a noncommutative $3$-manifold with an elliptic
5037: action of the three dimensional
5038: Heisenberg Lie algebra $\frach_3$ and an invariant trace $\tau$.
5039:
5040:
5041: \end{prop}
5042:
5043: \medskip
5044:
5045: It follows that one is exactly in the framework developed in
5046: \cite{C3}. We refer to \cite{Rieffel89} and \cite{aba-exel:1997}
5047: where these noncommutative manifolds were analyzed in terms of
5048: crossed products by Hilbert $C^*$-bimodules.
5049:
5050:
5051: Integration on the translation invariant volume form $dv$
5052: of $F_\varphi$ gives the $\frach_3$-invariant trace $\tau$,
5053: \begin{eqnarray}
5054: \label{trace}
5055: \tau(f)& = & \int f dv\,,\quad \forall f \in C^{\infty}(F_\varphi)\nonumber\\
5056: \tau(\xi \, W^k)& = &\tau(W^{\ast k} \, \eta^\ast)\,=\,0\,,\quad
5057: \forall k\neq 0
5058: \end{eqnarray}
5059: It follows in particular that the results of \cite{C3} apply to
5060: obtain the calculus. In particular the following gives the
5061: ``fundamental class" as a $3$-cyclic cocycle,
5062: \begin{equation}\label{3trace}
5063: \tau_3(a_0,\,a_1,\,a_2 ,\,a_3)=\,\sum
5064: \epsilon_{ijk}\,\tau(a_0\,\delta_i(a_1)\,\delta_j(a_2)
5065: \,\delta_k(a_3))
5066: \end{equation}
5067: where the $\delta_j$ are the generators of the action of $\frach_3$.
5068:
5069:
5070: \medskip The relation between the noncommutative spheres
5071: $S^3_\varphi$ and the noncommutative nilmanifolds $F_\varphi
5072: \times_{\sigma,\,\call} \mathbb {Z} $ is analyzed in
5073: \cite{CDV2}, \cite{CDV3} thanks to the computation of the
5074: Jacobian of the homomorphism $\theta$.
5075:
5076: \bigskip
5077:
5078:
5079:
5080:
5081: \section{$\Q$-lattices}\label{Qlatt}
5082:
5083: A class of examples of noncommutative spaces of relevance to
5084: number theory is given by the moduli spaces of $\Q$-lattices up to
5085: commensurability. These fall within the general framework of
5086: noncommutative spaces obtained as quotients of equivalence relations
5087: discussed in Section \ref{quotients}.
5088:
5089: A $\Q$-lattice in $\R^n$ consists of a pair $( \Lambda , \phi) $
5090: of a lattice $\Lambda\subset \R^n$ (a cocompact free abelian
5091: subgroup of $\R^n$ of rank $n$) together with a system of labels
5092: of its torsion points given by a homomorphism of abelian groups
5093: \begin{equation}\label{phimap}
5094: \phi : \Q^n/\Z^n \longrightarrow \Q\Lambda / \Lambda.
5095: \end{equation}
5096:
5097: Two $\Q$-lattices are commensurable,
5098: $$ (\Lambda_1, \phi_1) \sim (\Lambda_2, \phi_2), $$
5099: iff $\Q\Lambda_1=\Q\Lambda_2$ and
5100: $$ \phi_1 = \phi_2 \mod \Lambda_1 + \Lambda_2 $$
5101:
5102: In general, the map $\phi$ of \eqref{phimap} is just a group
5103: homomorphism. A $\Q$-lattice is said to be {\em invertible}
5104: is $\phi$ is an isomorphism. Two invertible $\Q$-lattices are
5105: commensurable if and only if they are equal.
5106:
5107: \begin{center}
5108: \begin{figure}
5109: \includegraphics[scale=0.95]{Qlat.eps}
5110: \caption{$\Q$-lattices: generic and invertible case.
5111: \label{FigQlat}}
5112: \end{figure}
5113: \end{center}
5114:
5115: The space $\cL_n$ of commensurabilty classes of
5116: $\Q$-lattices in $\R^n$ has the typical property of noncommutative
5117: spaces: it has the cardinality of the continuum but one cannot
5118: construct a countable collection of measurable functions that
5119: separate points of $\cL_n$.
5120: Thus, one can use noncommutative geometry to describe the quotient
5121: space $\cL_n$ through a noncommutative $C^*$-algebra $C^*(\cL_n)$.
5122:
5123: We consider especially the case of $n=1$ and $n=2$. One is also
5124: interested in the $C^*$-algebras describing $\Q$-lattices up to
5125: scaling, $\cA_1=C^*(\cL_1/\R_+^*)$ and $\cA_2=C^*(\cL_2/\C^*)$.
5126:
5127: In the 1-dimensional case, a $\Q$-lattice can always be written in the
5128: form
5129: \begin{equation}\label{1dimQlat}
5130: ( \Lambda , \phi) \, = (\lambda\, \Z,\lambda\,\rho)
5131: \end{equation}
5132: for some $\lambda>0$ and some
5133: \begin{equation}\label{rho}
5134: \rho \in \Hom(\Q/\Z,\Q/\Z)=\varprojlim \Z/n\Z = \hat\Z.
5135: \end{equation}
5136: By considering lattices up to scaling, we eliminate the factor
5137: $\lambda>0$ so that 1-dimensional $\Q$-lattices up to scale are
5138: completely specified by the choice of the element $\rho\in
5139: \hat\Z$. Thus, the algebra of coordinates of the space of
5140: 1-dimensional $\Q$-lattices up to scale is the commutative
5141: $C^*$-algebra
5142: \begin{equation}\label{ChatZ}
5143: C(\hat\Z)\simeq C^*(\Q/\Z),
5144: \end{equation}
5145: where we use Pontrjagin duality to get the identification in
5146: \eqref{ChatZ}.
5147:
5148: The equivalence relation of commensurability is implemented by the
5149: action of the semigroup $\N^\times$ on $\Q$-lattices. The
5150: corresponding action on the algebra \eqref{ChatZ} is by
5151: \begin{equation}\label{Nact}
5152: \alpha_n(f) (\rho)=\left\{ \begin{array}{lr} f(n^{-1} \rho) & \rho \in
5153: n\hat\Z \\ 0 & \text{otherwise.} \end{array}\right.
5154: \end{equation}
5155:
5156: Thus, the quotient of the space of 1-dimensional $\Q$-lattices up
5157: to scale by the commensurability relation and its algebra of
5158: coordinates is given by the semigroup crossed product
5159: \begin{equation}\label{semicross}
5160: C^*(\Q/\Z)\rtimes \N^\times .
5161: \end{equation}
5162: This is the Bost--Connes $C^*$-algebra introduced in \cite{BC}.
5163:
5164: It has a natural time evolution given by the covolume of a pair of
5165: commensurable $\Q$-lattices. It has symmetries (compatible with the
5166: time evolution) given by the group $\hat\Z^*=\GL_1(\A_f)/\Q^*$ and
5167: the KMS (Kubo--Martin--Schwinger) equilibrium states of the system
5168: have interesting arithmetic properties. Namely, the partition function
5169: of the system is the Riemann zeta function. There is a unique KMS
5170: state for sufficiently high temperature, while at low temperature the
5171: system undergoes a phase transition with spontaneous symmetry
5172: breaking. The pure phases (estremal KMS states) at low temperature
5173: are parameterized by elements in $\hat\Z^*$. They have an explicit
5174: expression in terms of polylogarithms at roots of unity. At zero
5175: temperature the extremal KMS states, evaluated on the elements of a
5176: rational subalgebra affect values that are algebraic numbers.
5177: The action on these values of the Galois group $\Gal(\bar\Q/\Q)$
5178: factors through its abelianization and
5179: is obtained (via the class field theory isomorphism $\hat\Z^* \cong
5180: \Gal(\Q^{ab}/\Q)$) as the action of symmetries on the algebra (\cf
5181: \cite{BC}, \cite{CoMa}, \cite{CoMajgp1} for details).
5182:
5183: \medskip
5184:
5185: In the 2-dimensional case, a $\Q$-lattice can be written in the form
5186: $$ (\Lambda,\phi)=(\lambda (\Z+\Z\tau),\lambda\rho), $$
5187: for some $\lambda\in \C^*$, some $\tau\in \H$, and some $\rho\in
5188: M_2(\hat\Z)=\Hom(\Q^2/\Z^2,\Q^2/\Z^2)$. Thus, the space of
5189: 2-dimensional $\Q$-lattices up to the scale factor $\lambda\in
5190: \C^*$ and up to isomorphisms, is given by
5191: \begin{equation}\label{2dQlat}
5192: M_2(\hat\Z)\times \H \mod \Gamma=\SL_2(\Z).
5193: \end{equation}
5194: The commensurability relation giving the
5195: space $\cL_2/\C^*$ is implemented by the partially
5196: defined action of $\GL_2^+(\Q)$.
5197:
5198: One considers in this case the quotient of the space
5199: \begin{equation}\label{tildeU}
5200: \tilde \cU:= \{ (g,\rho,\alpha)\in \GL_2^+(\Q)\times
5201: M_2(\hat\Z)\times \GL_2^+(\R)\, :\,\, g\rho\in M_2(\hat\Z) \}
5202: \end{equation}
5203: by the action of $\Gamma \times \Gamma$ given by
5204: \begin{equation}\label{Gamma2action}
5205: (\gamma_1,\gamma_2) \, (g,\rho,\alpha)=(\gamma_1 g \gamma_2^{-1},
5206: \gamma_2 \rho, \gamma_2 \alpha).
5207: \end{equation}
5208:
5209: The groupoid $\cR_2$ of the equivalence relation of
5210: commensurability on 2-dimensional $\Q$-lattices (not considered up
5211: to scaling for the moment) is a locally compact groupoid, which
5212: can be parameterized by the quotient of \eqref{tildeU} by
5213: $\Gamma\times \Gamma$ via the map $r: \tilde \cU \to \cR_2$,
5214: \begin{equation}\label{mapgr}
5215: r(g,\rho,\alpha)=\left( (\alpha^{-1}g^{-1} \Lambda_0, \alpha^{-1}
5216: \rho), (\alpha^{-1}\Lambda_0, \alpha^{-1} \rho)\right).
5217: \end{equation}
5218:
5219: We then consider the quotient by scaling.
5220: The quotient $\GL_2^+(\R)/\C^*$ can be identified with the
5221: hyperbolic plane $\H$ in the usual way.
5222: If $(\Lambda_k,\phi_k)$ $k=1,2$ are a pair of commensurable
5223: 2-dimensional $\Q$-lattices, then for any $\lambda\in \C^*$, the
5224: $\Q$-lattices $(\lambda \Lambda_k, \lambda \phi_k)$ are also
5225: commensurable, with
5226: $$ r(g,\rho,\alpha\lambda^{-1})=\lambda r(g,\rho,\alpha). $$
5227: However, the action of $\C^*$ on $\Q$-lattices is not free due to
5228: the presence of lattices (such as $\Lambda_0$ above) with nontrivial
5229: automorphisms. Thus, the quotient $Z=\cR_2/\C^*$ is no longer a
5230: groupoid. Still, one can define a convolution algebra for $Z$ by
5231: restricting the convolution product of $\cR_2$ to homogeneous
5232: functions of weight zero, where a function $f$ has weight $k$ if
5233: it satisfies
5234: $$ f(g,\rho,\alpha\lambda)=\lambda^k f(g,\rho,\alpha), \ \ \
5235: \forall \lambda \in \C^*. $$
5236: The space $Z$ is the quotient of the space
5237: \begin{equation}\label{Uspace}
5238: \cU:=\{ (g,\rho,z) \in \GL_2^+(\Q)\times M_2(\hat\Z)\times \H |
5239: g\rho\in M_2(\hat\Z) \}
5240: \end{equation}
5241: by the action of $\Gamma\times \Gamma$. Here the space
5242: $M_2(\hat\Z)\times \H$ has a partially defined action of
5243: $\GL_2^+(\Q)$ given by
5244: $$ g (\rho,z)=(g\rho, g(z)), $$
5245: where $g(z)$ denotes action as fractional linear transformation.
5246:
5247: Thus, the algebra of coordinates $\cA_2$ for the noncommutative space
5248: of commensurability classes of 2-dimensional $\Q$-lattices up
5249: to scaling is given by the following convolution algebra.
5250:
5251: Consider the space $C_c(Z)$ of continuous compactly supported
5252: functions on $Z$. These can be seen, equivalently, as functions on
5253: $\cU$ as in \eqref{Uspace} invariant under the
5254: $\Gamma\times\Gamma$ action $(g,\rho,
5255: z)\mapsto(\gamma_1g\gamma_2^{-1},\gamma_2 z)$.
5256: One endows $C_c(Z)$ with the convolution product
5257: \begin{equation}\label{Heckeprod2}
5258: (f_1*f_2)(g,\rho,z)= \displaystyle{\sum_{s\in \Gamma\backslash
5259: \GL_2^+(\Q): s\rho\in M_2(\hat\Z)}} f_1(gs^{-1},s\rho,s(z))
5260: f_2(s,\rho,z)
5261: \end{equation}
5262: and the involution
5263: $f^*(g,\rho,z)=\overline{f(g^{-1},g\rho,g(z))}$.
5264:
5265: Again there is a time evolution on this algebra, which is given by the
5266: covolume,
5267: \begin{equation}\label{evolution2}
5268: \sigma_t(f) (g,\rho,z) = \det(g)^{it} \, f(g,\rho,z).
5269: \end{equation}
5270: The partition function for this $\GL_2$ system is given by
5271: \begin{equation}\label{partition2}
5272: Z(\beta)= \sum_{m\in \Gamma\backslash M_2^+(\Z)}\det(m)^{-\beta}
5273: =\sum_{k=1}^\infty \sigma(k)\,
5274: k^{-\beta}=\zeta(\beta)\zeta(\beta-1),
5275: \end{equation}
5276: where $\sigma(k)=\sum_{d|k} d$. The form of the partition
5277: function suggests the possibility that two distinct phase
5278: transitions might happen at $\beta=1$ and $\beta=2$.
5279:
5280: The structure of KMS states for this system is analysed in
5281: \cite{CoMa}. The main result is the following.
5282:
5283: \begin{thm}\label{GL2KMS}
5284: The KMS$_\beta$ states of the $\GL_2$-system have the following
5285: properties:
5286: \begin{enumerate}
5287: \item In the range $\beta\leq 1$ there are no KMS states.
5288: \item In the range $\beta>2$ the set of extremal KMS states is
5289: given by the classical Shimura variety
5290: \begin{equation}\label{Ekmsbeta2}
5291: \cE_\beta \cong \GL_2(\Q)\backslash \GL_2(\A) /\C^*.
5292: \end{equation}
5293: \end{enumerate}
5294: \end{thm}
5295:
5296: The symmetries are more complicated than in the Bost--Connes case. In
5297: fact, in addition to symmetries given by automorphisms that commute
5298: with the time evolution, there are also symmetries by {\em
5299: endomorphisms} that play an important role. The resulting symmetry
5300: group is the quotient $\GL_2(\A_f)/\Q^*$. An important result of
5301: Shimura \cite{Shi} shows that this group is in fact the Galois group
5302: of the field $F$ of modular functions. The group $\GL_2(\A_f)$
5303: decomposes as a product
5304: \begin{equation}\label{SymmGL2}
5305: \GL_2(\A_f)=\GL_2^+(\Q) \GL_2(\hat\Z),
5306: \end{equation}
5307: where $\GL_2(\hat\Z)$ acts by automorphisms related to the
5308: deck transformations of the tower of the modular curves, while
5309: $\GL_2^+(\Q)$ acts by endomorphisms that move across levels in
5310: the modular tower.
5311:
5312: The modular field $F$ is the field of modular functions over
5313: $\Q^{ab}$, namely the union of the fields $F_N$ of modular
5314: functions of level $N$ rational over the cyclotomic field
5315: $\Q(\zeta_n)$, that is, such that the $q$-expansion in powers of
5316: $q^{1/N}=\exp(2\pi i \tau/N)$ has all coefficients in $\Q(e^{2\pi
5317: i/N})$.
5318:
5319: The action of the Galois group $\hat\Z^* \simeq \Gal(\Q^{ab}/\Q)$
5320: on the coefficients determines a homomorphism
5321: \begin{equation}\label{cyclhom}
5322: {\rm cycl}: \hat\Z^* \to \Aut(F).
5323: \end{equation}
5324:
5325: If $\tau\in \H$ is a generic point, then the evaluation map
5326: $f\mapsto f(\tau)$ determines an embedding $F\hookrightarrow \C$. We
5327: denote by $F_\tau$ the image in $\C$. This yields an
5328: identification
5329: \begin{equation}\label{GalFtau}
5330: \theta_\tau: \Gal(F_\tau/\Q)
5331: \stackrel{\simeq}{\to} \Q^* \backslash \GL_2(\A_f).
5332: \end{equation}
5333:
5334: There is an arithmetic algebra $\cA_{2,\Q}$ (defined over $\Q$) of
5335: unbounded multipliers of the $C^*$-algebra $\cA_2$, obtained by
5336: considering continuous functions on $Z$ (\cf
5337: \eqref{Uspace}), with finite support in the variable $g\in
5338: \Gamma\backslash\GL_2^+(\Q)$ and with the following properties.
5339: Let $p_N: M_2(\hat\Z)\to M_2(\Z/N\Z)$ be the canonical
5340: projection. With the notation $f_{(g,\rho)}(z) = f(g,\rho,z)$,
5341: we say that $f_{(g,\rho)}\in C(\H)$ is of level $N$ if
5342: $$ f_{(g,\rho)} = f_{(g,p_N(\rho))} \ \ \ \ \forall (g,\rho). $$
5343: We require that elements of $\cA_{2,\Q}$ have the $f_{(g,\rho)}$ of
5344: finite level with $ f_{(g,m)} \in F$ for all $(g,m)$. We also
5345: require that the action \eqref{cyclhom} on the coefficients
5346: of the q-expansion of the $f_{(g,m)}$ satisfies
5347: $$ f_{(g,\alpha(u)m)} = {\rm cycl}(u) \, f_{(g,m)}, $$
5348: for all $g\in \GL_2^+(\Q)$ diagonal and all $u\in \hat\Z^*$, with
5349: $$ \alpha(u)=\begin{pmatrix} u& 0 \\ 0 & 1 \end{pmatrix}, $$
5350: to avoid some ``trivial'' elements that would spoil the Galois
5351: action on values of states (\cf \cite{CoMa}, \cite{CoMajgp1}).
5352: The action of symmetries extends to $\cA_{2,\Q}$.
5353: We have then the following result (\cite{CoMa}):
5354:
5355:
5356: \begin{thm}\label{GalGL2infty}
5357: Consider a state $\varphi=\varphi_{\infty,L}\in \cE_\infty$, for a
5358: generic invertible $\Q$-lattice $L=(\rho,\tau)$. Then the
5359: values of the state on elements of the arithmetic
5360: subalgebra generate the image in $\C$ of the modular field,
5361: \begin{equation}\label{values}
5362: \varphi(\cA_{2,\Q})\subset F_\tau,
5363: \end{equation}
5364: and the isomorphism
5365: \begin{equation}\label{thetaphi1}
5366: \theta_\varphi : \Gal(F_\tau/\Q)
5367: \stackrel{\simeq}{\longrightarrow} \Q^* \backslash \GL_2(\A_f),
5368: \end{equation}
5369: given by
5370: \begin{equation}\label{thetaphi2}
5371: \theta_\varphi(\gamma)=\rho^{-1}\, \theta_\tau(\gamma) \, \rho,
5372: \end{equation}
5373: for $\theta_\tau$ as in \eqref{GalFtau}, intertwines the Galois
5374: action on the values of the state with the action of symmetries,
5375: \begin{equation}\label{intertwineGL2}
5376: \gamma\, \varphi(f) = \varphi( \theta_\varphi(\gamma) f), \ \ \ \
5377: \forall f\in \cA_{2,\Q}, \ \ \forall\gamma\in \Gal(F_\tau/\Q).
5378: \end{equation}
5379: \end{thm}
5380:
5381: \medskip
5382:
5383: A notion analogous to that of $\Q$-lattices can be given for other
5384: number fields $\K$. This notion was used in \cite{CMR} to construct a
5385: quantum statistical mechanical system for $\K$ an imaginary quadratic
5386: field. This system shares properties with both the Bost--Connes system
5387: of \cite{BC} and the $\GL_2$ system (2-dimensional $\Q$-lattices) of
5388: \cite{CoMa}.
5389:
5390: We assume that $\K=\Q(\sqrt{-d})$, $d$ a positive integer. Let
5391: $\tau\in \H$ be such that
5392: $\K=\Q(\tau)$ and $\cO = \Z + \Z \tau$ is the ring of integers of $\K$.
5393:
5394: A 1-dimensional $\K$-lattice
5395: $(\Lambda,\phi)$ is a finitely generated $\cO$-submodule $\Lambda\subset \C$,
5396: such that $\Lambda\otimes_\cO \K \cong \K$, together with a morphism of
5397: $\cO$-modules
5398: \begin{equation}\label{Kphi}
5399: \phi : \K/\cO \to \K\Lambda/\Lambda.
5400: \end{equation}
5401: A 1-dimensional $\K$-lattice is {\em invertible} if $\phi$ is an
5402: isomorphism of $\cO$-modules.
5403: A 1-dimensional $\K$-lattice is, in particular, a 2-dimensional
5404: $\Q$-lattice.
5405:
5406: We consider the
5407: notion of commensurability as in the case of $\Q$-lattices.
5408: Two 1-dimensional $\K$-lattices
5409: $(\Lambda_1,\phi_1)$ and
5410: $(\Lambda_2,\phi_2)$ are commensurable if $\K\Lambda_1=\K\Lambda_2$
5411: and $\phi_1=\phi_2$ modulo $\Lambda_1+\Lambda_2$.
5412: In particular, two 1-dimensional $\K$-lattices are commensurable iff
5413: the underlying $\Q$-lattices are commensurable.
5414:
5415: The algebra of the corresponding noncommutative space is a restriction
5416: of the algebra of the $\GL_2$-system to the subgroupoid of the
5417: equivalence of commensurability restricted to $\K$-lattices.
5418: The time evolution is also a restriction from the $\GL_2$-system.
5419:
5420: The resulting system has partition function the Dedekind zeta function
5421: $\zeta_\K(\beta)$ of the number field $\K$. Above the critical temperature
5422: $T=1$ there is a unique KMS state, while at lower temperatures the
5423: extremal KMS states are parameterized by elements of $\A_\K^*/\K^*$,
5424: where $\A_K=\A_{K,f}\times \C$ are the adeles of $\K$, with
5425: $\A_{K,f}=\A_f\otimes \K$. The KMS states at zero temperature,
5426: evaluated on the restriction to $\K$-lattices of the arithmetic
5427: algebra of the $\GL_2$-system, have an action of the Galois group
5428: $\Gal(\K^{ab}/\K)$ realized (via the class field theory isomorphism)
5429: through the action of symmetries (automorphisms and endomorphisms) of
5430: the system (\cf \cite{CMR}).
5431:
5432:
5433: \bigskip
5434:
5435: \section{Modular Hecke algebras}\label{hecke}
5436:
5437: Connes and Moscovici \cite{Co-Mosc1} defined modular Hecke algebras
5438: $\cA(\Gamma)$ of level $\Gamma$, a congruence subgroup of
5439: $\PSL_2(\Z)$. These extend both the ring of classical Hecke operators
5440: and the algebra of modular forms.
5441:
5442: Modular Hecke algebras encode two \textit{a priori} unrelated
5443: structures on modular forms, namely the algebra structure given by
5444: the pointwise product on one hand, and the action of the Hecke
5445: operators on the other. To any congruence subgroup $\Gamma$ of $\,
5446: \SL_2(\Z)$ corresponds a crossed product algebra
5447: $\cA(\Gamma)$, the \textit{modular Hecke algebra} of level $\Gamma$,
5448: which is a direct extension of both the ring of classical Hecke
5449: operators and of the algebra $\cM(\Gamma)$ of $\Gamma$-modular
5450: forms.
5451:
5452: These algebras can be obtained by considering the action of
5453: $\GL^+_2(\Q)$ on the algebra of modular forms on the full (adelic)
5454: modular tower, which yields the ``holomorphic part" of the ``ring of
5455: functions" of the noncommutative space of commensurability classes of
5456: 2-dimensional $\Q$-lattices, introduced in Section \ref{Qlatt}.
5457:
5458: With $\cM$ denoting the algebra of modular forms of arbitrary
5459: level, the elements of $\cA(\Gamma)$ are maps with finite support
5460: $$
5461: F: \Gamma\backslash \GL^+ (2, \mathbb{Q}) \to \cM \, ,
5462: \qquad \Gamma \alpha \mapsto F_{\alpha} \in \cM \, ,
5463: $$
5464: satisfying the covariance condition
5465: \begin{equation}
5466: F_{\alpha \gamma} \, = \, F_{\alpha} \vert \gamma \qqq\alpha \in
5467: \GL^+ (2, \mathbb{Q}) \, , \gamma \in \Gamma \, \nonumber
5468: \end{equation}
5469: and their product is given by convolution.
5470:
5471: More in detail,
5472: let $G=\PGL^+_2(\Q)$ and $\Gamma \subset \PSL_2(\Z)$ a finite
5473: index subgroup. The quotient map $\Gamma \backslash G \to \Gamma
5474: \backslash G / \Gamma$ is finite to one, and $\Gamma$ acts on $\C
5475: [\Gamma \backslash G]$. Let $\cH_k$ be the space of holomorphic
5476: functions $f: \H \to \C$ with polynomial growth, and with the
5477: action $|_k$, for $k\in 2\Z$, of $\PGL^+_2(\R)$ of the form
5478: $$ \left(f|_k \left(\begin{array}{cc} a& b\\ c& d \end{array}\right)\right)
5479: (z)= \frac{(ad-bc)^{k/2}}{(cz+d)^k} \, f\left( \frac{az+b}{cz+d}
5480: \right).
5481: $$
5482: This determines induced actions of $G$ and $\Gamma$ on $\cH_k$.
5483: The space of modular forms is obtained as $\cM_k(\Gamma)=
5484: \cH_k^\Gamma$, the invariants of this action.
5485:
5486: One can then define
5487: \begin{equation}\label{AkGamma}
5488: \cA_k (\Gamma) := \left( \C [ \Gamma \backslash G ] \otimes_\C
5489: \cH_k \right)^\Gamma,
5490: \end{equation}
5491: with respect to the right action of $\Gamma$,
5492: $$ \gamma : \sum_i (\Gamma g_i) \otimes f_i \mapsto \sum_i \Gamma g_i
5493: \gamma \otimes (f_i |_k \gamma).
5494: $$
5495: One considers the graded vector space $\cA_*(\Gamma) = \oplus_k
5496: \cA_k(\Gamma)$. The elements of $\cA_k(\Gamma)$ can be thought of
5497: as finitely supported $\Gamma$-equivariant maps
5498: $$ \phi: \Gamma\backslash G \to \cH_k \ \ \ \sum_i (\Gamma
5499: g_i)\otimes f_i \mapsto f_i. $$
5500: We can embed
5501: $$ \cA_*(\Gamma)\subset \hat\cA_*(\Gamma):= \Hom_\Gamma
5502: (\C[\Gamma\backslash G], \cH_k), $$ where we think of
5503: $\cA_*=\cH_*[\Gamma\backslash G]$ as polynomials in
5504: $\Gamma\backslash G$ with $\cH_*$ coefficients, and of $\hat\cA_*=
5505: \cH_* [[ \Gamma\backslash G ]]$ as formal power series, that is,
5506: $\Gamma$-equivariant maps $\phi: \Gamma \backslash G \to \cH_k$.
5507: There is on $\cA_*(\Gamma)$ an associative multiplication (\cf
5508: \cite{Co-Mosc1}), which makes
5509: $\cA_*(\Gamma)$ into a noncommutative ring. This is given by a
5510: convolution product. For any $\phi \in \cA_k(\Gamma)$, we have
5511: $\phi_g = \phi_{\gamma g}$, with $\phi_g=0$ off a finite subset of
5512: $\Gamma \backslash G$, and $\phi_g | \gamma = \phi_{g\gamma}$, so
5513: these terms are left $\Gamma$-invariant and right
5514: $\Gamma$-equivariant. For $\phi \in \cA_k(\Gamma)$ and $\psi \in
5515: \cA_\ell (\Gamma)$ we then define the convolution product as
5516: \begin{equation}\label{convAprod}
5517: ( \phi * \psi )_g := \sum_{(g_1,g_2)\in G\times_\Gamma G, g_1 g_2
5518: =g} (\phi_{g_1} | g_2) \phi_{g_2},
5519: \end{equation}
5520:
5521: The algebra $\cA_*(\Gamma)$ constructed this way has two
5522: remarkable subalgebras.
5523: \begin{itemize}
5524: \item $\cA_0(\Gamma)= \C [\Gamma \backslash G / \Gamma]$ is the
5525: algebra $\bT$ of Hecke operators.
5526: \item $\cM_k(\Gamma)\subset \cA_k(\Gamma)$ also gives a subalgebra
5527: $\cM_*(\Gamma) \subset \cA_*(\Gamma)$.
5528: \end{itemize}
5529:
5530: In particular observe that all the coefficients $\phi_g$ are
5531: modular forms. In fact, they satisfy $\phi_g | \gamma =
5532: \phi_{g\gamma}$, hence, for $\gamma \in \Gamma$, this gives
5533: $\phi_g | \gamma = \phi_g$.
5534:
5535: Notice however that the convolution product on $\cA_*$ does not
5536: agree with the Hecke action, namely the diagram
5537: \begin{eqnarray*}
5538: \diagram \cM_*(\Gamma) \otimes \bT \dto^{\iota}\rto^{H} &
5539: \cM_*(\Gamma) \dto^{\iota} \\
5540: \cA_*\otimes \cA_* \rto^{*}& \cA_*,
5541: \enddiagram
5542: \end{eqnarray*}
5543: with $\iota$ the inclusion of subalgebras and $H$ the Hecke
5544: action, is {\em not} commutative, nor is the symmetric one
5545: \begin{eqnarray*}
5546: \diagram \bT \otimes \cM_*(\Gamma) \dto^{\iota}\rto^{H} &
5547: \cM_*(\Gamma) \dto^{\iota} \\
5548: \cA_*\otimes \cA_* \rto^{*}& \cA_*.
5549: \enddiagram
5550: \end{eqnarray*}
5551: To get the correct Hecke action on modular forms from the algebra
5552: $\cA_*(\Gamma)$, one needs to introduce the augmentation map
5553: $$ \epsilon : \C [\Gamma \backslash G] \to \C $$
5554: extended to a map
5555: $$ \epsilon\otimes 1 : \C[\Gamma \backslash G] \otimes \cH_k \to
5556: \cH_k \ \ \ \sum [g]\otimes \phi_g \mapsto \sum \phi_g. $$
5557:
5558: One then obtains a commutative diagram
5559: \begin{eqnarray*}
5560: \diagram \bT \otimes \cM_*(\Gamma) \dto^{\iota}\rto^{H} &
5561: \cM_*(\Gamma) \\
5562: \cA_*\otimes \cA_* \rto^{*}& \cA_* \uto^{\epsilon \otimes 1}.
5563: \enddiagram
5564: \end{eqnarray*}
5565:
5566: In \cite{CoMoHopf} Connes and Moscovici introduced a Hopf algebra $\cH_1$
5567: associated to the transverse geometry of codimension one foliations.
5568: This is the universal enveloping algebra of a Lie algebra with basis
5569: $\{ \cX,\cY,\delta_n \, n\geq 1 \}$ satisfying, for $n,k,\ell \geq 1$,
5570: \begin{equation}\label{Liebra}
5571: [\cY,\cX]=\cX, \,\,\, [\cY,\delta_n] = n\,\delta_n, \,\,\,
5572: [\cX,\delta_n] =\delta_{n+1}, \,\,\, [\delta_k,\delta_\ell]=0,
5573: \end{equation}
5574: with coproduct an algebra homomorphism $\Delta: \cH_1 \to \cH_1
5575: \otimes \cH_1$ satisfying
5576: \begin{equation}\label{coprod-H1}
5577: \begin{array}{ll}
5578: \Delta \cY=& \cY \otimes 1 + 1 \otimes \cY, \\ \Delta \delta_1 =&
5579: \delta_1 \otimes 1 + 1 \otimes \delta_1, \\ \Delta \cX =& \cX\otimes 1
5580: + 1 \otimes \cX + \delta_1 \otimes \cY,
5581: \end{array}
5582: \end{equation}
5583: antipode the anti-isomorphism satisfying
5584: \begin{equation}\label{antipode-H1}
5585: S(\cY)=-\cY, \,\,\,\, S(\cX)=-\cX+\delta_1\cY, \,\,\,\,
5586: S(\delta_1)=-\delta_1,
5587: \end{equation}
5588: and co-unit $\epsilon(h)$ the constant term of $h\in \cH_1$.
5589:
5590: The Hopf algebra $\cH_1$ acts as symmetries of the modular Hecke
5591: algebras. This is a manifestation of the general fact that,
5592: while symmetries of ordinary commutative spaces are encoded by group
5593: actions, symmetries of noncommutative spaces are given by Hopf
5594: algebras.
5595:
5596: By comparing the actions of the Hopf algebra $\cH_1$, it is possible to
5597: derive an analogy (\cf \cite{Co-Mosc1}) between the modular Hecke
5598: algebras and the crossed product algebra of the action of a discrete
5599: subgroup of ${\rm Diff}(S^1)$ on polynomial functions on the frame
5600: bundle of $S^1$.
5601:
5602: In fact, for $\Gamma$ a discrete subgroup of ${\rm Diff}(S^1)$, and
5603: $X$ a smooth compact 1-dimensional manifold, consider as in
5604: \cite{Co-Mosc1} the algebra
5605: \begin{equation}\label{Jet-alg}
5606: \cA_\Gamma = C^\infty_c (J^1_+(X)) \rtimes \Gamma,
5607: \end{equation}
5608: where $J^1_+(X)$ is the oriented 1-jet bundle. This has an action of
5609: the Hopf algebra $\cH_1$ by
5610: \begin{equation}\label{H1act1}
5611: \begin{array}{ll}
5612: \cY(f U_\phi^*)=& y_1 \frac{\partial f}{\partial y_1} U_\phi^* \\[3mm]
5613: \cX(f U_\phi^*)=& y_1 \frac{\partial f}{\partial y} U_\phi^* \\[2mm]
5614: \delta_n (f U_\phi^*)=& y_1^n \frac{d^n}{d y^n}\left( \log
5615: \frac{d\phi}{dy} \right) f U_\phi^*,
5616: \end{array}
5617: \end{equation}
5618: with coordinates $(y,y_1)$ on $ J^1_+(X)\simeq X\times \R^+$.
5619: The trace $\tau$ defined by the volume form
5620: \begin{equation}\label{trace-tau}
5621: \tau(f U_\phi^*)= \left\{ \begin{array}{ll} \int_{J^1_+(X)} f(y,y_1)
5622: \, \frac{dy\wedge dy_1}{y_1^2} & \phi=1 \\[3mm]
5623: 0 & \phi\neq 1 \end{array}\right.
5624: \end{equation}
5625: satisfies
5626: \begin{equation}\label{tau-nu}
5627: \tau(h(a)) = \nu(h) \tau(a) \, \, \, \, \forall h \in\cH_1,
5628: \end{equation}
5629: with $\nu\in \cH_1^*$ satisfying
5630: \begin{equation}\label{nu}
5631: \nu(\cY)=1, \,\,\,\,\,\, \nu(\cX)=0, \,\,\,\,\,\, \nu(\delta_n)=0.
5632: \end{equation}
5633: The twisted antipode $\tilde S=\nu * S$ satisfies $\tilde S^2=1$ and
5634: \begin{equation}\label{twistS}
5635: \tilde S(\cY)=-\cY+1, \,\,\,\,\,\, \tilde S(\cX)=-\cX +\delta_1 \cY,
5636: \,\,\,\,\,\, \tilde S(\delta_1)=-\delta_1.
5637: \end{equation}
5638:
5639:
5640: The Hopf cyclic cohomology of a Hopf algebra is another fundamental
5641: tool in noncommutative geometry, which was developed by Connes and
5642: Moscovici in \cite{CoMoHopf}. They applied it to the
5643: computation of the local index formula for tranversely
5644: hypoelliptic operators on foliations. An action of a Hopf algebra on
5645: an algebra induces a characteristic map from the Hopf cyclic
5646: cohomology of the Hopf algebra to the cyclic cohomology of the
5647: algebra, hence the index computation can be done in terms of Hopf
5648: cyclic cohomology. The periodic Hopf cyclic cohomology of the Hopf
5649: algebra of transverse geometry is related to the Gelfand-Fuchs
5650: cohomology of the Lie algebra of formal vector fields \cite{CoMoHopf2}.
5651:
5652: In the case of the Hopf algebra $\cH_1$, there are three basic
5653: cyclic cocycles, which in the original context of transverse geometry
5654: correspond, respectively, to the Schwarzian derivative, the
5655: Godbillon-Vey class, and the transverse fundamental class.
5656:
5657: In particular, the Hopf cyclic
5658: cocycle associated to the Schwarzian derivative is of the form
5659: \begin{equation}\label{delta2'}
5660: \delta_2':= \delta_2 - \frac 12 \delta_1^2
5661: \end{equation}
5662: with
5663: \begin{equation}\label{delta2'act}
5664: \delta_2' (f U_\phi^*)= y_1^2 \left\{ \phi(y); y\right\} \, f
5665: U_\phi^*
5666: \end{equation}
5667: \begin{equation}\label{schw-der}
5668: \left\{ F ; x \right\} : = \frac{d^2}{dx^2} \left( \log \frac{dF}{dx}
5669: \right) - \frac 12 \left( \frac{d}{dx} \left( \log \frac{dF}{dx}
5670: \right) \right)^2.
5671: \end{equation}
5672:
5673: The action of the Hopf algebra $\cH_1$ on the modular Hecke algebra
5674: described in \cite{Co-Mosc1} involves the natural derivation on
5675: the algebra of modular forms initially introduced by Ramanujan, which
5676: corrects the ordinary
5677: differentiation by a logarithmic derivative of the Dedekind $\eta$
5678: function,
5679: \begin{equation}\label{Heck-rep}
5680: \cX:= \frac{1}{2\pi i} \frac{d}{dz} - \frac{1}{2\pi i}
5681: \frac{d}{dz}(\log \eta^4) \,\, \cY, \ \ \
5682: \cY (f) =\frac k2 \, f, \,\,\, \forall f \in \cM_k .
5683: \end{equation}
5684: The element $\cY$ is the grading operator
5685: that multiplies by $k/2$ forms of weight $k$, viewed as sections of
5686: the $(k/2)$th power of the line bundle of 1-forms. The element
5687: $\delta_1$ acts as multiplication by a form-valued cocycle on
5688: $\GL^+_2(\Q)$, which measures the
5689: lack of invariance of the section $\eta^4 dz$.
5690: More precisely, one has the following action of
5691: $\cH_1$ (Connes--Moscovici \cite{Co-Mosc1}):
5692:
5693: \begin{thm}\label{actH1hecke}
5694: There is an action of the Hopf algebra $\cH_1$ on the modular Hecke
5695: algebra $\cA(\Gamma)$ of level $\Gamma$, induced by an action on
5696: $\cA_{G^+(\Q)}:= \cM \rtimes G^+(\Q)$, for
5697: $\cM=\varinjlim_{N\to\infty} \cM(\Gamma(N))$, of the form
5698: \begin{equation}\label{actH1GQ}
5699: \begin{array}{ll}
5700: \cY(fU^*_\gamma)= & \cY(f)\, U^*_\gamma \\[3mm]
5701: \cX(fU^*_\gamma)= & \cX(f)\, U^*_\gamma \\[2mm]
5702: \delta_n(fU^*_\gamma)= & \frac{d^n}{dZ^n} \left( \log \frac{d(Z |_0
5703: \gamma)}{dZ} \right) (dZ)^n \, fU^*_\gamma,
5704: \end{array}
5705: \end{equation}
5706: with $\cX(f)$ and $\cY(f)$ as in \eqref{Heck-rep}, and
5707: \begin{equation}\label{coordZ}
5708: Z(z) = \int_{i\infty}^z \eta^4 dz.
5709: \end{equation}
5710: \end{thm}
5711:
5712: The cocycle \eqref{delta2'} associated to the Schwarzian derivative is
5713: represented by an inner derivation of $\cA_{G^+(\Q)}$,
5714: \begin{equation}\label{delta2'heck}
5715: \delta_2'(a) = [a, \omega_4],
5716: \end{equation}
5717: where $\omega_4$ is the weight four modular form
5718: \begin{equation}\label{omega4}
5719: \omega_4 = -\frac{E_4}{72}, \,\,\,\, \text{ with } \,\,\, E_4(q)= 1 +
5720: 240 \sum_{n=1}^\infty n^3 \frac{q^n}{1-q^n}, \,\,\,\,\, q=e^{2\pi i z},
5721: \end{equation}
5722: which is expressed as a Schwarzian derivative
5723: \begin{equation}\label{Somega4}
5724: \omega_4 = (2\pi i)^{-2} \, \{ Z; z\}.
5725: \end{equation}
5726:
5727: This result is
5728: used in \cite{Co-Mosc1} to investigate perturbations of the Hopf algebra
5729: action. The freedom one has in modifying the action by
5730: a 1-cocycle corresponds exactly to the data introduced by Zagier
5731: in \cite{Zagier}, defining "canonical" Rankin-Cohen algebras,
5732: with the derivation $\partial$ and the element $\Phi$ in Zagier's
5733: notation corresponding, respectively, to the action of the generator
5734: $\cX$ on modular forms and to $\omega_4=2\Phi$.
5735:
5736: The cocycle associated to the Godbillon-Vey class is described in
5737: terms of a 1-cocycle on $\GL^+_2(\Q)$ with values in
5738: Eisenstein series of weight two, which measures the lack of
5739: $\GL^+_2(\Q)$-invariance of the connection associated to the
5740: generator $\cX$. The authors derive from this an arithmetic presentation
5741: of the rational Euler class in $H^2(\SL_2(\Q),\Q)$ in
5742: terms of generalized Dedekind sums.
5743:
5744: The cocycle associated to the transverse fundamental class, on the
5745: other hand, gives rise to a natural extension of the first
5746: Rankin-Cohen bracket \cite{Zagier} from modular
5747: forms to the modular Hecke algebras.
5748:
5749: \medskip
5750:
5751: Rankin--Cohen algebras can be treated in different perspectives:
5752: Zagier introduced them and studied them with a
5753: direct algebraic approach (\cf \cite{Zagier}). There appears to be an
5754: interesting and deep connection to vertex operator algebras, which
5755: manifests itself in a form of duality between these two types of
5756: algebras.
5757:
5758: The Rankin--Cohen brackets are a family of brackets
5759: $[f,g\,]^{(k,\ell)}_n$ for $n\geq 0$, defined for $f,g \in R$, where
5760: $R$ is a graded ring with a derivation $D$. For
5761: $R=\oplus_{k\geq 0} R_k$, $D: R_k \to R_{k+2}$, $f\in R_k$ $g\in
5762: R_\ell$, the brackets $[,]_n : R_k\otimes R_\ell \to R_{k+\ell+2}$
5763: are given by
5764: \begin{equation}\label{RCbraD}
5765: [f,g\,]^{(k,\ell)}_n = \sum_{r+s=n} (-1)^r \left(\begin{array}{c}
5766: n+k-1 \\ s \end{array}\right) \left(\begin{array}{c} n+\ell -1 \\
5767: r \end{array}\right) D^r f \, D^s g.
5768: \end{equation}
5769:
5770: These Rankin--Cohen brackets induced by $(R_*, D) \Rightarrow
5771: (R_*, [,]_*)$ give rise to a {\em standard Rankin--Cohen algebra}, in
5772: Zagier's terminology (\cf \cite{Zagier}).
5773: There is an isomorphism of categories between graded rings with a
5774: derivation and standard Rankin--Cohen algebras.
5775:
5776: In the case of Lie algebras, one can first define a {\em standard
5777: Lie algebra} as the Lie algebra associated to an associative
5778: algebra $(A,*)\Rightarrow (A,[,])$ by setting $[X,Y]= X*Y-Y*X$ and
5779: then define an abstract Lie algebra as a structure $(A,[,])$ that
5780: satisfies {\em all} the algebraic identities satisfied by a
5781: standard Lie algebras, though it is not necessarily induced by an
5782: associative algebra. It is then a theorem that the antisymmetry of
5783: the bracket and the Jacobi identity are sufficient to determine
5784: all the other algebraic identities, hence one can take these as a
5785: definition of an abstract Lie algebra.
5786:
5787: Just as in the case of Lie algebras, we can define a Rankin--Cohen
5788: algebra $(R_*, [,]_*)$ as a graded ring $R_*$ with a family of
5789: degree $2n$ brackets $[,]_n$ satisfying all the algebraic
5790: identities of the standard Rankin--Cohen algebra. However, in this
5791: case there is no simple set of axioms that implies all the
5792: algebraic identities.
5793:
5794: The motivation for this structure lies in the fact that there is a
5795: very important example of a Rankin--Cohen algebra which is in fact
5796: non-standard. The example is provided by modular forms (\cf
5797: \cite{Zagier}).
5798:
5799: If $f\in \cM_k$ is a modular form satisfying
5800: $$ f\left( \frac{az+b}{cz+d} \right) = (cz+d)^k f(z), $$
5801: then it derivative is no longer a modular form, due to the
5802: presence of the second term in
5803: $$ f' \left(\frac{az+b}{cz+d} \right) =(cz+d)^{k+2} f(z) + kc
5804: (cz+d)^{k+1} f(z).
5805: $$
5806: On the other hand, if we have $f\in \cM_k$ and $g\in \cM_\ell$,
5807: the bracket
5808: $$ [f,g](z) := \ell f'(z) g(z) - k f(z) g'(z) $$
5809: is a modular form in $\cM_{k+\ell +2}$. Similarly, we can define
5810: an $n$-th bracket $[,]_n: \cM_k \otimes \cM_\ell \to
5811: \cM_{k+\ell+2}$. Here's the first few brackets:
5812: $$ \begin{array}{ll} [f,g]_0 =& f g \\[2mm]
5813: [f,g]_1 =& k f g' - \ell f' g \\[2mm]
5814: [f,g]_2 = & \left(\begin{array}{c} k+1\\ 2\end{array}\right) f g''
5815: - (k+1) (\ell +1) f'g' + \left(\begin{array}{c} \ell +1\\
5816: 2\end{array}\right)f'' g. \end{array} $$
5817:
5818:
5819: Notice that for the graded ring of modular forms we have $\cM_*
5820: (\Gamma)\subset \cH$, where $\cH$ is the vector space $\cH = {\rm
5821: Hol}(\H)_{polyn}$ of holomorphic functions on the upper half
5822: plane $\H$ with polynomial growth. This is closed under
5823: differentiation and $(\cH, D)$ induces a standard Rankin--Cohen
5824: algebra $(\cH, [,]_*)$. The inclusion $(\cM_*,[,]_*) \subset (\cH,
5825: [,]_*)$ is not closed under differentiation but it is closed under
5826: the brackets.
5827:
5828: A way of constructing non-standard Rankin--Cohen algebras is
5829: provided by Zagier's {\em canonical construction} (\cf
5830: \cite{Zagier}). One considers
5831: here the data $(R_*,D,\Phi)$, where $R_*$ is a graded ring with a
5832: derivation $D$ and with a choice of an element $\Phi \in R_4$, the
5833: {\em curvature}. One then defines the brackets by the formula
5834: \begin{equation}\label{RCbraDPhi}
5835: [f,g]^{(k,\ell)}_n = \sum_{r+s=n} (-1)^r \left(\begin{array}{c}
5836: n+k-1 \\ s \end{array}\right) \left(\begin{array}{c} n+\ell -1 \\
5837: r \end{array}\right)\, f_r \, g_s,
5838: \end{equation}
5839: where $f_0 = f$ and
5840: \begin{equation}\label{DPhi}
5841: f_{r+1} = D f_r + r(r+1) \Phi f_{r-1}.
5842: \end{equation}
5843: The structure $(R_*,[,]_*)$ obtained this way is a Rankin--Cohen
5844: algebra (see \cite{Zagier}).
5845:
5846:
5847: There is a gauge action on the curvature $\Phi$, namely, for any
5848: $\varphi \in R_2$ the transformation $D \mapsto D'$ and $\Phi
5849: \mapsto \Phi'$ with
5850: $$ D'(f) = D(f) + k \varphi f $$
5851: for $f \in \cM_k$ and
5852: \begin{equation}\label{gaugePhi}
5853: \Phi' = \Phi + \varphi^2 - D(\varphi)
5854: \end{equation}
5855: give rise to {\em the same} Rankin--Cohen algebra. Thus, all the
5856: cases where the curvature $\Phi$ can be gauged away to zero
5857: correspond to the standard case.
5858:
5859: \medskip
5860:
5861:
5862: The modular form $\omega_4$ of \eqref{Somega4} provides
5863: the curvature element $\omega_4 = 2\Phi$, and the gauge equivalence
5864: condition \eqref{gaugePhi} can be rephrased in terms of Hopf
5865: algebras as the freedom to change the $\cH_1$ action by a
5866: cocycle. In particular (\cf \cite{Co-Mosc1}), for the specified
5867: action, the resulting Rankin--Cohen structure is canonical but not
5868: standard, in Zagier's terminology.
5869:
5870: The 1-form $dZ=\eta^4 dz$ is, up to scalars, the only holomorphic
5871: differential on the elliptic curve $E=X_{\Gamma(6)} \cong
5872: X_{\Gamma_0(36)}$ of equation $y^2 = x^3 +1$, so that
5873: $dZ=\frac{dx}{y}$ in Weierstrass coordinates.
5874:
5875: \medskip
5876:
5877: The Rankin--Cohen brackets on modular forms can be extended to
5878: brackets $RC_n$ on the modular Hecke algebra, defined in terms of the
5879: action of the Hopf algebra $\cH_1$ of transverse geometry. In fact,
5880: more generally, it is shown in \cite{Co-Mosc2} that it is possible to
5881: define such Rankin--Cohen brackets on any associative algebra $\cA$
5882: endowed with an action of the Hopf algebra $\cH_1$ for which there
5883: exists an element $\Omega\in \cA$ such that
5884: \begin{equation}\label{innerdelta2'}
5885: \delta_2'(a) = [ \Omega, a ], \,\,\,\, \forall a\in \cA,
5886: \end{equation}
5887: and with $\delta_2'$ as in \eqref{delta2'}, and
5888: \begin{equation}\label{deltan0}
5889: \delta_n(\Omega)=0, \,\,\, \forall n\geq 1.
5890: \end{equation}
5891:
5892: \smallskip
5893:
5894: Under these hypotheses, the following result holds (Connes--Moscovici
5895: \cite{Co-Mosc2}):
5896:
5897: \begin{thm}\label{RCdef}
5898: Suppose given an associative algebra $\cA$ with an action of the Hopf
5899: algebra $\cH_1$ satisfying the conditions \eqref{innerdelta2'} and
5900: \eqref{deltan0}.
5901: \begin{enumerate}
5902: \item There exists Rankin--Cohen brackets $RC_n$ of the form
5903: \begin{equation}\label{RCnH1}
5904: RC_n(a,b)= \sum_{k=0}^n \frac{A_k}{k!} (2\cY+k)_{n-k}(a)\,
5905: \frac{B_{n-k}}{(n-k)!} (2\cY+n-k)_k (b),
5906: \end{equation}
5907: with $(\alpha)_r=\alpha(\alpha+1)\cdots(\alpha +r-1)$ and the
5908: coefficients $A_{-1}=0$, $A_0=1$, $B_0=1$, $B_1=\cX$,
5909: $$ A_{n+1}= S(\cX) A_n -n \Omega^0 \left(\cY - \frac{n-1}{2}\right)
5910: A_{n-1}, $$
5911: $$ B_{n+1}= \cX B_n - n \Omega \left(\cY-\frac{n-1}{2}\right) B_{n-1},
5912: $$
5913: and $\Omega^0$ the right multiplication by $\Omega$.
5914: \item When applied to the modular Hecke algebra $\cA(\Gamma)$, with
5915: $\Omega= \omega_4 = 2\Phi$, the
5916: above construction yields brackets \eqref{RCnH1} that are completely
5917: determined by their restriction to modular forms where they agree with
5918: the Rankin--Cohen brackets \eqref{RCbraDPhi}.
5919: \item The brackets \eqref{RCnH1} determine associative deformations
5920: \begin{equation}\label{assoc-def}
5921: a*b = \sum_n \hbar^n \, RC_n(a,b).
5922: \end{equation}
5923: \end{enumerate}
5924: \end{thm}
5925:
5926:
5927: \medskip
5928:
5929: The first of the steps described in Section \ref{road}, namely
5930: resolving the diagonal in $\cA(\Gamma)$, is not yet
5931: done for the modular Hecke algebras and should shed light on the
5932: important number theoretic problem of the interrelation of the Hecke
5933: operators with the algebra structure given by the pointwise product.
5934:
5935: The algebra $\cA(\Gamma)$ is deeply related to the algebra of the
5936: space of two dimensional $\Q$-lattices of Section \ref{Qlatt}.
5937:
5938:
5939:
5940: \section{Noncommutative moduli spaces, Shimura
5941: varieties}\label{shimura}
5942:
5943:
5944: It appears from the study of some significant cases that an important
5945: source of interesting noncommutative spaces is provided by the
5946: ``boundary'' of classical (algebro-geometric) moduli spaces, when one
5947: takes into account the possible presence of degenerations of classical
5948: algebraic varieties that give rise to objects no longer defined within
5949: the context of algebraic varieties, but which still make sense as
5950: noncommutative spaces.
5951:
5952: \smallskip
5953:
5954: An example of algebro-geometric moduli spaces which is sufficiently
5955: simple to describe but which at the same time exhibits a very rich
5956: structure is that of the modular curves. The geometry of modular
5957: curves has already appeared behind our discussion of the 2-dimensional
5958: $\Q$-lattices and of the modular Hecke algebras, through an
5959: associated class of functions: the modular functions that appeared in
5960: our discussion of the arithmetic algebra for the quantum statistical
5961: mechanical system of 2-dimensional $\Q$-lattices and the
5962: modular forms in the modular Hecke algebras.
5963:
5964: The modular curves, quotients of the hyperbolic plane $\H$ by the
5965: action of a subgroup $\Gamma$ of finite index of $\SL_2(\Z)$, are complex
5966: algebraic curves, which admit an arithmetic structure, as they are
5967: defined over cyclotomic number fields $\Q(\zeta_N)$. They are also
5968: naturally moduli spaces. The object they parameterize are elliptic
5969: curves (with some level structure). The modular curves have an
5970: algebro-geometric compactification obtained by adding finitely many
5971: cusp points, given by the points in $\P^1(\Q)/\Gamma$. These
5972: correspond to the algebro-geometric degeneration of the elliptic
5973: curve to $\C^*$. However, in addition to these degenerations, one can
5974: consider degenerations to noncommutative tori, obtained by a limit
5975: $q\to \exp(2\pi i \theta)$ in the modulus $q=\exp(2\pi i \tau)$ of the
5976: elliptic curve, where now $\theta$ is allowed to be also irrational.
5977: The resulting boundary $\P^1(\R)/\Gamma$ is a noncommutative space (in
5978: the sense of Section \ref{quotients}). It appeared in the string
5979: theory compactifications considered in \cite{CDS}. The arithmetic
5980: properties of the noncommutative spaces $\P^1(\R)/\Gamma$ were studied
5981: in \cite{ManMar}, \cite{Mar2} \cite{Mar3}.
5982:
5983: The modular curves, for varying finite index $\Gamma\subset
5984: \SL_2(\Z)$, form a tower of branched coverings. The projective limit
5985: of this tower sits as a connected component in the more refined {\em
5986: adelic} version of the modular tower, given by the quotient
5987: \begin{equation}\label{Sh2}
5988: \GL_2(\Q)\backslash \GL_2(\A) /\C^*,
5989: \end{equation}
5990: where, as usual, $\A=\A_f\times \R$ denotes the adeles of $\Q$, with
5991: $\A_f=\hat\Z\otimes \Q$ the finite adeles.
5992:
5993: The space \eqref{Sh2} is also a moduli space. In fact, it belongs to
5994: an important class of algebro-geometric moduli spaces of great
5995: arithmetic significance, the Shimura varieties $Sh(G,X)$, where the
5996: data $(G,X)$ are given by a reductive algebraic group $G$ and a
5997: hermitian symmetric domain $X$. The pro-variety \eqref{Sh2}
5998: is the Shimura variety $Sh(\GL_2,\H^\pm)$, where $\H^\pm=\GL_2(\R)
5999: /\C^*$ is the union of upper and lower half plane in $\P^1(\C)$.
6000:
6001: \smallskip
6002:
6003: We have mentioned above that the spaces $\P^1(\R)/\Gamma$ describe
6004: degenerations of elliptic curves to noncommutative tori. This type of
6005: degeneration corresponds to degenerating a lattice $\Lambda=\Z+\Z\tau$
6006: to a pseudolattice $L=\Z+\Z\theta$ (see \cite{Manin01RM} for a detailed
6007: discussion of this viewpoint and its implications in noncommutative
6008: geometry and in arithmetic). In terms of the space \eqref{Sh2}, it
6009: corresponds to degenerating the archimedean component, namely
6010: replacing $\GL_2(\R)$ by $M_2(\R)^\cdot=M_2(\R)\smallsetminus \{ 0
6011: \}$, nonzero $2\times 2$-matrices. However, when one is working with
6012: the adelic description as in \eqref{Sh2}, one can equally consider the
6013: possibility of degenerating a lattice at the non-archimedean
6014: components. This brings back directly the notion of $\Q$-lattices of
6015: Section \ref{Qlatt}.
6016:
6017: In fact, it was shown in \cite{CMR2} that the notions of 2-dimensional
6018: $\Q$-lattices and commensurability can be reformulated in terms of
6019: Tate modules of elliptic curves and isogeny. In these terms, the space
6020: of $\Q$-lattices corresponds to non-archimedean degenerations of the
6021: Tate module, which correspond to the ``bad quotient''
6022: \begin{equation}\label{Sh2nc}
6023: \GL_2(\Q)\backslash M_2(\A_f)\times \GL_2(\R) /\C^*.
6024: \end{equation}
6025: The combination of these two types of degenerations yields a
6026: ``noncommutative compactification'' of the Shimura variety
6027: $Sh(\GL_2,\H^\pm)$ which is the algebra of the ``bad quotient''
6028: \begin{equation}\label{Sh2ncC}
6029: \GL_2(\Q)\backslash M_2(\A)^\cdot /\C^*,
6030: \end{equation}
6031: where $M_2(\A)^\cdot$ are the elements of $M_2(\A)$ with nonzero
6032: archimedean component. One recovers the Shimura variety
6033: $Sh(\GL_2,\H^\pm)$ as the set of classical points (extremal KMS states
6034: at zero temperature) of the quantum statistical mechanical system
6035: associated to the noncommutative space \eqref{Sh2nc} (\cf \cite{CoMa},
6036: \cite{CMR2}).
6037:
6038: \medskip
6039:
6040: More generally, Shimura varieties are moduli spaces for certain
6041: types of motives or as moduli spaces of Hodge structures (\cf \eg
6042: \cite{Mil2}). A Hodge structure $(W,h)$ is a pair of a finite
6043: dimensional $\Q$-vector space $W$ and a homomorphism $h:\bS\to
6044: \GL(W_\R)$, of the real algebraic group
6045: $\bS=\Res_{\C/\R}\bG_m$, with
6046: $W_\R=W\otimes \R$. This determines a decomposition
6047: $W_\R\otimes \C=\oplus_{p,q} W^{p,q}$ with
6048: $\overline{W^{p,q}}=W^{q,p}$ and $h(z)$ acting on $W^{p,q}$ by
6049: $z^{-p}\bar z^{-q}$. This gives a Hodge filtration and a weight
6050: filtration $W_\R=\oplus_k W_k$ where $W_k=\oplus_{p+q=m} W^{p,q}$.
6051: The Hodge structure $(W,h)$ has weight $m$ if $W_\R=W_m$. It is
6052: rational if the weight filtration is defined over $\Q$. A Hodge
6053: structure of weight $m$ is polarized if there is a morphism of
6054: Hodge structures $\psi:W\otimes W\to \Q(-m)$, such that
6055: $(2\pi i)^m\psi(\cdot, h(i)\cdot)$ is symmetric and positive
6056: definite. Here $\Q(m)$ is the rational Hodge structure of weight
6057: $-2m$, with $W=(2\pi i)^m \Q$ with the action $h(z)=(z\bar z)^m$.
6058: For a rational $(W,h)$, the subspace of $W\otimes \Q(m)$ fixed by
6059: the $h(z)$, for all $z\in \C^*$, is the space of ``Hodge cycles''.
6060:
6061: One can then view Shimura varieties $Sh(G,X)$ as moduli spaces of
6062: Hodge structures in the following way. Let $(G,X)$ be a Shimura datum
6063: and $\rho:G \to \GL(V)$ a faithful representation. Since $G$ is
6064: reductive, there is a finite family of tensors $\tau_i$ such that
6065: \begin{equation}\label{tensors}
6066: G=\{ g\in \GL(V): g\tau_i=\tau_i \}.
6067: \end{equation}
6068: A point $x\in X$ is by construction a $G(\R)$ conjugacy class of
6069: morphisms $h_x: \bS\to G$, with suitable properties.
6070:
6071: Consider data of the form $((W,h),\{ s_i \},\phi)$, where
6072: $(W,h)$ is a rational Hodge structure, $\{ s_i \}$ a finite
6073: family of Hodge cycles, and $\phi$ a $K$-level structure, for some
6074: $K\subset G(\A_f)$, namely a $K$-orbit of $\A_f$-modules {\em
6075: isomorphisms} $\phi: V(\A_f) \to W(\A_f)$, which maps $\tau_i$ to
6076: $s_i$. Isomorphisms of such data are isomorphisms $f:W\to W'$ of
6077: rational Hodge structures, sending $s_i\mapsto s_i'$,
6078: and such that $f\circ \phi=\phi' k$, for some $k\in K$.
6079:
6080: We assume that there exists an isomorphism of $\Q$-vector spaces
6081: $\beta:W\to V$ mapping $s_i \mapsto \tau_i$ and $h$ to $h_x$, for some
6082: $x\in X$.
6083:
6084: One denotes by ${\rm Hodge}(G,X,K)$ the set data $((W,h),\{ s_i \},\phi)$.
6085: The Shimura variety $$Sh_K(G,X)=G(\Q)\backslash X\times
6086: G(\A_f)/K$$ is the moduli space of the isomorphism classes of data
6087: $((W,h),\{ s_i \},\phi)$, namely there is a map of ${\rm Hodge}(G,X,K)$
6088: to $Sh_K(G,X)$ (seen over $\C$), that descends to a bijection on
6089: isomorphism classes ${\rm Hodge}(G,X,K)/\sim$.
6090:
6091: In such cases also one can consider degenerations of these data, both
6092: at the archimedean and at the nonarchimedean components. One then
6093: considers data $((W,h),\{ s_i \}, \phi,\tilde\beta)$, with
6094: a non-trivial {\em homomorphism} $\tilde\beta: W\to V$, which is a
6095: morphism of Hodge structures, and such that $\tilde\beta(\ell_{s_i})\subset
6096: \ell_{\tau_i}$. This yields noncommutative spaces inside which the
6097: classical Shimura variety sits as the set of classical points.
6098:
6099: Quantum statistical machanical systems associated to Shimura varieties
6100: have been recently studied by Ha and Paugam in \cite{HaPa}. Given a
6101: faithful representation $\rho:G\to \GL(V)$ as above, there is an
6102: ``enveloping semigroup'' $M$, that is, a normal irreducible semigroup
6103: $M\subset \End(V)$ such that $M^\times =G$. Such semigroup can be used
6104: to encode the degenerations of the Hodge data described above.
6105: The data $(G,X,V,M)$ then determine a noncommutative space which
6106: describes the ``bad quotient'' $Sh^{nc}_K(G,X)=G(\Q)\backslash X\times
6107: M(\A_f)$ and is a moduli space for the
6108: possibly non-invertible data $((W,h),\{ s_i \},\phi)$.
6109: Its ``set of classical points'' is the Shimura variety
6110: $Sh(G,X)$. The actual construction of the algebras involves
6111: some delicate steps, especially to handle the presence of stacky
6112: singularities (\cf \cite{HaPa}).
6113:
6114:
6115: \section{The adeles class space and the spectral realization}\label{adeles}
6116:
6117:
6118:
6119: In this section we describe a noncommutative space, the adele
6120: class space $X_\K$, associated to any global field $\K$, which leads
6121: to a spectral realization of the zeros of the Riemann zeta function
6122: for $\K=\Q$ and more generally of $L$-functions associated to Hecke
6123: characters. It also gives a geometric interpretation of the
6124: Riemann-Weil explicit formulas of number theory as a trace formula.
6125: This space is closely related for $\K=\Q$ with the space of
6126: commensurability classes of $\Q$-lattices described above. Rather
6127: than starting directly with its description we first put the
6128: problem of finding the geometry of the set of prime numbers in the
6129: proper perspective.
6130:
6131: \medskip
6132: {\bf The set of primes}
6133: \medskip
6134:
6135: One of the main problems of arithmetic is to understand the
6136: distribution of the set of prime numbers
6137: $$
6138: \{2, 3, 5, 7, 11, 13, 17, 19, 23, 29, 31, 37, 41, 43, 47, 53, 59,
6139: 61, 67, 71, \ldots \}
6140: $$
6141: as a subset of the integers. To that effect one introduces the
6142: counting function
6143: $$
6144: \pi(n)=\,{\rm number}\; {\rm of}\; {\rm primes}\; p\leq n
6145: $$
6146: The problem is to understand the behavior of $\pi(x)$ when $x\to
6147: \infty$. One often hears that the problem comes from the lack of a
6148: ``simple" formula for $\pi(x)$. This is not really true and for
6149: instance in 1898 H. Laurent \cite{laurent} gave the following
6150: formula whose validity is an easy exercise in arithmetic,
6151: \begin{small}
6152: $$
6153: \pi(n)=\, 2+\,\sum_{k=5}^n\,\frac{e^{2\pi i \Gamma(k)/k}-1}{e^{-2\pi i
6154: /k}-1} \,,
6155: $$
6156: \end{small}
6157: where $ \Gamma(k)=(k-1)!$ is the Euler Gamma function.
6158:
6159: \begin{figure}
6160: \begin{center}
6161: \includegraphics[scale=0.80]{pin3.eps}
6162: \end{center}
6163: \caption{Graphs of $\pi(x)$ and Li$(x)$ \label{pili1}}
6164: \end{figure}
6165:
6166: The problem with this formula is that it has no bearing on the
6167: asymptotic expansion of $\pi(x)$ when $x\to \infty$. Such an
6168: expansion was guessed by Gauss in the following form,
6169: $$
6170: \pi(x)=\, \int_0^x\,\frac{du}{\log(u)}\,+\;R(x)
6171: $$
6172: where the integral logarithm admits the asymptotic expansion,
6173: $$Li(x) =\int_0^x\,\frac{du}{\log(u)}
6174: \sim \sum \,(k-1)!\,\frac{x}{\log(x)^k}$$
6175:
6176: \begin{figure}
6177: \begin{center}
6178: \includegraphics[scale=0.80]{pin4.eps}
6179: \end{center}
6180: \caption{Graph of $\pi(x)-$ Li$(x)$ \label{pili2}}
6181: \end{figure}
6182:
6183: The key issue then is the size of the remainder $R(x)$.
6184:
6185: \medskip
6186: {\bf The Riemann hypothesis}
6187: \medskip
6188:
6189: It asserts that this size is governed by
6190: \begin{equation}\label{rh}
6191: R(x)=\,O(\sqrt{x}\,\log(x)) .
6192: \end{equation}
6193: The Riemann Hypothesis is in fact a conjecture on the zeros of the
6194: zeta function
6195: \begin{equation}\label{zeta}
6196: \zeta(s)=\,\sum_1^\infty\,n^{-s} ,
6197: \end{equation}
6198: whose definition goes back to Euler,
6199: who showed the fundamental factorization
6200: \begin{equation}\label{eulerprod}
6201: \zeta(s)=\,\prod_{\mathcal P}\,(1-\,p^{-s})^{-1}\,.
6202: \end{equation}
6203:
6204: It extends to a meromorphic function in the whole complex plane $\C$
6205: and fulfills the functional equation
6206: \begin{equation}\label{functequ}
6207: \pi^{-s/2}\,\Gamma(s/2)\,\zeta(s)=\,\pi^{-(1-s)/2}\,\Gamma((1-s)/2)\,\zeta(1-s),
6208: \end{equation}
6209: so that the function
6210: \begin{equation}\label{zetaq}
6211: \zeta_\Q(s)=\pi^{-s/2}\,\Gamma(s/2)\,\zeta(s)
6212: \end{equation}
6213: admits the symmetry $s\mapsto 1-s$.
6214:
6215: The Riemann conjecture asserts that all zeros of $\zeta_\Q$ are on
6216: the critical line $\frac{1}{2} +\,i\,\R$. The reason why the
6217: location of the zeros of $\zeta_\Q$ controls the size of the
6218: remainder in \eqref{rh} is the explicit formulas that relate primes
6219: with these zeros. Riemann proved the following first instance of an
6220: ``explicit formula"
6221: \begin{equation}\label{explicit1}
6222: \pi'(x)=\,Li(x)-\,\sum_\rho\,Li(x^\rho)+\,\int_x^\infty\,\frac{1}{u^2-1}\;\frac{du}{u
6223: \log u}-\,\log 2
6224: \end{equation}
6225: where the sum is over the non-trivial (\ie complex) zeros of the
6226: zeta function, and where the function $\pi'(x)$ is given by
6227: $$
6228: \pi'(x)=\,\pi(x)\,+\,\,\frac{1}{2}\,\pi(x^{\frac{1}{2}})+
6229: \,\frac{1}{3}\,\pi(x^{\frac{1}{3}})\,+\cdots
6230: $$
6231: which gives by the Moebius inversion formula
6232: $$
6233: \pi(x)=\,\sum \,\mu(m)\,\frac{1}{m}\,\pi'(x^{\frac{1}{m}}) .
6234: $$
6235:
6236: \bigskip
6237: \begin{figure}
6238: \begin{center}
6239: \includegraphics[scale=0.80]{rszeta.eps}
6240: \end{center}
6241: \caption{Zeros of zeta \label{zeros}}
6242: \end{figure}
6243:
6244:
6245: \medskip
6246: {\bf The generalized Riemann hypothesis}
6247: \medskip
6248:
6249: The explicit formulas of Riemann were put in more modern form by A.
6250: Weil, as
6251: \begin{equation}\label{explicit}
6252: \,\wh h(0) + \wh h(1) -\sum_\rho\,\wh h(\rho)=\,\sum_{v}
6253: \int'_{\K^*_v} \, \frac{h(u^{-1})}{
6254: |1-u|}\, d^* u \, ,
6255: \end{equation}
6256: where $\K$ is now an arbitrary global field, $v\in \Sigma_\K$ varies
6257: among the places of $\K$ and the integral is taking place over the
6258: locally compact field $\K_v$ obtained by completion of $\K$ at the
6259: place $v$. Also $\int'$ is the pairing with the distribution on
6260: $\K_v$ which agrees with ${du \over \vert 1-u \vert}$ for $u \not= 1$
6261: and whose Fourier transform (relative to a selfdual choice of
6262: additive characters $\alpha_v$) vanishes at $1$. By definition a
6263: {\em global field} is a (countable) discrete cocompact subfield in a
6264: locally compact ring. This ring depends functorially on $\K$ and is
6265: called the ring $\A_\K$ of adeles of $\K$. The quotient
6266:
6267: \begin{equation}\label{ideleclass}
6268: C_\K=\,{\rm GL}_1(\A_\K)/{\rm GL}_1(\K)
6269: \end{equation}
6270: is the locally compact group of idele classes of $\K$ which plays a
6271: central role in class field theory. In Weil's explicit formula the
6272: test function $h$ is in the Bruhat-Schwartz space $\cS(C_\K)$. The
6273: multiplicative groups ${\rm GL}_1(\K_v)=\K_v^*$ are embedded
6274: canonically as cocompact subgroups of $C_\K$. The sum on the left
6275: hand side takes place over the zeros of $L$-functions associated to
6276: Hecke characters. The function $\wh h$ is the Fourier transform of
6277: $h$. The generalized Riemann conjecture asserts that all the zeros
6278: of these $L$-functions are on the critical line $\frac{1}{2}
6279: +\,i\,\R$. This was proved by Weil when the global field $\K$ has
6280: non-zero characteristic, but remains open in the case when $\K$ is of
6281: characteristic zero, in which case it is a number field \ie a finite
6282: algebraic extension of the field $\Q$ of rational numbers.
6283:
6284:
6285: \medskip
6286: {\bf
6287: Quantum Chaos $\to$ Riemann Flow ? }
6288: \medskip
6289:
6290: For $E>0$ let $N(E)$ be the number of zeros of the Riemann zeta
6291: function $\zeta_\Q$ whose imaginary parts are in the open interval
6292: $]0,E[$. Riemann proved that the step function $N(E)$ can be written
6293: as the sum
6294: $$N(E)=\, \langle N(E) \rangle + N_{\rm osc} (E)$$
6295: of a smooth approximation $ \langle N(E) \rangle$ and a purely
6296: oscillatory function $N_{\rm osc} (E)$ and gave the following
6297: explicit form
6298:
6299: \begin{equation}\label{riemcount} \langle N(E) \rangle = \,\frac{E}{2\,\pi}\,(\log
6300: \frac{E}{2\,\pi} -1) + \frac{7}{8} +\,o(1)
6301: \end{equation}
6302:
6303: for the smooth approximation.
6304:
6305: There is a striking analogy between the behavior of the step
6306: function $N(E)$ and that of the function counting the number of
6307: eigenvalues of the Hamiltonian $H$ of the quantum system obtained
6308: after quantization of a chaotic dynamical system, which is at
6309: centerstage in the theory of {\em quantum chaos}. A comparison of
6310: the asymptotic expansions of the oscillatory terms in both cases,
6311: namely
6312:
6313: $$N_{\rm osc} (E) \sim \frac{1}{ \pi} \sum_{\gamma_p} \sum_{m=1}^{\infty}
6314: \frac{1}{ m} \, \frac{1}{ 2{\rm sh} \left( \frac{m\lambda_p }{
6315: 2}\right)} \, \sin (m \, E \, T_{\gamma}^{\#})
6316: $$
6317:
6318: for the quantization of a chaotic dynamical system, and
6319:
6320:
6321: $$
6322: N_{\rm osc} (E) \sim \frac{-1 }{ \pi} \sum_p \sum_{m=1}^{\infty}
6323: \frac{1}{ m} \, \frac{1 }{ p^{m/2}} \, \sin \, (m \, E \, \log \, p)
6324: \,
6325: $$
6326:
6327: for the Riemann zeta function, gives precious indications on the
6328: hypothetical {\em Riemann flow} that would make it possible to identify the
6329: zeros of zeta as the spectrum of an Hamiltonian. For instance the
6330: periodic orbits of the flow should be labeled by the prime numbers
6331: and the corresponding periods $T_p$ should be given by the $\log p$.
6332: However a closer look reveals an overall minus sign that forbids any
6333: direct comparison.
6334:
6335:
6336: \bigskip
6337: \begin{figure}
6338: \begin{center}
6339: \includegraphics[scale=0.90]{graph1.eps}
6340: \end{center}
6341: \caption{Counting zeros of zeta \label{zeros1}}
6342: \end{figure}
6343:
6344:
6345: \medskip
6346: {\bf
6347: Spectral realization as an absorption spectrum}
6348: \medskip
6349:
6350:
6351: The above major sign obstruction was bypassed in \cite{CoRH} using
6352: the following basic distinction between observed spectra in physics.
6353: When the light coming from a hot chemical element is decomposed
6354: through a prism, it gives rise to bright emission lines on a dark
6355: background, and the corresponding frequencies are a signature of its
6356: chemical composition. When the light coming from a distant star is
6357: decomposed through a prism, it gives rise to dark lines, called
6358: absorption lines, on a white background. The spectrum of the light
6359: emitted by the sun was the first observed example of an absorption
6360: spectrum. In this case the absorption lines were discovered by
6361: Fraunhofer. The chemicals in the outer atmosphere of the star absorb
6362: the corresponding frequencies in the white light coming from the
6363: core of the star.
6364:
6365: \bigskip
6366: \begin{figure}
6367: \begin{center}
6368: \includegraphics[scale=0.8]{spectres7.eps}
6369: \end{center}
6370: \caption{The two kinds of Spectra \label{spectra}}
6371: \end{figure}
6372:
6373: The simple idea then is that, because of the minus sign above, one
6374: should look for the spectral realization of the zeros of zeta not as
6375: a usual emission spectrum but as an absorption spectrum. Of course
6376: by itself this idea does not suffice to get anywhere since one needs
6377: the basic dynamical system anyway. The adele class space, namely the
6378: quotient
6379: $$
6380: X_\K=\A_\K/ \K^*
6381: $$
6382: introduced in \cite{CoRH}, does the job as shown there. The action of
6383: the idele class group $C_\K$ on the adele class space is simply given
6384: by multiplication. In particular the idele class group $C_\K$ acts on
6385: the suitably defined Hilbert space $L^2(X_\K)$ and the zeros of
6386: $L$-functions give the absorption spectrum, with non-critical zeros
6387: appearing as resonances.
6388:
6389: \medskip
6390:
6391: Exactly as adeles, the adele class space $X_\K$ involves all the
6392: places of $\K$. If in order to simplify, one restricts to a finite
6393: set of places one still finds a noncommutative space but one can
6394: analyze the action of the analogue of $C_\K$ and compute its trace
6395: after performing a suitable cutoff (necessary in all cases to see
6396: the missing lines of an absorption spectrum). One gets a trace
6397: formula
6398: $$
6399: {\rm Trace} \, (R_{\Lambda} \, U(h)) = 2h (1) \log' \Lambda +
6400: \sum_{v \in S} \int'_{\K^*_v} \frac{h(u^{-1})}{ \vert 1-u \vert} \,
6401: d^* u + o(1),
6402: $$
6403: where the terms in the right hand side are exactly the same as in Weil's explicit
6404: formula \eqref{explicit}. This is very encouraging since at least it
6405: gives geometric meaning to the complicated terms of \eqref{explicit}
6406: as the contributions of the periodic orbits to the computation of
6407: the trace.
6408:
6409: In particular it gives a perfect interpretation of the smooth
6410: function $ \langle N(E) \rangle $ approximating the counting $N(E)$
6411: of the zeros of zeta, from counting the number of states of the
6412: one dimensional quantum system with Hamiltonian
6413: $$
6414: h (q, p) = \, 2\pi q\,p
6415: $$
6416: which is just the generator of the scaling group. Indeed the
6417: function $\frac{E}{2\,\pi}\,(\log \frac{E}{2\,\pi} -1)$ of Riemann
6418: formula \eqref{riemcount} appears as the number of missing degrees
6419: of freedom in the number of quantum states for the above system, as
6420: one obtains from the simple computation of the area of the region
6421: $$
6422: B_+=\,\{(p,q)\in[0,\Lambda]^2\,;\,h(p,q)\leq E\}
6423: $$
6424:
6425: $${\rm Area}(B_+)=\frac{E}{2\pi} \times 2 \log \Lambda\,
6426: - \frac{E}{2\pi} \left( \log \frac{E}{2\pi} - 1 \right)
6427: $$
6428:
6429: while the term $\frac{E}{2\pi} \times 2 \log \Lambda$ corresponds to
6430: the number of degrees of freedom of white light. A careful
6431: computation gives not only the correction term of $\frac{7}{8}$ in
6432: \eqref{riemcount} but all the remaining $o(1)$ terms.
6433:
6434:
6435:
6436: \bigskip
6437: \begin{figure}
6438: \begin{center}
6439: \includegraphics[scale=0.8]{gr0.eps}
6440: \end{center}
6441: \caption{Counting quantum states}\label{countingstates}
6442: \end{figure}
6443:
6444:
6445: Finally it was shown in \cite{CoRH} that the generalized Riemann
6446: hypothesis is equivalent to the validity of a {\em global} trace
6447: formula, but this is only one of many equivalent reformulations of the
6448: Riemann hypothesis.\footnote{There is a running joke, inspired
6449: by the European myth of Faust, about a mathematician trying to
6450: bargain with the devil for a proof of the Riemann hypothesis ...}
6451:
6452: \medskip
6453:
6454:
6455:
6456: \section{Thermodynamics of endomotives and the Tehran program}\label{program}
6457:
6458: In many ways the great virtue of a problem like RH comes from the
6459: developments that it generates. At first sight it does not appear as
6460: having any relation with geometry, and its geometric nature
6461: gradually emerged in the twentieth century mainly because of the
6462: solution of Weil in the case of global fields of positive
6463: characteristic.
6464:
6465: We outline a program, current joint
6466: work of Katia Consani and the two authors, to adapt Weil's proof for
6467: the case of global fields of positive characteristic to the case of
6468: number fields.\footnote{This program was first announced in a
6469: lecture at IPM Tehran in September 2005, hence we refer to it as ``the Tehran
6470: program''.}
6471:
6472:
6473:
6474: \medskip
6475: {\bf
6476: Function Fields}
6477: \medskip
6478:
6479: Given a global field $\K$ of positive characteristic, there exists a
6480: finite field $\F_q$ and a smooth projective curve $C$ defined over
6481: $\F_q$ such that $\K$ is the field of $\F_q$-valued rational
6482: functions on $C$. The analogue (Artin, Hasse, Schmidt) of the zeta
6483: function is
6484: $$
6485: \zeta_\K(s)=\,\prod_{v\in\Sigma_K}\,(1-q^{-f(v)s})^{-1}
6486: $$
6487: where $\Sigma_\K$ is the set of places of $\K$ and $f(v)$ is the
6488: degree (see below) of the place $v\in \Sigma_\K$.
6489:
6490: The functional equation takes the form
6491: $$
6492: q^{(g-1)(1-s)}\,\zeta_\K(1-s)=\,q^{(g-1)s} \,\zeta_\K(s)
6493: $$
6494: where $g$ is the genus of $C$.
6495:
6496: The analogue of the Riemann conjecture for such global fields was
6497: proved by Weil (1942) who developed algebraic geometry in that
6498: context. Weil's proof rests on two steps.
6499:
6500: \begin{itemize}
6501:
6502: \item (A) Explicit Formula
6503: \medskip
6504:
6505: \item (B) Positivity
6506:
6507:
6508: \end{itemize}
6509:
6510:
6511: Both are based on the geometry of the action of the Frobenius on the
6512: set $C(\bar \F_q)$ of points of $C$ over an algebraic closure $\bar
6513: \F_q$ of $\F_q$. This set $C(\bar \F_q)$ maps canonically to the
6514: set $\Sigma_\K$ of places of $\K$ and the degree of a place $v\in
6515: \Sigma_\K$ is the number of points in the orbit of the
6516: Frobenius acting on the fiber of the projection
6517: $$
6518: C(\bar \F_q) \to \Sigma_\K\,.
6519: $$
6520: The analogue
6521: $$\#\{C(\F_{q^j})\}=\,\sum\,(-1)^k\,
6522: {\rm Tr}({\rm Fr}^{* j}|H^k_{{\rm et}}(\bar C,\Q_\ell))$$
6523: of the Lefschetz fixed point formula makes it possible to compute
6524: the number $\#\{C(\F_{q^j})\}$ of points with coordinates in the
6525: finite extension $\F_{q^j}$ from the action of ${\rm Fr}^*$ in the
6526: etale cohomology group $H^1_{{\rm et}}(\bar C,\Q_\ell)$, which does
6527: not depend upon the choice of the $\ell$-adic coefficients
6528: $\Q_\ell$.
6529:
6530: This shows that the zeta function is a rational fraction
6531: $$
6532: \zeta_\K(s)=\,\frac{P(q^{-s})}{(1-q^{-s})(1-q^{1-s})}
6533: $$
6534:
6535: where the polynomial $P$ is the characteristic polynomial of the
6536: action of ${\rm Fr}^*$ in $H^1$.
6537:
6538:
6539:
6540: The analogue of the Riemann conjecture for global fields of
6541: characteristic $p$ means that its eigenvalues \ie the complex
6542: numbers $\lambda_j$ of the factorization
6543: $$
6544: P(T)=\,\prod \,(1-\lambda_j\, T)
6545: $$
6546: are of modulus $|\lambda_j|=q^{1/2}$.
6547:
6548: The main ingredient in the proof of Weil is the notion of
6549: correspondence, given by divisors in $C\times C$. They can be viewed
6550: as multivalued maps,
6551: $$Z\;:\;C\to C,\ \ \ \ P\mapsto Z(P).$$
6552:
6553: Two correspondences are equivalent if
6554: they differ by a principal divisor,
6555: $$
6556: U\sim V \Leftrightarrow U-V=\,(f)
6557: $$
6558: The composition of correspondences is
6559: $$
6560: Z=\,Z_1 \star Z_2 \,, \ \ \ Z_1 \star Z_2(P)=\,Z_1 ( Z_2(P))
6561: $$
6562: and the adjoint is given using the transposition
6563: $\sigma(x,y)=\,(y,x)$ by
6564: $$
6565: Z'=\, \sigma(Z)\,.
6566: $$
6567: The degree $d(Z)$ of a correspondence is defined, independently of a
6568: generic point $P\in C$ by,
6569: $$
6570: d(Z)=\,Z \,\bullet \,(P\times C)\,,
6571: $$
6572: where $\bullet$ is the intersection number. One has a similar
6573: definition of the codegree
6574: $$ \quad d^{\,'}(Z)=\,Z
6575: \,\bullet \,(C\times P) $$
6576:
6577: Weil defines the {\em Trace} of a
6578: correspondence as follows
6579: $$
6580: {\rm Trace}(Z)=\,d(Z)+\,d^{\,'}(Z)-\,Z\bullet \Delta
6581: $$
6582: where $\Delta$ is the identity correspondence. The main step in
6583: Weil's proof is
6584:
6585: \smallskip
6586:
6587: \begin{thm} (Weil)
6588: The following positivity holds : ${\rm Trace}(Z \star Z')>0$ unless
6589: $Z$ is a trivial class.
6590: \end{thm}
6591:
6592: \smallskip
6593:
6594: Clearly
6595: if one wants to have any chance at imitating the steps of Weil's
6596: proof of RH for the case of number fields one needs to have an
6597: analogue of the points of $C(\bar \F_q)$ and the action of the
6598: Frobenius, of the etale cohomology and of the unramified extensions
6599: $\K\otimes_{\F_q}\,\F_{q^n}$ of $\K$.
6600:
6601: \medskip
6602: {\bf
6603: Endomotives and Galois action}
6604: \medskip
6605:
6606: The adele class space $X_\K$ of a global field admits a natural
6607: action of the idele class group $C_\K$ and as such is on the adelic
6608: side of the class field theory isomorphism. In order to obtain a
6609: description of this space which is closer to geometry one needs to
6610: pass to the Galois side of class field theory. In the case
6611: $\K=\Q$, it is possible to present the adele class space in a fairly simple
6612: manner not involving adeles thanks to its intimate relation with
6613: the space of $1$-dimensional $\Q$-lattices of section \ref{Qlatt}.
6614: The direct interpretation of the action of the Galois group of $\bar
6615: \Q/\Q$ on the values of fabulous states for the BC-system (section
6616: \ref{Qlatt}) then suggests that one should be able to construct
6617: directly the space $X_\Q$ with a canonical action of the Galois
6618: group of $\bar \Q/\Q$.
6619:
6620: This was done in \cite{CCM} thanks to an extension of the notion of
6621: Artin motives, called {\em endomotives}. Following Grothendieck, one
6622: can reformulate Galois theory over a field $\K$ as the equivalence of
6623: the category of reduced commutative finite dimensional algebras over
6624: $\K$ with the category of continuous actions of the Galois group $G$
6625: of $\bar \K/\K$ on finite sets. By construction the algebra of the
6626: BC-system is a crossed product of a commutative algebra $A$ by a
6627: semi-group. When working over $\K=\Q$ which is essential in the
6628: definition of fabulous states, the algebra $A$ is simply the group
6629: ring $\Q[\Q/\Z]$ of the torsion group $\Q/\Z$. Thus
6630: $$
6631: A=\, \varinjlim \,A_n \,,\quad A_n=\,\Q[\Z/n\Z]
6632: $$
6633: and we are dealing with a projective limit of Artin motives. The key
6634: point then is to keep track of the corresponding action of the
6635: Galois group $G$ of $\bar \K/\K$, $\K=\Q$. The Galois-Grothendieck
6636: correspondence associates to a reduced commutative finite
6637: dimensional algebra $B$ over $\K$ the set of characters of $B$ with
6638: values in $\bar \K$ together with the natural action of $G$. This
6639: action is non-trivial for the algebras $A_n=\,\Q[\Z/n\Z]$ where it
6640: corresponds to the cyclotomic theory.
6641:
6642: One can then recover the Bost--Connes system with its natural Galois symmetry
6643: in a conceptual manner which extends to the general context of
6644: semigroup actions on projective systems of Artin motives. These
6645: typically arise from self-maps of algebraic varieties.
6646: Given a pointed algebraic
6647: variety $(Y,y_0)$ over a field $\K$ and a countable unital abelian
6648: semigroup $S$ of finite endomorphisms of $(Y,y_0)$, unramified over
6649: $y_0\in Y$ one constructs a projective system of Artin motives $X_s$
6650: over $\K$ from these data as follows. For $s\in S$, one sets
6651: \begin{equation}\label{Xs}
6652: X_s=\{ y\in Y:\, s(y)=y_0 \}.
6653: \end{equation}
6654: For a pair $s,s'\in S$, with $s'=sr$, the map $\xi_{s',s}: X_{sr}\to
6655: X_s$ is given by
6656: \begin{equation}\label{Xsr}
6657: X_{sr} \ni y \mapsto r(y)\in X_s.
6658: \end{equation}
6659: This defines a projective system indexed by the semigroup $S$ itself
6660: with partial order given by divisibility. We let $X=\varprojlim_s
6661: X_s$.
6662:
6663: Since $s(y_0)=y_0$, the base point $y_0$ defines a component $Z_s$
6664: of $X_s$ for all $s\in S$. Let $\xi_{s',s}^{-1}(Z_s)$ be the inverse
6665: image of $Z_s$ in $X_{s'}$. It is a union of components of $X_{s'}$.
6666: This defines a projection $e_s$ onto an open and closed subset
6667: $X^{e_s}$ of the projective limit $X$. One then shows (\cite{CCM})
6668: that the semigroup $S$ acts on the projective limit $X$ by partial
6669: isomorphisms $\rho_s: X \to X^{e_s}$ defined by the property that
6670: \begin{equation}\label{endoS}
6671: \xi_{su}(\rho_s(x))=\xi_u(x), \, \forall u\in S, \forall x\in X.
6672: \end{equation}
6673:
6674: The BC-system is obtained from the pointed algebraic variety
6675: $(\bG_m(\Q),1)$ where the affine group scheme $\bG_m$ is the
6676: multiplicative group. The semigroup $S$ is the semigroup of non-zero
6677: endomorphisms of $\bG_m$. These correspond to maps of the form
6678: $u\mapsto u^n$ for some non-zero $n\in \Z$, and one restricts to
6679: $n\in \N^*$.
6680:
6681: In this class of examples one has an ``equidistribution'' property,
6682: by which the uniform normalized counting measures $\mu_s$ on $X_s$
6683: are compatible with the projective system and define a probability
6684: measure on the limit $X$. Namely, one has
6685: \begin{equation}\label{measlim}
6686: \xi_{s',s} \mu_s = \mu_{s'},\ \ \ \ \forall s,s'\in S.
6687: \end{equation}
6688: This follows from the fact that the number of preimages of a point
6689: under $s\in S$ is equal to $\deg s$. This provides exactly the data
6690: which makes it possible to perform the thermodynamical analysis of such
6691: endomotives. This gives a rather unexplored new territory since even
6692: the simplest examples beyond the BC-system remain to be
6693: investigated. For instance let $Y$ be an elliptic curve defined over
6694: $\K$. Let $S$ be the semigroup of non-zero endomorphisms of $Y$.
6695: This gives rise to an example in the general class described above.
6696: When the elliptic curve has complex multiplication, this gives rise
6697: to a system which, in the case of a maximal order, agrees with the
6698: one constructed in \cite{CMR}. In the case without complex
6699: multiplication, this provides an example of a system where the
6700: Galois action does not factor through an abelian quotient.
6701:
6702:
6703:
6704: \medskip
6705: {\bf
6706: Frobenius as dual of the time evolution}
6707: \medskip
6708:
6709: The Frobenius is such a universal symmetry in characteristic $p$,
6710: owing to the linearity of the map $x\mapsto x^p$ that it is very hard to
6711: find an analogue of such a far reaching concept in characteristic
6712: zero. As we now explain, the classification of type III
6713: factors provides the basic ingredient which when combined with
6714: cyclic cohomology makes it possible to analyze the thermodynamics of a
6715: noncommutative space and get an analogue of the action of the
6716: Frobenius on etale cohomology.
6717:
6718:
6719:
6720: The key ingredient is that noncommutativity generates a time
6721: evolution at the ``measure theory" level. While it had been long
6722: known by operator
6723: algebraists that the theory of
6724: von-Neumann algebras represents
6725: a far reaching extension of measure theory, the
6726: main surprise which occurred at the
6727: beginning of the seventies in (Connes \cite{Co_2}) following
6728: Tomita's theory is that such an algebra $M$ inherits
6729: from its noncommutativity a god-given time evolution:
6730: \begin{equation}
6731: \delta\,: \quad \R \quad\longrightarrow \quad{\rm Out}\,M=\,{\rm
6732: Aut}\,M/{\rm Inn}\,M
6733: \end{equation}
6734:
6735: where ${\rm Out}\,M ={\rm Aut}\,M/{\rm Inn}\,M$ is the quotient of
6736: the group of automorphisms of $M$ by the normal subgroup of inner
6737: automorphisms. This led in \cite{Co_2} to the reduction from type
6738: III to type II and their automorphisms and eventually to the
6739: classification of injective factors.
6740:
6741:
6742: \noindent They are classified by their module,
6743: \begin{equation}
6744: {\rm Mod} (M) \mathop{\subset}_{\sim} \ \R_+^* \, , \label{eq:(20)}
6745: \end{equation}
6746: which is a virtual closed subgroup of $\R_+^*$ in the sense of
6747: G.~Mackey, i.e. an ergodic action of $\R_+^*$, called the flow of
6748: weights \cite{CT}. This invariant was first defined and used in
6749: \cite{Co_2}, to show in particular the existence of hyperfinite
6750: factors which are not isomorphic to Araki-Woods factors.
6751:
6752: The ``measure theory" level \ie the set-up of von-Neumann algebras
6753: does not suffice to obtain the relevant cohomology theory and one
6754: needs to be given a weakly dense subalgebra $\cA\subset M$ playing
6755: the role of smooth functions on the noncommutative space. This
6756: algebra will play a key role when cyclic cohomology is used at a
6757: later stage. At first one only uses its norm closure $A=\bar \cA$ in
6758: $M$ and assumes that it is globally invariant under the modular
6759: automorphism group $\sigma_t^\varphi$ of a faithful normal state
6760: $\varphi$ on $M$. One can then proceed with the thermodynamics of
6761: the $C^*$ dynamical system $(A,\sigma_t)$. By a very simple
6762: procedure assuming that KMS states at low temperature are of type I,
6763: one obtains a ``cooling morphism" $\pi$ which is a morphism of
6764: algebras from the crossed product $\hat \cA=\,\cA
6765: \rtimes_{\sigma}\R$ to a type I algebra of compact operator valued
6766: functions on a canonical $\R^*_+$-principal bundle
6767: $\tilde\Omega_\beta$ over the space $\Omega_\beta$ of type I
6768: extremal KMS$_\beta$ states fulfilling a suitable regularity
6769: condition (\cf \cite{CCM}). Any $\varepsilon\in \Omega_\beta$ gives
6770: an irreducible representation $\pi_\varepsilon$ of $\cA$ and the
6771: choice of its essentially unique extension to $\hat \cA$ determines
6772: the fiber of the $\R^*_+$-principal bundle $\tilde\Omega_\beta$. The
6773: cooling morphism is then given by,
6774: \begin{equation}
6775: \pi_{\varepsilon,H}(\int \,x(t)\,U_t\,dt)=\,\int
6776: \,\pi_\varepsilon(x(t))\,e^{itH}\,dt .
6777: \end{equation}
6778: This morphism is equivariant for the dual action $\theta_\lambda \in
6779: \Aut(\hat \cA)$ of $\R^*_+$,
6780: \begin{equation}\label{dualaction}
6781: \theta_\lambda(\int \,x(t)\,U_t\,dt)=\,\int
6782: \,\lambda^{it}\,x(t)\,U_t\,dt.
6783: \end{equation}
6784:
6785: The key point is that the range of the morphism $\pi$ is contained
6786: in an algebra of functions on $\tilde\Omega_\beta$ with values in
6787: trace class operators. In other words modulo a Morita equivalence
6788: one lands in the commutative world provided one lowers the
6789: temperature.
6790:
6791:
6792: The interesting space is obtained by ``distillation" and is simply
6793: given by the cokernel of the cooling morphism $\pi$ but this does
6794: not make sense in the category of algebras and algebra homomorphisms
6795: since the latter is not even an additive category. This is where
6796: cyclic cohomology enters the scene : the category of cyclic modules
6797: is an abelian category with a natural functor from the category of
6798: algebras and algebra homomorphisms.
6799:
6800: Cyclic modules are modules of the cyclic category $\Lambda$ which
6801: is a small category, obtained by enriching with {\it cyclic
6802: morphisms} the familiar {\it simplicial category} $\Delta$ of
6803: totally ordered finite sets and increasing maps. Alternatively,
6804: $\Lambda$ can be defined by means of its ``cyclic covering'', the
6805: category $E \Lambda$. The latter has one object $(\Zb , n)$ for each
6806: $n \geq 0$ and the morphisms $f : (\Zb , n) \to (\Zb , m)$ are given
6807: by non decreasing maps $f : \Zb \to \Zb \ $, such that $ f(x+n) =
6808: f(x)+m \qqq x \in \Zb$. One has $\Lambda = E \Lambda / \Zb$, with
6809: respect to the obvious action of $\Zb$ by translation. To any
6810: algebra $\cA$ one associates a module $\cA^{\natural}$ over the category
6811: $\Lambda$ by assigning to each $n$ the $(n+1)$ tensor power
6812: $\cA\otimes \cA\cdots \otimes \cA$. The cyclic morphisms correspond to the
6813: cyclic permutations of the tensors while the face and degeneracy
6814: maps correspond to the algebra product of consecutive tensors and
6815: the insertion of the unit. The corresponding functor $\cA\to
6816: \cA^\natural$ gives a linearization of the category of associative
6817: algebras and cyclic cohomology appears as a derived functor.
6818:
6819: One can thus define the {\em distilled} module $D(\cA,\varphi)$ as
6820: the cokernel of the cooling morphism and consider the action of
6821: $\R^*_+$ (obtained from the above equivariance) in the cyclic
6822: homology group $HC_0(D(\cA,\varphi))$. As shown in \cite{CCM} this
6823: in the simplest case of the BC-system gives a cohomological
6824: interpretation of the above spectral realization of the zeros of the
6825: Riemann zeta function (and of Hecke L-functions).
6826:
6827: One striking feature is that the KMS strip (\cf Figure \ref{FigKMS})
6828: becomes canonically identified in the process with the critical
6829: strip of the zeta function (recall that $\beta >1$) by
6830: multiplication by $i=\sqrt{ -1}$.
6831:
6832: This cohomological interpretation combines with the above theory of
6833: endomotives to give a natural action of the Galois group $G$ of
6834: $\bar\Q/\Q$ on the above cohomology. This action factorizes to the
6835: abelianization $G^{ab}$ and the corresponding decomposition
6836: according to characters of $G^{ab}$ corresponds to the spectral
6837: realization of L-functions.
6838:
6839: The role of the invariant $S(M)$ in the classification of factors or
6840: of the more refined flow of weights mentioned above, is very similar
6841: to the role of the module of local or global fields and the Brauer
6842: theory of central simple algebras. In fact there is a striking
6843: parallel (see \cite{CCM}) between the lattice of unramified
6844: extensions $\K\to \,\K\otimes_{\F_q}\,\F_{q^n}$ of a global field of
6845: characteristic $p$ and the lattice of extensions of a factor $M$ by
6846: the crossed product algebras $M\to \,M\rtimes_{ \sigma_T}\Z$. Using
6847: the algebraic closure of $\F_q$ \ie the operation $\K\to
6848: \K\otimes_{\F_q}\,\bar \F_{q}$ corresponds to passing to the dual
6849: algebra $M\to \,M\rtimes_\sigma \R$ and the dual action corresponds
6850: to the Frobenius automorphism when as above the appropriate
6851: cohomological operations (distillation and $HC_0)$ are performed.
6852:
6853: \bigskip
6854: \bigskip
6855: \begin{center}
6856: \begin{tabular}{|c|c|}
6857: \hline & \\
6858: \bf{Global field $\K$} &\ \bf{Factor $M$} \\
6859: &\\ \hline & \\
6860: ${\rm Mod}\,\K\subset \R_+^*$ &\ ${\rm Mod}\,M\subset \R_+^*$ \\
6861: &\\
6862: \hline & \\
6863: $\K\to \,\K\otimes_{\F_q}\,\F_{q^n}
6864: $ &\ $M\to \,M\rtimes_{ \sigma_T}\Z$ \\
6865: &\\ \hline & \\ $K\to
6866: \K\otimes_{\F_q}\,\bar \F_{q}$ &\ $M\to \,M\rtimes_\sigma \R$ \\
6867: &\\ \hline & \\ \bf{Points}
6868: $C(\bar \F_{q})$ &\ $\Gamma\subset$ \bf{$X_\Q$} \\
6869: &\\ \hline
6870: \end{tabular}
6871: \end{center}
6872:
6873: \bigskip
6874: \bigskip
6875: \bigskip
6876:
6877: \medskip
6878: A notable difference with the original Hilbert space theoretic
6879: spectral realization of \cite{CoRH} is that while in the latter case
6880: only the critical zeros were appearing directly (the possible
6881: non-critical ones appearing as resonances), in the cyclic homology
6882: set-up it is more natural to use everywhere the ``rapid decay"
6883: framework (advocated in \cite{Meyer}) so that all zeros appear on
6884: the same footing. This eliminates the difficulty coming from the
6885: potential non-critical zeros, so that the trace formula is much
6886: easier to prove and reduces to the Riemann-Weil explicit formula.
6887: However, it was not obvious how to obtain a direct geometric proof
6888: of this formula from the $S$-local trace formula of \cite{CoRH}.
6889: This was done in \cite{Meyer}, showing that the noncommutative
6890: geometry framework makes it possible to give a geometric
6891: interpretation of the Riemann-Weil explicit formula. While the
6892: spectral side of the trace formula is given by the action on the
6893: cyclic homology of the distilled space, the geometric side is given
6894: as follows \cite{CCM}.
6895:
6896:
6897: \medskip
6898:
6899: \begin{thm} \label{geom} Let $h
6900: \in S (C_\K)$. Then the following holds:
6901: \begin{equation}\label{trace3}
6902: \Trace \; \left(\vartheta(h)|_{ \cH^1} \right)=\,\wh h(0) + \wh h(1)
6903: -\,\Delta\bullet \Delta\;h(1) -\,\sum_{v} \int_{(\K^*_v,e_{\K_v})}'
6904: \, \frac{h(u^{-1})}{
6905: |1-u|}\, d^* u \, .
6906: \end{equation}
6907: \end{thm}
6908:
6909: \medskip
6910:
6911: We refer to \cite{CCM} for the detailed notations which are
6912: essentially those of \cite{CoRH}. The origin of the terms in the
6913: geometric side of the trace formula comes from the Lefschetz
6914: formula by Atiyah-Bott \cite{atbo} and its adaptation by
6915: Guillemin-Sternberg (\cf \cite{gui}) to the distribution theoretic
6916: trace for flows on manifolds, which is a variation on the theme of
6917: \cite{atbo}. For the action of $C_\K$ on the adele class space
6918: $X_\K$ the relevant periodic orbits on which the computation
6919: concentrates turn out to form also the classical points of the
6920: noncommutative space $X_\K \backslash C_\K$ distilled in the above
6921: sense from $X_\K$. This ``classical" subspace of $X_\K\backslash
6922: C_\K$ is given by
6923: \begin{equation}\label{defcurve}
6924: \Gamma_\K=\,\cup \; C_\K\,[v]\,,\quad v\in \Sigma_\K .
6925: \end{equation}
6926: where for each place $v\in\Sigma_\K$, one lets $[v]$ be the adele
6927: \begin{equation}\label{base}
6928: [v]_w=1\qqq w\neq v\,,\quad [v]_v=0\,.
6929: \end{equation}
6930:
6931:
6932: In the function field case, one has a {\em non-canonical}
6933: isomorphism of the following form.
6934:
6935:
6936: \begin{prop} \label{Zspace}
6937: Let $\K$ be the function field of an algebraic curve $C$ over
6938: $\F_q$. Then the action of the Frobenius on $Y=C(\bar \F_q)$ is
6939: isomorphic to the action of $q^\Z$ on the quotient
6940: $$
6941: \Gamma_\K/C_{\K,1}.
6942: $$
6943: \end{prop}
6944:
6945:
6946: \begin{center}
6947: \begin{figure}
6948: \includegraphics[scale=1]{primecurve4.eps}
6949: \caption{The classical points of the adeles class space
6950: \label{Figpoints}}
6951: \end{figure}
6952: \end{center}
6953:
6954:
6955:
6956: In the case $\K=\Q$ the space $\Gamma_\Q/C_{\Q,1}$ appears as the
6957: union of periodic orbits of period $\log p$ under the action of
6958: $C_\Q/C_{\Q,1}\sim \R$ (\cf Figure \ref{Figpoints}). This gives a
6959: first approximation to the sought for space $Y=C(\bar \F_q)$ in
6960: characteristic zero. One important refinement is obtained from the
6961: subtle nuance between the adelic description of $X_\Q$ and the finer
6962: description in terms of the endomotive obtained from the pointed
6963: algebraic variety $(\bG_m(\Q),1)$. The second description keeps
6964: track of the Galois symmetry and as in proposition \ref{Zspace} the
6965: isomorphism of the two descriptions is non canonical.
6966:
6967:
6968:
6969: \smallskip
6970:
6971: At this point we have, in characteristic zero, several of the
6972: geometric notions which are the analogues of the ingredients of
6973: Weil's proof and it is natural to try and imitate the steps of his
6974: proof. The step (A), \ie the explicit formula is taken care of by
6975: Theorem \ref{geom}. That what remains is to prove a positivity is a
6976: well known result of A. Weil (\cf \cite{EB}) which states that RH
6977: is equivalent to the positivity of the distribution entering in the
6978: explicit formulae. Thanks to the above $H^1$ obtained as the cyclic
6979: homology of the distilled module, Weil's reformulation can be stated
6980: as follows.
6981:
6982: \begin{thm} \label{pos} The following two conditions are
6983: equivalent.
6984: \begin{itemize}
6985: \item All $L$-functions with Gr\"ossencharakter on $\K$ satisfy the Riemann
6986: Hypothesis.
6987: \item $\Trace\,\vartheta(f\,\star\,
6988: f^\sharp)|_{ \cH^1} \, \geq \,0$, for all $f\in\cS (C_\K)$.
6989: \end{itemize}
6990: \end{thm}
6991:
6992:
6993: Here we used the notation
6994: \begin{equation}\label{conv}
6995: f=\,f_1 \star f_2 \,,\ \ \ \text{ with } \ (f_1 \star f_2)(g)=\,\int
6996: \,f_1(k)\, f_2(k^{-1}\,g) \,d^*g
6997: \end{equation}
6998: for the convolution of functions, using the multiplicative Haar
6999: measure $d^*g$, and for the adjoint
7000: \begin{equation}\label{adj1}
7001: f\to f^{\sharp}\,,\quad f^{\sharp}(g)=\,\vert\, g\vert^{-1}\,\bar
7002: f(g^{-1}) .
7003: \end{equation}
7004:
7005: The role of the specific correspondences used in Weil's proof of RH
7006: in positive characteristic is played by the test functions $ f\in
7007: \cS (C_\K) $. More precisely the scaling map which replaces $f(x)$
7008: by $f(g^{-1} x)$ has a graph, namely the set of pairs $(x, g^{-1} x)
7009: \in X_\K \times X_\K$, which we view as a correspondence $Z_g$.
7010: Then, given a test function $f$ on the ideles classes, one assigns
7011: to $f$ the linear combination
7012: \begin{equation}\label{corresp}
7013: Z(f)=\,\int f(g) Z_g d^*g
7014: \end{equation}
7015: of the above graphs, viewed as a ``divisor" on $X_\K \times X_\K$.
7016:
7017: The analogs of the degrees $d(Z)$ and codegrees $d^{\,'}(Z)=\,d(Z')$
7018: of correspondences in the context of Weil's proof are given, for the
7019: degree, by
7020: \begin{equation}\label{degree}
7021: d(Z(h))=\,\wh h(1)=\,\int\,h(u)\, \vert u \vert \, d^* \, u ,
7022: \end{equation}
7023: so that the degree $d(Z_g)$ of the correspondence $Z_g$ is equal to
7024: $|g|$. Similarly, for the codegree one has
7025: \begin{equation}\label{codegree}
7026: d^{\,'}(Z(h))=\,d(Z(\bar h^\sharp))=\,\int\,h(u)\, d^* \, u =\,\wh
7027: h(0) ,
7028: \end{equation}
7029: so that the codegree $d^{\,'}(Z_g)$ of the correspondence $Z_g$ is
7030: equal to $1$.
7031:
7032:
7033: One of the major difficulties is to find the replacement for the
7034: principal divisors which in Weil's proof play a key role as an ideal
7035: in the algebra of correspondences on which the trace vanishes. At
7036: least already one can see that there is an interesting subspace $V$
7037: of the linear space of correspondences described above on which the
7038: trace also vanishes. It is given by the subspace
7039: \begin{equation}\label{V}
7040: V\subset \cS(C_\K) \,,\quad V=\{\,g(x)=\, \sum\,\xi(k\,x)\,|\,\xi\in
7041: \cS(\A_\K)_0\} ,
7042: \end{equation}
7043: where the subspace $\cS(\A_\K)_0\subset \cS(\A_\K)$ is defined by
7044: the two boundary conditions
7045: $$
7046: \xi(0)=0 \,,\quad \int \,\xi(x)\,dx\,=\,0 .
7047: $$
7048:
7049: \medskip
7050:
7051:
7052: \begin{lem} \label{vanish}
7053: For any $f\in V\subset \cS (C_\K)$, one has
7054: $$
7055: \vartheta(f)|_{ \cH^1}=\,0 .
7056: $$
7057: \end{lem}
7058:
7059: \medskip
7060:
7061: This shows that the Weil pairing of Theorem \ref{pos} admits a huge
7062: radical given by all functions which extend to adeles and gives
7063: another justification for working with the above cohomology $H^1$.
7064: In particular one can
7065: modify arbitrarily the degree and codegree of the correspondence
7066: $Z(h)$ by adding to $h$ an element of the radical $V$ using a subtle
7067: failure of Fubini's theorem. We will show in a forthcoming paper
7068: \cite{CoCM} that several of the steps of Weil's proof can be
7069: transposed in the framework described above.
7070:
7071: This constitutes a clear motivation to develop noncommutative
7072: geometry much further. One can write a very tentative form of a
7073: dictionary from the language of algebraic geometry (in the case of
7074: curves) and that of noncommutative geometry. The dictionary is
7075: summarized in the following table. It should be stressed that the
7076: main problem is to find the correct translation in the right column
7077: (non-commutative geometry) of the well established notion of
7078: principal divisor in the (algebraic geometry) left column. The table
7079: below is too rough in that respect since one does not expect to be
7080: able to work in the usual ``primary" theory which involves periodic
7081: cyclic homology and index theorems. Instead one expects that both
7082: the unstable cyclic homology and the finer invariants of spectral
7083: triples arising from transgression will play an important role. Thus
7084: the table below should be taken as a very rough first approximation,
7085: and a motivation for developing the missing finer notions in the
7086: right column.
7087:
7088: \bigskip
7089: \bigskip\bigskip\bigskip\bigskip\bigskip
7090:
7091:
7092:
7093:
7094: \begin{small}
7095: \begin{center}
7096: \begin{tabular}{|c|c|}
7097: \hline & \\
7098: Virtual correspondences & bivariant $K$-theory class $\Gamma$ \\ & \\
7099: \hline & \\
7100: Modulo torsion & $KK(A,B\otimes {\rm II}_1)$ \\ & \\
7101: \hline & \\
7102: Effective correspondences & Epimorphism of $C^*$-modules \\
7103: & \\ \hline & \\
7104: Principal correspondences & Compact morphisms \\
7105: & \\ \hline & \\
7106: Composition & cup product in $KK$-theory \\
7107: & \\ \hline & \\
7108: Degree of correspondence & Pointwise index $d(\Gamma)$ \\
7109: & \\ \hline & \\
7110: $\deg D(P)\geq g \Rightarrow \sim$ effective & $d(\Gamma)>0
7111: \Rightarrow \exists K, \Gamma+K$ onto \\
7112: & \\ \hline & \\
7113: Adjusting the degree & Fubini step \\
7114: by trivial correspondences & on the test functions \\
7115: & \\ \hline & \\
7116: Frobenius correspondence & Correspondence $Z_g$ \\
7117: & \\ \hline & \\
7118: Lefschetz formula & bivariant Chern of $Z(h)$ \\
7119: &
7120: (localization on graph $Z(h)$) \\ & \\ \hline & \\
7121: Weil trace unchanged & bivariant Chern unchanged
7122: \\ by principal divisors & by compact
7123: perturbations
7124: \\
7125: & \\ \hline
7126: \end{tabular}
7127: \end{center}
7128: \end{small}
7129:
7130:
7131: \newpage
7132:
7133:
7134:
7135:
7136: \begin{thebibliography}{99}
7137:
7138: \bibitem{ANS} J.~Aastrup, R.~Nest, E.~Schrohe,
7139: {\it A Continuous Field of C*-algebras and the
7140: Tangent Groupoid for Manifolds with Boundary}, ArXiv
7141: math.FA/0507317.
7142:
7143:
7144: \bibitem{aba-exel:1997}
7145: B.~Abadie, R.~Exel, {\em Hilbert $C^*$-bimodules
7146: over commutative $C^*$-algebras and an isomorphism condition
7147: for quantum Heisenberg manifolds}, Rev. Math. Phys. Vol.9 (1997)
7148: no. 4, 411--423.
7149:
7150: \bibitem{aba-eil-exel:1998}
7151: B.~Abadie, S.~Eilers, R.~Exel,
7152: {\em Morita equivalence for crossed products by Hilbert $C^*$-bimodules}.
7153: Trans. Amer. Math. Soc. Vol.350, (1998) no. 8, 3043--3054.
7154:
7155: \bibitem{AndPut} J.E.~Anderson, I.F.~Putnam, {\em Topological
7156: invariants for substitution tilings and their associated $C\sp
7157: *$-algebras}. Ergodic Theory Dynam. Systems 18 (1998), no. 3, 509--537.
7158:
7159: \bibitem{Anonym} Anonymous Persian manuscript, {\em Fi tadakhul
7160: al-ashkal al-mutashabiha aw mutawafiqa}, Biblioth\`eque Nationale, Paris,
7161: Ancien fonds persan 169.
7162:
7163: \bibitem{AntChris} C.~Antonescu, E.~Christensen, {\em Spectral triples
7164: for AF $C^*$-algebras and metrics on the Cantor set}, preprint
7165: arXiv math.OA/0309044.
7166:
7167: \bibitem{Ard1} Farhad Ardalan, Amir H. Fatollahi, {\em Point-Like
7168: Structure in Strings and Non-Commutative Geometry},
7169: Phys.Lett. B408 (1997) 157-163.
7170:
7171: \bibitem{Ard2} F. Ardalan, H. Arfaei, N. Sadooghi, {\em
7172: On the Anomalies and Schwinger Terms in Noncommutative Gauge
7173: Theories}, hep-th/0507230.
7174:
7175: \bibitem{ArdBak} Nader Ardalan and Laleh Bakhtiar, {\em The sense of
7176: unity: the Sufi tradition in Persian architecture}, The University of
7177: Chicago Press, 1973.
7178:
7179: \bibitem{art-tat-vdb:1990}
7180: M.~Artin, J.~Tate, and M.~Van~den Bergh,
7181: {\em Some algebras associated to automorphisms of elliptic curves}. The
7182: Grothendieck Festschrift. vol.I.
7183: Prog. Math., Vol.86:33--85, 1990.
7184:
7185: \bibitem{atbo} M.F.~Atiyah and R.~Bott, {\it A Lefschetz fixed
7186: point formula for elliptic complexes: I}, Annals of Math, Vol.86
7187: (1967) 374--407.
7188:
7189: \bibitem{AT} M.F.~Atiyah, {\it Global theory of elliptic
7190: operators}. Proc. Internat. Conf. on {\it Functional Analysis and
7191: Related Topics} (Tokyo, 1969). University of Tokyo press, Tokyo
7192: 1970, 21-30.
7193:
7194: \bibitem{BaMo} Michael Baake and Robert V. Moody (Editors), {\em
7195: Directions in mathematical quasicrystals}, CRM Monograph Series, 13.
7196: American Mathematical Society, Providence, RI, 2000. viii+379 pp.
7197:
7198: \bibitem{BFSS} T. Banks, W. Fischler, S.H. Shenker, L. Susskind,
7199: {\em M theory as a Matrix model: a conjecture},
7200: Phys.Rev. D55 (1997) 5112-5128. hep-th/9610043.
7201:
7202: \bibitem{BauCo} P.~Baum and A.~Connes,
7203: {\it Geometric K-theory for Lie groups and
7204: foliations}, Preprint IHES (M/82/), 1982, L'enseignement
7205: Mathematique, t.46 (2000), 1--35.
7206:
7207: \bibitem{BCH} P. Baum, A. Connes, and N. Higson, {\it Classifying space
7208: for proper -actions and K-theory of group $C^*$-algebras}, Contemp.
7209: Math. Vol.167 (1994), 241--291.
7210:
7211: \bibitem{BayFron} F.~Bayen and C.~Fronsdal.
7212: {\em Quantization on the sphere}.
7213: J. Math. Phys. Vol.22 (1981) N.7, 1345--1349.
7214:
7215: \bibitem{Bellissard1}
7216: J.~Bellissard,
7217: {\it Noncommutative geometry and quantum Hall effect}.
7218: In: Proceedings of the International Congress of Mathematicians,
7219: Vol.\ 1, 2 (Z\"urich, 1994), pages 1238--1246, Basel, 1995. Birkh\"auser.
7220:
7221: \bibitem{Bellissard2}
7222: J.~Bellissard, A.~van Elst, and H.~Schulz-Baldes.
7223: {\it The noncommutative geometry of the quantum Hall effect}.
7224: J. Math. Phys., 35 (1994) N.10, 5373--5451.
7225:
7226: \bibitem{BellC} J.~Bellissard, {\em The noncommutative geometry of
7227: aperiodic solids}, in ``Geometric and topological methods for
7228: quantum field theory (Villa de Leyva, 2001)'', 86--156, World
7229: Scientific, 2003.
7230:
7231: \bibitem{BF} M-T.~Benameur, T.~Fack,
7232: {\em On von Neumann spectral triples},
7233: math.KT/0012233.
7234:
7235: \bibitem{BDV} R.~Berger, M.~Dubois-Violette, M.~Wambst, {\it Homogeneous
7236: algebras}, math.QA/0203035.
7237:
7238: \bibitem{BoGia} C.G.~Bollini, J.J.~Giambiagi {\em Lowest order
7239: divergent graphs in $\nu$-dimensional space} Phys.Lett. B40 (1972)
7240: 566--568.
7241:
7242: \bibitem{EB} E.~Bombieri, {\em Problems of the Millenium:
7243: The Riemann Hypothesis}, Clay mathematical Institute (2000).
7244:
7245: \bibitem{Born} M.~Born and P.~Jordan.
7246: {\it Zur Quantenmechanik},
7247: Z. Phys., 34 (1925) 858--888.
7248:
7249: \bibitem{BKLR} D.~Borthwick, S.~Klimek, A.~Lesniewski, M.~Rinaldi, {\em
7250: Supersymmetry and Fredholm modules over quantized spaces}.
7251: Comm. Math. Phys. 166 (1994), no. 2, 397--415.
7252:
7253: \bibitem{BC} J.-B. Bost and A. Connes, {\it Hecke Algebras, Type
7254: III factors and phase transitions with spontaneous symmetry
7255: breaking in number theory}, Selecta Mathematica, New
7256: Series Vol.1, N.3 (1995), 411--457.
7257:
7258: \bibitem{BMa1} P.~Bouwknegt, K.~Hannabuss, V.~Mathai, {\em $T$-duality
7259: for principal torus bundles}. J. High Energy Phys. 2004, no. 3,
7260: 018, 10 pp. (electronic).
7261:
7262: \bibitem{BMa2} P.~Bouwknegt, J.~Evslin, V.~Mathai, {\em $T$-duality:
7263: topology change from $H$-flux}. Comm. Math. Phys. 249 (2004),
7264: no. 2, 383--415.
7265:
7266: \bibitem{BoHa} M.~Boyle, D.~Handelman, {\em Orbit equivalence, flow
7267: equivalence, and ordered cohomology}, Israel J. Math. 95 (1996)
7268: 169--210.
7269:
7270: \bibitem{Bratteli} O.~Bratteli.
7271: {\it Inductive limits of finite dimensional $C^*$-algebras}.
7272: Trans. Amer. Math. Soc., 171 (1972) 195--234.
7273:
7274: \bibitem{BM} P.~Breitenlohner, D.~Maison, {\em Dimensional
7275: renormalization and the action principle}. Comm. Math. Phys.
7276: Vol.52 (1977), N.1, 11--38.
7277:
7278: \bibitem{BruSee} J.~Br\"uning, R.~Seeley, {\it Regular singular
7279: asymptotics}, Adv. in Math. 58 (1985) 133--148.
7280:
7281: \bibitem{BryNi} J-L. Brylinski and V. Nistor, {\it Cyclic Cohomology
7282: of Etale Groupoids}, K-theory, Vol.8 (1994), 341--365.
7283:
7284: \bibitem{Bulatov} M.S.~Bulatov, {\em Geometricheskaya garmonizatsiya v
7285: arkhitekture Srednei Azii IX-XV vv.}, Nauka, 2nd ed., Moskva 1988,
7286: 360 pp.
7287:
7288: \bibitem{Burgh}
7289: D. Burghelea, {\it The cyclic homology of the group rings}, Comment.
7290: Math. Helv. Vol.60 (1985) 354--365.
7291:
7292: \bibitem{CHMM} A.~Carey, K.~Hannabuss, V.~Mathai, P.~McCann,
7293: {\em Quantum Hall Effect on the hyperbolic plane},
7294: Commun. Math. Physics, Vol.190, no. 3 (1998) 629--673.
7295:
7296: \bibitem{CPRS} Alan L.~Carey, J.~Phillips, A.~Rennie, F. A.~Sukochev
7297: {\em The Local Index Formula in Semifinite von Neumann Algebras I:
7298: Spectral Flow},
7299: math.OA/0411021.
7300:
7301: \bibitem{CPRS1} Alan L.~Carey, J.~Phillips, A.~Rennie, F. A.~Sukochev
7302: {\em The Local Index Formula in Semifinite von Neumann Algebras II:
7303: The Even Case},
7304: math.OA/0411021.
7305:
7306: \bibitem{Castera} J.M.~Castera, {\em Zellijs, muqarnas and
7307: quasicrystals}, ISAMA 99 Proceedings, pp.99--104, 1999.
7308:
7309: \bibitem{Castera2} J.M.~Castera, {\em Arabesques}, ACR editions,
7310: 1999, 486 pp.
7311:
7312: \bibitem{ChakraPal} P.S.~Chakraborti, A.~Pal, {\em Equivariant
7313: spectral triple on the quantum $SU(2)$-group}, $K$-Theory 28
7314: (2003), no. 2, 107--126.
7315:
7316: \bibitem{Chakra} T.~Chakraborti, P.~Pietil\"anen,
7317: {\em The Quantum Hall Effects}, Second Edition, Springer 1995.
7318:
7319: \bibitem{C-C} A.~Chamseddine and A.~Connes, {\it Universal
7320: formulas for noncommutative geometry actions}, Phys. Rev.
7321: Letters, Vol.77, N.24, (1996) 4868--4871.
7322:
7323: \bibitem{C-C2} A.~Chamseddine and A.~Connes, {\it The spectral
7324: action principle}, Commun. Math. Phys. 186 (1997) 731--750.
7325:
7326: \bibitem{C-C3} A.~Chamseddine and A.~Connes, {\it Scale
7327: Invariance in the spectral
7328: action}, Arxiv hep-th/0512169.
7329:
7330: \bibitem{ChGr} J. Cheeger, M. Gromov, {\it $L^2$-cohomology and group
7331: cohomology}, Topology, Vol.25 (1986), 189--215.
7332:
7333: \bibitem{chou} A.~Chou, {\it The Dirac Operator on Spaces
7334: with Conical Singularities
7335: and Positive Scalar Curvatures}, Trans. Am. Math. Soc. 289 (1985),
7336: no. 1, 1--40.
7337:
7338: \bibitem{IvChris} E.~Christensen, C.~Ivan, {\it Sums of two
7339: dimensional spectral triples}, ArXiv math.OA/0601024
7340:
7341: \bibitem{Collins} J.~Collins, {\em Renormalization}, Cambridge
7342: Monographs in Math. Physics, Cambridge University Press, 1984.
7343:
7344: \bibitem{Co_2} A.~Connes, {\it Une classification des facteurs de type III},
7345: Ann. Sci. Ecole Norm. Sup., {\bf 6}, n.~4 (1973), 133-252.
7346:
7347: \bibitem{Co78} A.~Connes, {\it Sur la th\'eorie non commutative de
7348: l'int\'egration} Alg\`ebres d'op\'erateurs (S\'em., Les Plans-sur-Bex,
7349: 1978), p. 19-143, Lecture Notes in Math., 725, Springer, Berlin,
7350: 1979.
7351:
7352: \bibitem{Co90} A.~Connes, {\it G\'eom\'etrie Noncommutative}.
7353: InterEditions, Paris, 1990.
7354:
7355: \bibitem{Co94} A.~Connes, {\it Noncommutative geometry}.
7356: Academic Press, 1994.
7357:
7358: \bibitem{Co-Thom} A.~Connes, {\em An analogue of the Thom isomorphism
7359: for crossed products of a $C^*$-algebra by an action of $\R$},
7360: Adv. in Math. 39 (1981), no. 1, 31--55.
7361:
7362: \bibitem{Co1} A.~Connes, {\it Cyclic cohomology and the
7363: transverse fundamental class of a foliation}. In: Geometric
7364: methods in operator algebras (Kyoto, 1983). Pitman Res. Notes in
7365: Math., 123, Longman, Harlow 1986, 52--144.
7366:
7367: \bibitem{C3} A.~Connes, {\it $C^*$ alg\`ebres et
7368: g\'eom\'etrie differentielle}. C.R. Acad. Sci. Paris,
7369: Ser.~A-B , 290 (1980) 599--604.
7370:
7371: \bibitem{ConnesCH} A.~Connes, {\em Noncommutative
7372: differential geometry}. Inst. Hautes Etudes Sci. Publ.
7373: Math., 62, 1985, 257--360.
7374:
7375: \bibitem{Co-fr} A.~Connes, {\it Compact metric spaces, Fredholm
7376: modules, and hyperfiniteness}, Ergod. Th. Dynam. Sys. (1989) 9,
7377: 207--220.
7378:
7379: \bibitem{Cotheta} A.~Connes, {\it On the Chern character of $\theta$-summable
7380: Fredholm modules} Comm. Math. Phys. 139 (1991), n° 1, p. 171-181.
7381:
7382: \bibitem{Co3} A.~Connes, {\it Gravity coupled with matter and
7383: foundation of noncommutative geometry}. Commun. Math. Phys.,
7384: 182 (1996) 155--176.
7385:
7386: \bibitem{Coo2} A.~Connes, {\it Noncommutative geometry and
7387: reality}. Journal of Math. Physics, 36, n.11 (1995) 6194--6231.
7388:
7389: \bibitem{ConnesS3} A.~Connes, {\em Geometry from the spectral point of
7390: view}. Lett. Math. Phys. 34 (1995), no. 3, 203--238.
7391:
7392: \bibitem{CoRH} A.~Connes, {\it Trace formula in Noncommutative
7393: Geometry and the zeros of the Riemann zeta function}. Selecta
7394: Mathematica. New Ser. 5 (1999) 29--106.
7395:
7396: \bibitem{Co-Qgr} A.~Connes, {\em Cyclic cohomology, quantum group
7397: symmetries and the local index formula for ${\rm SU}\sb q(2)$}.
7398: J. Inst. Math. Jussieu 3 (2004), no. 1, 17--68.
7399:
7400: \bibitem{CCM} A.~Connes, C.~Consani, M.~Marcolli,
7401: {\em Noncommutative geometry and motives:
7402: the thermodynamics of endomotives}, math.QA/0512138.
7403:
7404: \bibitem{CoCM} A.~Connes, C.~Consani, M.~Marcolli, {\em The Weil proof
7405: and the geometry of the adeles class space}, in preparation.
7406:
7407: \bibitem{CDS} A.~Connes, M.~Douglas, A.~Schwarz,
7408: {\it Noncommutative geometry and Matrix theory: compactification
7409: on tori}.
7410: J. High Energy Phys. (1998) no. 2, Paper 3, 35 pp. (electronic)
7411:
7412: \bibitem{CDV1} A.~Connes, M.~Dubois-Violette, {\em Noncommutative
7413: finite-dimensional manifolds. I. Spherical manifolds and related
7414: examples}. Comm. Math. Phys. 230 (2002), no. 3, 539--579.
7415:
7416: \bibitem{CDV2} A.~Connes, M.~Dubois-Violette, {\em Moduli space and
7417: structure of noncommutative 3-spheres}. Lett. Math. Phys. 66
7418: (2003), no. 1-2, 91--121.
7419:
7420: \bibitem{CDV3} A.~Connes, M.~Dubois-Violette, {\em Non commutative
7421: finite-dimensional manifolds II. Moduli space and structure of
7422: noncommutative 3-spheres}. ArXiv math.QA/0511337.
7423:
7424: \bibitem{CLa} A.~Connes and G.~Landi, {\it Noncommutative manifolds,
7425: the instanton algebra and isospectral
7426: deformations}. Comm. Math. Phys. 221 (2001), no. 1, 141--159.
7427:
7428: \bibitem{CL} A.~Connes, J.~Lott, {\em Particle models and noncommutative
7429: geometry}, in ``Recent advances in field theory''
7430: (Annecy-le-Vieux, 1990). Nuclear Phys. B Proc. Suppl. 18B
7431: (1990), 29--47 (1991).
7432:
7433: \bibitem{CoMa} A.~Connes, M.~Marcolli, {\em From Physics to Number
7434: Theory via Noncommutative Geometry.
7435: Part I: Quantum Statistical Mechanics of $Q$-lattices},
7436: to appear in the volume
7437: ``Frontiers in Number Theory Physics and Geometry''
7438: math.NT/0404128.
7439:
7440: \bibitem{CoMajgp1} A.~Connes, M.~Marcolli, {\em $\Q$-lattices: quantum
7441: statistical mechanics and
7442: Galois theory}, Journal of Geometry and Physics, Vol. 56 (2005) N.1,
7443: 2--23.
7444:
7445: \bibitem{CoMar-an} A.~Connes, M.~Marcolli, {\em Anomalies, Dimensional
7446: Regularization and Noncommutative Geometry}, in preparation.
7447:
7448: \bibitem{CoMarBook} A.~Connes, M.~Marcolli, {\em Noncommutative geometry,
7449: from quantum fields to motives} (tentative title), book in preparation.
7450: To appear as a co-publication of
7451: the American Mathematical Society, Colloquium Publications Series, and
7452: Hindustan Book Agency, Texts and Readings in Mathematics Series.
7453:
7454: \bibitem{CMR} A.~Connes, M.~Marcolli, N.~Ramachandran, {\em KMS states
7455: and complex multiplication}, to appear in Selecta Mathematica.
7456:
7457: \bibitem{CMR2} A.~Connes, M.~Marcolli, N.~Ramachandran, {\em KMS states
7458: and complex multiplication. Part II}, to appear in the proceedings of
7459: the Abel Symposium 2005.
7460:
7461: \bibitem{CoMoind} A.~Connes, H.~Moscovici,
7462: {\em Cyclic cohomology, the Novikov conjecture and hyperbolic
7463: groups}, Topology, Vol. 29 (1990) no. 3, 345--388.
7464:
7465: \bibitem{CoMo} A.~Connes and H.~Moscovici, {\it The local
7466: index formula in noncommutative geometry}. GAFA, 5 (1995)
7467: 174--243.
7468:
7469: \bibitem{CoMoHopf} A.~Connes and H.~Moscovici, {\it Hopf Algebras,
7470: Cyclic Cohomology and the Transverse Index Theorem}.
7471: Commun. Math. Phys. 198 (1998) 199--246.
7472:
7473: \bibitem{CoMoHopf2} A.~Connes and H.~Moscovici, {\it Differentiable
7474: cyclic cohomology and Hopf algebraic structures in transverse
7475: geometry}. Essays on geometry and related topics, Vol. 1, 2, 217--255,
7476: Monogr. Enseign. Math., 38, Enseignement Math., Geneva, 2001.
7477:
7478: \bibitem{Co-Mosc1} A.~Connes and H.~Moscovici, {\em Modular Hecke
7479: algebras and their Hopf symmetry}, Mosc. Math. J. 4 (2004), no. 1,
7480: 67--109, 310 (math.QA/0301089).
7481:
7482: \bibitem{Co-Mosc2} A.~Connes and H.~Moscovici, {\em Rankin-Cohen
7483: Brackets and the Hopf Algebra of Transverse Geometry}, Mosc. Math. J.
7484: 4 (2004), no. 1, 111--130, 311.
7485: (math.QA/0304316).
7486:
7487: \bibitem{CoMo-back} A.~Connes and H.~Moscovici,
7488: {\em Background independent geometry and Hopf cyclic cohomology},
7489: math.QA/0505475.
7490:
7491: \bibitem{CoMoBook} A.~Connes and H.~Moscovici, {\it Spectral Geometry
7492: of Noncommutative Spaces} (tentative title), book in preparation.
7493:
7494: \bibitem{CoRi} A.~Connes and M.~Rieffel, {\it Yang-Mills for
7495: noncommutative two tori}. Contemp. Math. Oper. Algebra Math. Phys.
7496: AMS (1987) 237--266.
7497:
7498: \bibitem{CoSk} A.~Connes and G.~Skandalis,
7499: {\it The longitudinal index theorem for
7500: foliations} Publ. Res. Inst. Math. Sci. 20 (1984), n° 6, p.
7501: 1139-1183.
7502:
7503: \bibitem{CT} A.~Connes and M.~Takesaki,
7504: {\it The flow of weights on factors of type III}, Tohoku Math. J.
7505: {\bf 29} (1977), 473-575.
7506:
7507: \bibitem{CM1} C.~Consani, M.~Marcolli, {\em Noncommutative
7508: geometry, dynamics, and $\infty$-adic Arakelov geometry}. Selecta
7509: Math. (N.S.) 10 (2004), no. 2, 167--251.
7510:
7511: \bibitem{CM2} C.~Consani, M.~Marcolli, {\em New perspectives in
7512: Arakelov geometry}. Number theory, 81--102, CRM Proc. Lecture Notes,
7513: 36, Amer. Math. Soc., Providence, RI, 2004.
7514:
7515: \bibitem{Crai} M. Crainic, {\it Cyclic Cohomology
7516: of Etale Groupoids, the general case}, K-theory, Vol.17
7517: (1999), 319--362.
7518:
7519: \bibitem{Cu} J.~Cuntz, {\it A class of $C^*$--algebras and topological
7520: Markov chains II: reducible chains and the $Ext$--functor for
7521: $C^*$--algebras}, Invent. Math. 63 (1981) 25--40.
7522:
7523: \bibitem{CuKrie} J.~Cuntz, W.~Krieger, {\it A class of $C^*$--algebras
7524: and topological Markov chains}, Invent. Math. 56 (1980) 251--268.
7525:
7526: \bibitem{Davidson} K.R.~Davidson, {\it $C^*$-algebras by example}.
7527: American Mathematical Society, Providence, RI, 1996.
7528:
7529: \bibitem{Dix} J.~Dixmier, {\it Existence de traces non
7530: normales}. C.R. Acad. Sci. Paris, Ser. A-B, 262 (1966) A1107--A1108.
7531:
7532: \bibitem{D-V} M.~Dubois-Violette, {\it Generalized differential
7533: spaces with $d^N=0$ and
7534: the $q$-differential calculus},
7535: Czech J. Phys. 46 (1997) 1227--1233.
7536:
7537: \bibitem{D-V1} M.~Dubois-Violette, {\it
7538: $d^N=0$ : Generalized homology},
7539: K-Theory, 14 (1998) 371--404.
7540:
7541: \bibitem{D-V2}
7542: M.~Dubois-Violette, {\it Lectures on graded differential algebras
7543: and noncommutative geometry}, In Y.~Maeda and H.~Moriyoshi, editors,
7544: {\em Noncommutative
7545: Differential Geometry and Its Applications to Physics}, pages 245--306.
7546: Shonan, Japan, 1999, Kluwer Academic Publishers, 2001.
7547:
7548: \bibitem{D-V3} M.~Dubois-Violette and M.~Henneaux, {\it
7549: Generalized cohomology for irreducible tensor fields of mixed Young
7550: symmetry type}
7551: Lett. Math. Phys. 49 (1999) 245--252.
7552:
7553: \bibitem{D-V4} M.~Dubois-Violette and M.~Henneaux, {\it
7554: Tensor fields of mixed Young symmetry type and $N$-complexes},
7555: Commun. Math. Phys. 226 (2002) 393--418.
7556:
7557: \bibitem{FGV} H.~Figueroa, J.M.~Gracia-Bond\'ia, J.~Varilly, {\em
7558: Elements of Noncommutative Geometry}, Birkh\"auser, 2000.
7559:
7560: \bibitem{Gabo1} D. Gaboriau, {\it Invariants $\ell^2$ de relations
7561: d'\'equivalence et de groupes}, Publ. Math. Inst. Hautes \'Etudes
7562: Sci. No. 95 (2002), 93--150.
7563:
7564: \bibitem{Gabo2} D. Gaboriau, {\it Approximate dimension for
7565: equivalence relations and groups}, in preparation.
7566:
7567: \bibitem{GGISV} V.~Gayral, J.M.~Gracia-Bond\'ia, B.~Iochum,
7568: T.~Sch\"ucker, J.C.~V\'arilly, {\em Moyal planes are spectral
7569: triples}. Comm. Math. Phys. 246 (2004), no. 3, 569--623.
7570:
7571: \bibitem{GPS} T.~Giordano, I. F.~Putnam and C. Fr.~Skau,
7572: {\it Topological orbit equivalence and
7573: $C^*$-crossed products}. J. Reine Angew. Math. 469
7574: (1995), 51--111
7575:
7576: \bibitem{GuIso1} D.~Guido, T.~Isola, {\em Dimensions and singular
7577: traces for spectral triples, with applications to fractals},
7578: Journ. Funct. Analysis, 203 (2003), 362-400. (math.OA/0202108)
7579:
7580: \bibitem{GuIso2} D.~Guido, T.~Isola, {\em Dimensions and spectral
7581: triples for fractals in $R^N$}, math.OA/0404295.
7582:
7583: \bibitem{gui} V.~Guillemin, {\it Lectures on spectral theory of
7584: elliptic operators}, Duke Math. J., Vol.44 (1977) N.3, 485--517.
7585:
7586: \bibitem{GuVai} D.~Gurevich and L.~Vainerman, {\it Non-commutative
7587: analogues of $q$-special polynomials and a $q$-integral on a quantum sphere}.
7588: J. Phys. A, 31 (1998) 7, 1771--1780.
7589:
7590: \bibitem{HaPa} E.~Ha, F.~Paugam, {\em Bost-Connes-Marcolli systems for
7591: Shimura varieties. I. Definitions and formal analytic properties},
7592: IMRP 2005, Vol.5 (2005) 237--286. (math.OA/0507101)
7593:
7594: \bibitem{Hajac} P.M.~Hajac, M.~Khalkhali, B.~Rangipour,
7595: Y.~Sommerhaeuser, {\em Hopf-cyclic homology and cohomology
7596: with coefficients}. math.KT/0306288.
7597:
7598: \bibitem{Hall}
7599: E.H.~Hall, {\it On a new action of the magnet on electric currents},
7600: Amer. J. of Math. 2 (1879) 287.
7601:
7602: \bibitem{EHaw} E.~Hawkins, {\em Hamiltonian gravity and noncommutative
7603: geometry}. Comm. Math. Phys. 187 (1997), no. 2, 471--489.
7604:
7605: \bibitem{Heisenberg} W.~Heisenberg, {\it \"Uber quantentheoretische
7606: Umdeutung kinematischer und
7607: mechanischer Beziehungen}, Z. Phys. 33 (1925) 879--893.
7608:
7609: \bibitem{HKR} G.~Hochschild, B.~Kostant, A.~Rosenberg, {\em Differential
7610: forms on regular affine algebras}. Trans. Amer. Math. Soc.
7611: Vol.102 (1962) 383--408.
7612:
7613: \bibitem{tHV} G.~'t Hooft, M.~Veltman, {\em Regularization and
7614: renormalization of gauge fields} Nuclear Physics B, Vol.44, N.1
7615: (1972), 189--213.
7616:
7617: \bibitem{Hu} S.~Hurder, {\em Secondary classes and the von Neumann algebra
7618: of a foliation}, {\it Preprint MSRI Berkeley} (1983).
7619:
7620: \bibitem{JLO} A.~Jaffe, A.~Lesniewski, K.~Osterwalder,
7621: {\em Quantum K-theory : the Chern character},
7622: Commun. Math. Phys., 118 (1988)
7623: 1--14.
7624:
7625: \bibitem{Julg} P.~Julg, {\em Travaux de N. Higson et
7626: G. Kasparov sur la conjecture de
7627: Baum-Connes}. Seminaire Bourbaki, (1997-98), Expose 841,
7628: Vol.252 (1998), 151--183.
7629:
7630: \bibitem{K-W} W.~Kalau and M.~Walze, {\it Gravity,
7631: noncommutative geometry and the Wod\-zi\-cki residue}.
7632: J. of Geom. and Phys., 16 (1995) 327--344.
7633:
7634: \bibitem{Kasp} G.G.~Kasparov, {\it The operator $K$-functor
7635: and extensions of $C^*$ algebras}. Izv. Akad. Nauk SSSR, Ser. Mat.,
7636: 44 (1980) 571--636; Math. USSR Izv., 16 (1981) 513--572.
7637:
7638: \bibitem{Kast} D.~Kastler, {\it The Dirac operator and
7639: gravitation}. Commun. Math. Phys., 166 (1995) 633--643.
7640:
7641: \bibitem{Kast2} D.~Kastler, {\em A detailed account of Alain Connes'
7642: version of the standard model in noncommutative geometry. I, II},
7643: Rev. Math. Phys. 5 (1993), no. 3, 477--532. {\em III},
7644: Rev. Math. Phys. 8 (1996), no. 1, 103--165.
7645:
7646: \bibitem{khal} M.~Khalkhali,{\em Very Basic Noncommutative
7647: Geometry}, arXiv:math.KT/0408416. Lecture Notes Series IPM Tehran,
7648: Vol.5, 2005.
7649:
7650: \bibitem{kr1} M.~Khalkhali and B.~Rangipour,
7651: {\em A new cyclic module for Hopf
7652: algebras}, $K$-Theory 27 (2002), no. 2, 111--131.
7653:
7654: \bibitem{kr2} M.~Khalkhali and B.~Rangipour, {\em Invariant cyclic homology},
7655: $K$-Theory 28 (2003), no. 2, 183--205.
7656:
7657: \bibitem{kr3} M.~Khalkhali and B.~Rangipour,
7658: {\em Cyclic cohomology of (extended) Hopf
7659: algebras},
7660: Noncommutative geometry and quantum groups,
7661: (Warsaw, 2001), 59--89, Banach Center Publ., 61,
7662: Polish Acad. Sci., Warsaw, 2003.
7663:
7664: \bibitem{KlimSchmu} A.~Klimyk and K.~Schmuedgen, {\em Quantum groups
7665: and their representations}, Springer Verlag, 1998.
7666:
7667: \bibitem{vonKlitz}
7668: K.~von Klitzing, G.~Dorda, M.~Pepper, {\em
7669: New method for high--accuracy determination of the fine--structure
7670: constant based on quantized hall resistance}.
7671: Phys. Rev. Lett. 45 (1980) N.6, 494--497.
7672:
7673: \bibitem{KontIHES97} M.~Kontsevich, {\it Deformation quantization of
7674: Poisson manifolds I}, IHES preprint M/97/72.
7675:
7676: \bibitem{Kontsevich93} M.~Kontsevich, {\it Formal (non)commutative
7677: symplectic geometry}.
7678: In: The Gel'fand Mathematical Seminars, 1990--1992, pages
7679: 173--187. Birkh\"auser Boston, Boston, MA, 1993.
7680:
7681: \bibitem{Kontsevich00} M.~Kontsevich and A.L.~Rosenberg, {\it
7682: Noncommutative smooth spaces}.
7683: In: The Gelfand Mathematical Seminars, 1996--1999, pages
7684: 85--108. Birkh\"auser Boston, Boston, MA, 2000.
7685:
7686: \bibitem{Laff} V.~Lafforgue,
7687: {\it K-th\'eorie bivariante pour les alg\`ebres de Banach et conjecture
7688: de Baum-Connes}, Inv. Math. Vol.149 (2002), 1--97.
7689:
7690: \bibitem{landi} G.~Landi, {\em An Introduction to
7691: Noncommutative Spaces and their
7692: Geometry}. Lecture Notes in Physics: Monographs m51,
7693: Springer-Verlag, 1997. ISBN 3-540-63509-2.
7694:
7695: \bibitem{LapFra} M.L.~Lapidus, M.~van Frankenhuysen, {\it Fractal
7696: geometry and number theory. Complex dimensions of fractal strings and
7697: zeros of zeta functions}. Birkh\"auser Boston, Inc., Boston, MA, 2000.
7698:
7699: \bibitem{Lau} B.~Laughlin, {\it Quantized Hall conductivity in two
7700: dimensions}. Phys. Rev. B, 23 (1981) 5632.
7701:
7702: \bibitem{laurent} H.~Laurent, {\it Sur la th\'eorie des nombres
7703: premiers}. C.R. Acad. Sci. Paris, 126 (1898) 809--810.
7704:
7705: \bibitem{Lesch} M.~Lesch, {\it Operators of Fuchs type, conical
7706: singularities, and asymptotic methods}, B. G. Teubner
7707: Verlagsgesellschaft mbH, Stuttgart, 1997.
7708:
7709: \bibitem{Lescure} J-M.~Lescure, {\em Triplets spectraux pour les
7710: vari\'et\'es \'a singularit\'e conique
7711: isol\'ee}. Bull. Soc. Math. France 129 (2001), no. 4, 593--623.
7712:
7713: \bibitem{ManDim} Yu.I.~Manin, {\it The notion of dimension in geometry
7714: and algebra}, to appear in Bulletin of the AMS (math.AG/0502016).
7715:
7716: \bibitem{Manin01RM} Yu.I.~Manin, {\it Real multiplication and
7717: noncommutative geometry}, The legacy of Niels Henrik Abel, 685--727,
7718: Springer, Berlin, 2004. (math.AG/0202109)
7719:
7720: \bibitem{Manin91NC} Yu.I.~Manin, {\em Topics in noncommutative geometry}.
7721: Princeton University Press, Princeton, NJ, 1991.
7722:
7723: \bibitem{Man-hyp} Yu.I.~Manin, {\em Three--dimensional hyperbolic geometry
7724: as $\infty$--adic Arakelov geometry}, Invent. Math. 104 (1991)
7725: 223--244.
7726:
7727: \bibitem{[M]} Yu.I.~Manin, {\it Quantum groups and
7728: noncommutative geometry}. Centre Recherche Math. Univ.
7729: Montr\'eal 1988.
7730:
7731: \bibitem{ManMar} Yu.I.~Manin, M.~Marcolli, {\em Continued fractions,
7732: modular symbols, and noncommutative geometry}. Selecta Math. (N.S.)
7733: 8 (2002), no. 3, 475--521.
7734:
7735: \bibitem{MaPa} Yu.I.~Manin, A.A.~Panchishkin, {\em Introduction to
7736: modern number theory. Fundamental problems, ideas and
7737: theories}. Second edition. Encyclopaedia of Mathematical Sciences,
7738: 49. Springer-Verlag, 2005. xvi+514 pp.
7739:
7740: \bibitem{Mar} M.~Marcolli, {\em Arithmetic noncommutative geometry},
7741: University Lecture Series, 36. American Mathematical Society, 2005.
7742: xii+136 pp.
7743:
7744: \bibitem{Mar2} M.~Marcolli, {\em Limiting modular symbols and the
7745: Lyapunov spectrum}. J. Number Theory 98 (2003), no. 2, 348--376.
7746:
7747: \bibitem{Mar3} M.~Marcolli, {\em Modular curves, $C^*$-algebras, and
7748: chaotic cosmology}, math-ph/0312035.
7749:
7750: \bibitem{MM1} M.~Marcolli and V.~Mathai, {\em Twisted index theory on
7751: good orbifolds, II: fractional quantum numbers}, Communications in
7752: Mathematical Physics, Vol.217, no.1 (2001) 55--87.
7753:
7754: \bibitem{MM2} M.~Marcolli and V.~Mathai, {\em Twisted index theory on good
7755: orbifolds, I: noncommutative Bloch theory}, Communications in
7756: Contemporary Mathematics, Vol.1 (1999) 553--587.
7757:
7758: \bibitem{MM3} M.~Marcolli and V.~Mathai, {\em Towards the fractional
7759: quantum Hall effect: a noncommutative geometry perspective},
7760: cond-mat/0502356.
7761:
7762: \bibitem{RoMa} V.~Mathai, J.~Rosenberg, {\em $T$-duality for
7763: torus bundles with $H$-fluxes via noncommutative topology}.
7764: Comm. Math. Phys. 253 (2005), no. 3, 705--721
7765:
7766: \bibitem{mather} J.~Mather, {\it Commutators of diffeomorphisms.
7767: II}, Comment. Math. Helv. 50 (1975), 33--40.
7768:
7769: \bibitem{Meyer} R.~Meyer, {\em On a representation of the idele class
7770: group related to primes and zeros of $L$-functions}, preprint arXiv
7771: math.NT/0311468.
7772:
7773: \bibitem{Mil2} J.S.~Milne, {\em Shimura varieties: the geometric
7774: side of the zeta function}, preprint available at
7775: www.jmilne.org
7776:
7777: \bibitem{Mi-S} J.~Milnor and D.~Stasheff, {\it Characteristic
7778: classes}. Ann. of Math. Stud., Princeton
7779: University Press, Princeton, N.J. 1974.
7780:
7781: \bibitem{Month} B.~Monthubert, {\it
7782: Pseudodifferential calculus on manifolds with corners and groupoids}
7783: Proc. Amer. Math. Soc. 127 (1999) no 10, 2871--2881. ArXiv
7784: math.KT/0507601.
7785:
7786: \bibitem{Month1} B.~Monthubert and V.~Nistor, {\it
7787: A topological index theorem for manifolds with corners}
7788:
7789: \bibitem{Nagy} D.~Nagy, {\em Visual mathematics: a missing link in a
7790: split culture}. Vis. Math. 1 (1999), no. 2, 20 pp. (electronic).
7791:
7792: \bibitem{NS} N.~Nekrasov and A.~Schwarz, {\it Instantons in
7793: noncommutative $\R^4$ and (2,0) superconformal six dimensional
7794: theory}. Comm. Math. Phys. 198 (1998), no. 3, 689--703.
7795:
7796: \bibitem{PaTu} W.~Parry, S.~Tuncel, {\it Classification problems in
7797: ergodic theory}, London Math. Soc. Lecture Notes Series 67, 1982.
7798:
7799: \bibitem{Penrose} R.~Penrose, {\em Pentaplexity: a class of
7800: nonperiodic tilings of the plane}. Math. Intelligencer 2 (1979/80),
7801: no. 1, 32--37.
7802:
7803: \bibitem{pims:1997}
7804: M.~V.~Pimsner, {\em A class of C*-algebra
7805: generalizing both Cuntz-Krieger algebras and crossed products by
7806: $Z$}. Free probablility Theory, 189-212. Fields Inst. Comm. Vol.12,
7807: American Math. Society, 1997.
7808:
7809: \bibitem{PV} M.~Pimsner and D.~Voiculescu, {\it Exact sequences
7810: for $K$ groups and Ext group of certain crossed product
7811: $C^*$-algebras}. J. Operator Theory, 4 (1980) 93--118.
7812:
7813: \bibitem{Poli} A.~Polishchuk, {\em Noncommutative two-tori with real
7814: multiplication as noncommutative projective varieties}.
7815: J. Geom. Phys. 50 (2004), no. 1-4, 162--187.
7816:
7817: \bibitem{Pusch} M.~Puschnigg, {\em Characters of Fredholm modules
7818: and a problem of Connes}, Preprint 2005.
7819:
7820: \bibitem{Putn} I.~Putnam, {\it $C^*$--algebras from Smale spaces},
7821: Can. J. Math. 48 (1996) N.1 175--195.
7822:
7823: \bibitem{PutSpi} I.~Putnam, J.~Spielberg, {\it The structure of ${\rm
7824: C}^*$--algebras associated with hyperbolic dynamical systems},
7825: J. Funct. Anal. 163 (1999) 279--299.
7826:
7827: \bibitem{Ramis} J-P.~Ramis, {\it S\'eries divergentes et th\'eories
7828: asymptotiques}. Bull. Soc. Math. France, 121(Panoramas et Syntheses,
7829: suppl.) 74, 1993.
7830:
7831: \bibitem{Ren} J.~Renault, {\em A groupoid approach to $C^*$-algebras}.
7832: Lecture Notes in Math. 793, Springer, Berlin, 1980.
7833:
7834: \bibitem{Ri2} M.A.~Rieffel, {\it $C^*$-algebras associated
7835: with irrational rotations}. Pacific J. Math., 93 (1981)
7836: 415--429.
7837:
7838: \bibitem{Rieffel} M.A.~Rieffel, {\it The cancellation theorem
7839: for projective modules over irrational rotation $C^*$-algebras}.
7840: Proc. London Math. Soc., 47 (1983) 285--302.
7841:
7842: \bibitem{Rieffel89} M.A.~Rieffel, {\it Deformation quantization of
7843: Heisenberg manifolds}. Comm. Math. Phys. 122 (1989) 4, 531--562.
7844:
7845: \bibitem{Rob} G.~Robertson, {\it Boundary actions for affine buildings
7846: and higher rank Cuntz--Krieger algebras}, in J.~Cuntz, S.~Echterhoff
7847: (eds.) ``${\rm C}^*$--algebras'', Springer Verlag 2001, pp.~ 182--202.
7848:
7849: \bibitem{Rosenberg98} A.L.~Rosenberg, {\em Noncommutative schemes}.
7850: Compositio Math. 112 (1998) N.1, 93--125.
7851:
7852: \bibitem{Rosenberg95} A.L.~Rosenberg.
7853: {\em Noncommutative algebraic geometry and representations of
7854: quantized algebras}.
7855: Kluwer Academic Publishers Group, Dordrecht, 1995.
7856:
7857: \bibitem{Rue2} D.~Ruelle, {\it Non--commutative algebras for
7858: hyperbolic diffeomorphisms}, Invent. Math. 93 (1988) 1--13.
7859:
7860: \bibitem{Schu} T.~Schucker, {\it Spin group and almost
7861: commutative geometry}. preprint hep-th/0007047.
7862:
7863: \bibitem{Witten} N.~Seiberg and E.~Witten, {\it String theory
7864: and noncommutative geometry}. J. High Energy Phys. 1999, no. 9, Paper
7865: 32, 93 pp.
7866:
7867: \bibitem{Shahin} M.M.~Sheikh-Jabbari,
7868: {\em Tiny Graviton Matrix Theory: DLCQ of IIB Plane-Wave
7869: String Theory, A Conjecture}, JHEP 0409 (2004) 017.
7870: hep-th/0406214.
7871:
7872: \bibitem{Shi} G.~ Shimura, {\em Arithmetic theory of automorphic
7873: functions}, Iwanami Shoten and Princeton 1971.
7874:
7875:
7876: \bibitem{Skan1} G.~Skandalis, {\em Approche de la conjecture de Novikov
7877: par la
7878: cohomologie cyclique}. Seminaire Bourbaki, (1990-91), Expose 739,
7879: Ast\'erisque No.201-203 (1992), 299--320.
7880:
7881: \bibitem{Skan2} G.~Skandalis, {\em Progr\`es r\'ecents
7882: sur la conjecture de Baum-Connes.
7883: Contribution de Vincent Lafforgue}, S\'eminaire Bourbaki, Vol.1999-2000,
7884: Expose 869, Ast\'erisque No. 276 (2002), 105--135.
7885:
7886: \bibitem{skau} C.~ Skau,
7887: {\it Orbit structure of topological dynamical systems and
7888: its invariants}, Operator algebras and quantum field theory (Rome,
7889: 1996), 533--544, Internat. Press, Cambridge, MA, 1997.
7890:
7891: \bibitem{smi-sta:1992}
7892: S.P.~Smith and J.T. Stafford.
7893: {\em Regularity of the four dimensional Sklyanin algebra}.
7894: Compos. Math., Vol.83 (1992) 259--289.
7895:
7896: \bibitem{Soibelman01} Y.~Soibelman, {\it Quantum tori, mirror symmetry
7897: and deformation theory}. Lett. Math. Phys. 56 (2001) 99--125.
7898:
7899: \bibitem{Spi} J.~Spielberg, {\it Free--product groups, Cuntz--Krieger
7900: algebras, and covariant maps}. Internat. J. Math. 2 (1991), no. 4,
7901: 457--476.
7902:
7903: \bibitem{SDLSV} W.~van Suijlekom, L.~Dabrowski, G.~Landi, A.~Sitarz,
7904: J.C.~Varilly, {\em Local index formula for
7905: $SU_q(2)$}, math.QA/0501287
7906:
7907: \bibitem{SDLSV2} W.~van Suijlekom, L.~Dabrowski, G.~Landi, A.~Sitarz,
7908: J.C.~Varilly, {\em The Dirac operator on
7909: $SU_q(2)$}, math.QA/0411609
7910:
7911: \bibitem{Sull} D.~Sullivan, {\it Geometric periodicity and
7912: the invariants of manifolds}. Lecture Notes in Math., 197,
7913: Springer, 1971.
7914:
7915: \bibitem{Sun} T.~Sunada,
7916: {\em A discrete analogue of periodic magnetic Schr\"odinger operators},
7917: Contemp. Math. Vol.173 (1994) 283--299.
7918:
7919: \bibitem{Tak} M.~Takesaki, {\it Tomita's theory of modular
7920: Hilbert algebras and its applications}. Lecture Notes in Math., 28,
7921: Springer, 1970.
7922:
7923: \bibitem{Velt2} M.~Veltman, {\em Diagrammatica}, Cambridge
7924: University Press, 1994.
7925:
7926: \bibitem{vdWaerden} B.L.~van der Waerden,
7927: {\em Sources of Quantum Mechanics}. Dover, New York, 1967.
7928:
7929: \bibitem{Wafa} Abu'l-Wafa' al-Buzjani, {\em Kitab fima yahtaju ilayhi
7930: al-sani` min a`mal al-handasa}. Milan manuscript (Arabic): Biblioteca
7931: Ambrosiana, Arab. 68.
7932:
7933: \bibitem{Wassermann88} A.~Wassermann, {\em Ergodic actions of compact
7934: groups on operator
7935: algebras. III. Classification for ${\rm SU}(2)$}.
7936: Invent. Math. 93 (1988) 2, 309--354.
7937:
7938: \bibitem{Winkel} H.E.~Winkelnkemper,
7939: {\it The graph of a foliation}, Ann. Global Anal.
7940: Geom. 1 (1983), no 3, 51--75.
7941:
7942: \bibitem{Wit1} E.~Witten, {\em Bound states of strings and p-Branes},
7943: Nuclear Physics B 460 (1996) 335--360.
7944:
7945: \bibitem{Wo} M.~Wodzicki, {\it Noncommutative residue, Part
7946: I. Fundamentals}. In $K$-theory, arithmetic and geometry.
7947: Lecture Notes in Math., 1289, Springer, Berlin 1987.
7948:
7949: \bibitem{Zagier} D.~Zagier, {\em Modular forms and differential
7950: operators}, Proc. Indian Acad. Sci. Math. Sci. (K.G.Ramanathan
7951: memorial issue) 104 (1994) N.1, 57--75.
7952:
7953: \end{thebibliography}
7954: \end{document}
7955: \end
7956: