Lectures On Arithmetic Noncommutative Geometry
Lectures On Arithmetic Noncommutative Geometry
Lectures On Arithmetic Noncommutative Geometry
Noncommutative Geometry
Matilde Marcolli
And indeed there will be time
To wonder “Do I dare?” and, “Do I dare?”
Time to turn back and descend the stair.
...
Do I dare
Disturb the Universe?
...
For I have known them all already, known them all;
Have known the evenings, mornings, afternoons,
I have measured out my life with coffee spoons.
...
I should have been a pair of ragged claws
Scuttling across the floors of silent seas.
...
No! I am not Prince Hamlet, nor was meant to be;
Am an attendant lord, one that will do
To swell a progress, start a scene or two
...
At times, indeed, almost ridiculous–
Almost, at times, the Fool.
...
We have lingered in the chambers of the sea
By sea-girls wreathed with seaweed red and brown
Till human voices wake us, and we drown.
Chapter 1. Ouverture 5
1. The NCG dictionary 7
2. Noncommutative spaces 8
3. Spectral triples 9
Chapter 2. Noncommutative modular curves 17
1. Modular curves 17
2. The noncommutative boundary of modular curves 24
3. Modular interpretation: noncommutative elliptic curves 24
4. Limiting modular symbols 29
5. Hecke eigenforms 40
6. Selberg zeta function 43
7. The modular complex and K-theory of C∗ -algebras 44
8. Intermezzo: Chaotic Cosmology 45
Chapter 3. Quantum statistical mechanics and Galois theory 53
1. Quantum Statistical Mechanics 54
2. The Bost–Connes system 58
3. Noncommutative Geometry and Hilbert’s 12th problem 63
4. The GL2 system 65
5. Quadratic fields 72
Chapter 4. Noncommutative geometry at arithmetic infinity 81
1. Schottky uniformization 81
2. Dynamics and noncommutative geometry 88
3. Arithmetic infinity: archimedean primes 93
4. Arakelov geometry and hyperbolic geometry 96
5. Intermezzo: Quantum gravity and black holes 100
6. Dual graph and noncommutative geometry 104
7. Arithmetic varieties and L–factors 109
8. Archimedean cohomology 115
Bibliography 125
3
CHAPTER 1
Ouverture
many proofs have not been included in the text, the reader will find
references to the relevant literature, where complete proofs are provided
(in particular [29], [36], and [70]).
More explicitly, the text is organized as follows:
• We start by recalling a few preliminary notions of noncommu-
tative geometry (following [20]).
• The second chapter describes how various arithmetic proper-
ties of modular curves can be seen by their “noncommutative
boundary”. This part is based on the joint work of Yuri Manin
and the author. The main references are [70], [71], [72].
• We review briefly the work of Connes and the author [29] on
the noncommutative geometry of commensurability classes of
Q-lattices. The relation of the noncommutative space of com-
mensurability classes of Q-lattices to the Hilbert 12th prob-
lem of explicit class field theory is based on ongoing work of
Connes, Ramachandran and the author [31], on the original
work of Bost and Connes [8] and on Manin’s real multiplica-
tion program [62] [63].
• The noncommutative geometry of the fibers at “arithmetic in-
finity” of varieties over number fields is based on joint work of
Consani and the author, for which the references are [36], [37],
[38], [39], [40]. This chapter also contains a detailed account
of Manin’s formula for the Green function of Arakelov geom-
etry for arithmetic surfaces, based on [66], and a proposed
physical interpretation of this formula, as in [69].
Similarly, at the level of topology one often sees phenomena that are
closer to rigid analytic geometry. This is the case, for instance, with the
noncommutative tori Tθ , which already at the C ∗ -algebra level exhibit
moduli that behave much like moduli of one-dimensional complex tori
(elliptic curves) in the commutative case.
In the context we are going to discuss this richer structure of non-
commutative spaces is crucial, as it permits us to use tools like C ∗ -
algebras (topology) to study the properties more rigid spaces like alge-
braic or arithmetic varieties.
2. Noncommutative spaces
The way to assign the algebra of coordinates to a quotient space
X = Y / ∼ can be explained in a short slogan as follows:
• Functions on Y with f (a) = f (b) for a ∼ b. Poor!
• Functions fab on the graph of the equivalence relation. Good!
The second description leads to a noncommutative algebra, as the
product, determined by the groupoid law of the equivalence relation,
has the form of a convolution product (like the product of matrices).
For sufficiently nice quotients, even though the two notions are
not the same, they are related by Morita equivalence, which is the
suitable notion of “isomorphism” between noncommutative spaces. For
more general quotients, however, the two notions truly differ and the
second one is the only one that allows one to continue to make sense
of geometry on the quotient space.
A very simple example illustrating the above situation is the fol-
lowing (cf. [23]). Consider the topological space Y = [0, 1] × {0, 1}
with the equivalence relation (x, 0) ∼ (x, 1) for x ∈ (0, 1). By the first
method one only obtains constant functions C, while by the second
method one obtains
{f ∈ C([0, 1]) ⊗ M2 (C) : f (0) and f (1) diagonal }
which is an interesting nontrivial algebra.
The idea of preserving the information on the structure of the equiv-
alence relation in the description of quotient spaces has analogs in
Grothendieck’s theory of stacks in algebraic geometry.
2.1. Morita equivalence. In noncommutative geometry, isomor-
phisms of C ∗ -algebras are too restrictive to provide a good notion of
isomorphisms of noncommutative spaces. The correct notion is pro-
vided by Morita equivalence of C ∗ -algebras.
We have equivalent C∗ -algebras A1 ∼ A2 if there exists a bimodule
M, which is a right Hilbert A1 module with an A1 -valued inner product
3. SPECTRAL TRIPLES 9
h·, ·iA1 , and a left Hilbert A2 -module with an A2 -valued inner product
h·, ·iA2 , such that we have:
• We obtain all Ai as the closure of the span of
{hξ1 , ξ2 iAi : ξ1 , ξ2 ∈ M}.
• ∀ξ1 , ξ2 , ξ3 ∈ M we have
hξ1 , ξ2 iA1 ξ3 = ξ1 hξ2 , ξ3 iA2 .
• A1 and A2 act on M by bounded operators,
ha2 ξ, a2 ξiA1 ≤ ka2 k2 hξ, ξiA1 ha1 ξ, a1 ξiA2 ≤ ka1 k2 hξ, ξiA2
for all a1 ∈ A1 , a2 ∈ A2 , ξ ∈ M.
This notion of equivalence roughly means that one can transfer
modules back and forth between the two algebras.
2.2. The tools of noncommutative geometry. Once one iden-
tifies in a specific problem a space that, by its nature of quotient of the
type described above, is best described as a noncommutative space,
there is a large set of well developed techniques that one can use to
compute invariants and extract essential information from the geome-
try. The following is a list of some such techniques, some of which will
make their appearance in the cases treated in these notes.
• Topological invariants: K-theory
• Hochschild and cyclic cohomology
• Homotopy quotients, assembly map (Baum-Connes)
• Metric structure: Dirac operator, spectral triples
• Characteristic classes, zeta functions
We will recall the necessary notions when needed. We now begin
by taking a closer look at the analog in the noncommutative world of
Riemannian geometry, which is provided by Connes’ notion of spectral
triples.
3. Spectral triples
Spectral triples are a crucial notion in noncommutative geometry.
They provide a powerful and flexible generalization of the classical
structure of a Riemannian manifold. The two notions agree on a com-
mutative space. In the usual context of Riemannian geometry, the
definition of the infinitesimal element ds on a smooth spin manifold
can be expressed in terms of the inverse of the classical Dirac opera-
tor D. This is the key remark that motivates the theory of spectral
triples. In particular, the geodesic distance between two points on the
manifold is defined in terms of D −1 (cf. [20] §VI). The spectral triple
10 1. OUVERTURE
Here the notation limω takes into account the fact that the sequence
k
1 X
S(k, T ) := λj (T )
log k j=1
defined as
1
Z
(1.3) Trω (a|D|−n ).
a :=
V
The usual notion of integration on a Riemannian spin manifold M
can be recovered in this context (cf. [20], [58]) through the formula (n
even): Z
f dv = 2n−[n/2]−1 π n/2 nΓ(n/2) Trω (f |D|−n ).
M
Here D is the classical Dirac operator on M associated to the metric
that determines the volume form dv, and f in the right hand side is
regarded as the multiplication operator acting on the Hilbert space of
square integrable spinors on M .
3.2. Zeta functions. An important function associated to the
Dirac operator D of a spectral triple (A, H, D) is its zeta function
X
(1.4) ζD (z) := Tr(|D|−z ) = Tr(Π(λ, |D|))λ−z ,
λ
These zeta functions are related to the heat kernel e−t|D| by Mellin
transform
Z ∞
1
(1.8) ζa,D (z) = tz−1 Tr(a e−t|D| ) dt
Γ(z) 0
where
X
(1.9) Tr(a e−t|D| ) = Tr(a Π(λ, |D|))e−tλ =: θa,D (t).
λ
3. SPECTRAL TRIPLES 13
Similarly,
∞
1
Z
(1.10) ζa,D (s, z) = θa,D,s (t) tz−1 dt
Γ(z) 0
with
X
(1.11) θa,D,s (t) := Tr(a Π(λ, |D|))e(s−λ)t .
λ
The results of this section are mostly based on the joint work of
Yuri Manin and the author [70], with additional material from [71]
and [72]. We include necessary preliminary notions on modular curves
and on noncommutative tori, respectively based on the papers [68] and
[15], [85].
We first recall some aspects of the classical theory of modular
curves. We will then show that these notions can be entirely recovered
and enriched through the analysis of a noncommutative space associ-
ated to a compactification of modular curves, which includes noncom-
mutative tori as possible degenerations of elliptic curves.
1. Modular curves
Let G be a finite index subgroup of the modular group Γ = PSL(2, Z),
and let XG denote the quotient
(2.1) XG := G\H2
|q|
1
Figure 4. The fundamental domain for the Jacobi uni-
formization of the elliptic curve
We set
pn (α) := Pn (k1 (α), . . . , kn (α)), qn (α) := Qn (k1 (α), . . . , kn (α))
so that pn (α)/qn (α) is the sequence of convergents to α. We also set
pn−1 (α) pn (α)
gn (α) := ∈ GL(2, Z).
qn−1 (α) qn (α)
Written in terms of the continued fraction expansion, the shift T is
given by
T : [k0 , k1 , k2 , . . .] 7→ [k1 , k2 , k3 , . . .].
The properties of the shift (2.24) can be used to extend the no-
tion of modular symbols to geodesics with irrational ends ([70]). Such
geodesics correspond to infinite geodesics on the modular curve XG ,
which exhibit a variety of interesting possible behaviors, from closed
geodesics to geodesics that approximate some limiting cycle, to geodesics
that wind around different homology class exhibiting a typically chaotic
behavior.
These level sets are uncountable dense T –invariant subsets of [0, 1], of
varying Hausdorff dimension [83]. The Lyapunov spectrum measures
how the Hausdorff dimension varies, as a function h(c) = dimH (Lc ).
ζn
zn
z n+1
ζn+1
p p
α n-1
p
n+1 β n
q
q q n
n-1 n+1
where we take h(x, t) = −2σ log |T 0 (x, t)|. Clearly this operator is well
suited for capturing the dynamical properties of the map T as it is
defined as a weighted sum over preimages. On the other hand, there
is another operator that can be associated to a dynamical system and
which typically has better spectral properties, but is less clearly related
to the dynamics. The best circumstances are when these two agree (for
a particular value of the parameter). The other operator is the Perron–
Frobenius operator P. This is defined by the relation
Z Z
(2.32) f (g ◦ T ) dµLeb = (P f ) g dµLeb .
[0,1]×P [0,1]×P
1 X̀ −1
= {gk (β) · 0, gk−1 (β) · i∞}G .
λ(β)`
k=1
This shows that, in this case, the limiting modular symbols are
linear combinations of classical modular symbols, with coefficients in
the field generated over Q by the Lyapunov exponents λ(β) of the
quadratic irrationalities.
In terms of geodesics on the modular curve, this is the case where
the geodesic has a limiting cycle given by the closed geodesic {0, g(0)}G
(Figure 6).
There is then the “generic case”, where, contrary to the previous
example, the geodesics wind around many different cycles in such a
way that the resulting homology class averages out to zero over long
distances (Figure 7).
4. LIMITING MODULAR SYMBOLS 39
this yields
Z Z Z
{{∗, β}} f (β, t) dµ(β, t) = f dµ ϕ h dµ
[0,1]×P [0,1]×P [0,1]×P
Z
1 X
= f dµ ϕ(s).
2#P s∈P
P
It is then easy to check that the sum s∈P ϕ(s) = 0 since each term in
the sum changes sign under the action of the inversion σ ∈ PGL(2, Z)
with σ 2 = 1, but the sum is globally invariant under σ.
The argument extends to the case of other T -invariant subsets of
[0, 1] × P for which the corresponding Perron–Frobenius operator LE,δE
has analogous properties (cf. [71]). The weak convergence proved
by this type of argument can be improved to strong convergence by
applying the strong law of large numbers to the “random variables”
ϕk = ϕ ◦ T k (cf. [71] for details). The result effectively plays the role
of an ergodic theorem for the shift T on E.
5. Hecke eigenforms
A very important question is what happens to modular forms at the
noncommutative boundary of the modular curves. There is a variety
of phenomena in the theory of modular forms that hint to the fact that
a suitable class of “modular forms” survives on the noncommutative
boundary. Zagier introduced the term “quantum modular forms” to
denote this important and yet not sufficiently understood class of exam-
ples. Some aspects of modular forms “pushed” to the noncommutative
boundary were analyzed in [70], in the form of certain averages involv-
ing modular symbols and Dirichlet series related to modular forms of
weight 2. We recall here the main results in this case.
We shall now consider the case of congruence subgroups G = Γ0 (p),
for p a prime.
Let ω be a holomorphic differential (also called a differential of the
first kind) on the modular curve XΓ0 (p) = Γ0 (p)\H. Let Φ = ϕ∗ (ω)/dz
be the pullback under the projection ϕ : H → XΓ0 (p) .
Let Φ be an eigenfunction for all the Hecke operators
d−1
XX n/d b
Tn =
0 d
d|n b=0
with (p, n) = 1,
Tn Φ = cn Φ.
5. HECKE EIGENFORMS 41
one obtains Z 1
`(f, β) dµLeb (β) =
0
" #Z
(p) i∞
ζ(1 + t) Lω (2 + t)
− (p) φ∗ (ω),
ζ(2 + t) ζ (2 + t)2 0
This defines, for <(t) > 0 and for almost all β, a homology class in
H1 (XG , cusps, R) such that the integral average
Z
`(f, β) = ω
C(f,β)
An estimate
n
qn+1 (β) + qn (β) qn (β) −(5+2t)nλ(β)
X
0, ∼e ϕ ◦ T k (s)
qn+1 (β)1+t qn+1 (β) G k=1
6. SELBERG ZETA FUNCTION 43
K1 (A o Γ) o K1 (A o Γσ ) ⊕ K1 (A o Γτ ) o K1 (A)
β̃ β
Since for fixed u the model is given by a Kasner space-time, the behav-
ior of this universe can be approximated for certain large intervals of
time by a Kasner metric. In fact, the evolution is divided into Kasner
eras and each era into cycles. During each era the mixmaster universe
goes through a volume compression. Instead of resulting in a collapse,
as with the Kasner metric, high negative curvature develops resulting
in a bounce (transition to a new era) which starts again a behavior
approximated by a Kasner metric, but with a different value of the
parameter u. Within each era, most of the volume compression is due
8. INTERMEZZO: CHAOTIC COSMOLOGY 47
to the scale factors along one of the space axes, while the other scale
factors alternate between phases of contraction and expansion. These
alternating phases separate cycles within each era.
More precisely, we are considering a metric
(2.44) ds2 = −dt2 + a(t)dx2 + b(t)dy 2 + c(t)dz 2 ,
generalizing the Kasner metric (2.42). We still require that (2.44)
admits SO(3) symmetry on the space-like hypersurfaces, and that it
dt
presents a singularity at t → 0. In terms of logarithmic time dΩ = − abc ,
the mixmaster universe model of Belinsky, Khalatnikov, and Lifshitz
admits a discretization with the following properties:
1. The time evolution is divided in Kasner eras [Ωn , Ωn+1 ], for n ∈ Z.
At the beginning of each era we have a corresponding discrete value of
the parameter un > 1 in (2.43).
2. Each era, where the parameter u decreases with growing Ω, can
be subdivided in cycles corresponding to the discrete steps un , un − 1,
un − 2, etc. A change u → u − 1 corresponds, after acting with the
permutation (12)(3) on the space coordinates, to changing u to −u,
hence replacing contraction with expansion and conversely. Within
each cycle the space-time metric is approximated by the Kasner metric
(2.42) with the exponents pi in (2.43) with a fixed value of u = un −k >
1.
3. An era ends when, after a number of cycles, the parameter un
falls in the range 0 < un < 1. Then the bouncing is given by the
transition u → 1/u which starts a new series of cycles with new Kasner
parameters and a permutation (1)(23) of the space axis, in order to
have again p1 < p2 < p3 as in (2.43).
Thus, the transition formula relating the values un and un+1 of two
successive Kasner eras is
1
un+1 = ,
un − [un ]
which is exactly the shift of the continued fraction expansion, T x =
1/x − [1/x], with xn+1 = T xn and un = 1/xn .
The previous observation is the key to a geometric description of
solutions of the mixmaster universe in terms of geodesics on a modular
curve (cf. [70]).
Theorem 8.1. Consider the modular curve XΓ0 (2) . Each infinite
geodesic γ on XΓ0 (2) not ending at cusps determines a mixmaster uni-
verse.
48 2. NONCOMMUTATIVE MODULAR CURVES
In fact, an infinite geodesic on XΓ0 (2) is the image under the quotient
map
πΓ : H2 × P → Γ\(H2 × P) ∼ = XG ,
where Γ = PGL(2, Z), G = Γ0 (2), and P = Γ/G ∼ = P1 (F2 ) = {0, 1, ∞},
of an infinite geodesic on H2 ×P with ends on P1 (R)×P. We consider the
elements of P1 (F2 ) as labels assigned to the three space axes, according
to the identification
0 = [0 : 1] 7→ z
(2.45) ∞ = [1 : 0] 7→ y
1 = [1 : 1] 7→ x.
As we have seen, geodesics can be coded in terms of data (ω − , ω + , s)
with the action of the shift T .
The data (ω, s) determine a mixmaster universe, with the kn =
[un ] = [1/xn ] in the Kasner eras, and with the transition between
subsequent Kasner eras given by xn+1 = T xn ∈ [0, 1] and by the per-
mutation of axes induced by the transformation
−kn 1
1 0
acting on P1 (F2 ). It is easy to verify that, in fact, this acts as the
permutation 0 7→ ∞, 1 7→ 1, ∞ 7→ 0, if kn is even, and 0 7→ ∞,
1 7→ 0, ∞ 7→ 1 if kn is odd, that is, after the identification (2.45), as
the permutation (1)(23) of the space axes (x, y, z), if kn is even, or as
the product of the permutations (12)(3) and (1)(23) if kn is odd. This
is precisely what is obtained in the mixmaster universe model by the
repeated series of cycles within a Kasner era followed by the transition
to the next era.
Data (ω, s) and T m (ω, s), m ∈ Z, determine the same solution up
to a different choice of the initial time.
There is an additional time-symmetry in this model of the evolu-
tion of mixmaster universes (cf. [5]). In fact, there is an additional
parameter δn in the system, which measures the initial amplitude of
each cycle. It is shown in [5] that this is governed by the evolution of
a parameter
δn+1 (1 + un )
vn =
1 − δn+1
which is subject to the transformation across cycles vn+1 = [un ] + vn−1 .
By setting yn = vn−1 we obtain
1
yn+1 = ,
(yn + [1/xn ])
8. INTERMEZZO: CHAOTIC COSMOLOGY 49
hence we see that we can interpret the evolution determined by the data
(ω + , ω − , s) with the shift T either as giving the complete evolution of
the u-parameter towards and away from the cosmological singularity,
or as giving the simultaneous evolution of the two parameters (u, v)
while approaching the cosmological singularity.
This in turn determines the complete evolution of the parameters
(u, δ, Ω), where Ωn is the starting time of each cycle. For the explicit
recursion Ωn+1 = Ωn+1 (Ωn , xn , yn ) see [5].
The result of Theorem 4.5 on the unique T -invariant measure on
[0, 1] × P
δ(s) dx
(2.46) dµ(x, s) = ,
3 log(2) (1 + x)
implies that the alternation of the space axes is uniform over the time
evolution, namely the three axes provide the scale factor responsible
for volume compression with equal frequencies.
The Perron-Frobenius operator for the shift (2.24) yields the density
of the invariant measure (2.46) satisfying L1 f = f . The top eigenvalue
ησ of Lσ is related to the topological pressure by ησ = exp(P (σ)).
This can be estimated numerically, using the technique of [4] and the
integral kernel operator representation of §1.3 of [70].
The interpretation of solutions in terms of geodesics provides a nat-
ural way to single out and study certain special classes of solutions on
the basis of their geometric properties. Physically, such special classes
of solutions exhibit different behaviors approaching the cosmological
singularity.
For instance, the data (ω + , s) corresponding to an eventually peri-
odic sequence k0 k1 . . . km . . . of some period a0 . . . a` correspond to those
geodesics on XΓ0 (2) that asymptotically wind around the closed geo-
desic identified with the doubly infinite sequence . . . a0 . . . a` a0 . . . a` . . ..
Physically, these universes exhibit a pattern of cycles that recurs peri-
odically after a finite number of Kasner eras.
Another special class of solutions is given by the Hensley Cantor
sets (cf. [72]). These are the mixmaster universes for which there is a
fixed upper bound N to the number of cycles in each Kasner era.
In terms of the continued fraction description, these solutions cor-
respond to data (ω + , s) with ω + in the Hensley Cantor set EN ⊂ [0, 1].
The set EN is given by all points in [0, 1] with all digits in the continued
fraction expansion bounded by N (cf. [51]). In more geometric terms,
these correspond to geodesics on the modular curve XΓ0 (2) that wander
only a finite distance into a cusp.
50 2. NONCOMMUTATIVE MODULAR CURVES
general case
invertible case
Im z = β
iβ F(t + iβ) = ϕ(σt(b)a)
F(t) = ϕ(aσt(b))
Im z = 0 0
Our definition of KMS∞ state is stronger than the one often adopted
in the literature, which simply uses the existence, for each a, b ∈ A, of
a bounded holomorphic function Fa,b (z) on the upper half plane such
that Fa,b (t) = ϕ(aσt (b)). It is easy to see that this notions, which we
1. QUANTUM STATISTICAL MECHANICS 57
simply call the ground states of the system, is in fact weaker than the
notion of KMS∞ states given in Definition 1.2. For example, in the
simplest case of the trivial time evolution, all states are ground states,
while only tracial states are KMS∞ states. Another advantage of our
definition is that for all 0 < β ≤ ∞, the KMSβ states form a Choquet
simplex and we can therefore consider the set Eβ of its extremal points.
These give a good notion of points for the underlying noncommutative
space. This will be especially useful in connection to an arithmetic
structure specified by an arithmetic subalgebra AQ of A on which one
evaluates the KMS∞ states. This will play a key role in the relation
between the symmetries of the system and the action of the Galois
group on states ϕ ∈ E∞ evaluated on AQ .
1.1. Symmetries. An important role in quantum statistical me-
chanics is played by symmetries. Typically, symmetries of the algebra
A compatible with the time evolution induce symmetries of the equi-
librium states Eβ at different temperatures. Especially important are
the phenomena of symmetry breaking. In such cases, there is a global
underlying group G of symmetries of the algebra A, but in certain
ranges of temperature the choice of an equilibrium state ϕ breaks the
symmetry to a smaller subgroup Gϕ = {g ∈ G : g ∗ ϕ = ϕ}, where
g ∗ denotes the induced action on states. Various systems can exhibit
one or more phase transitions, or none at all. A typical situation in
physical systems sees a unique KMS state for all values of the parame-
ter above a certain critical temperature (β < βc ). This corresponds to
a chaotic phase such as randomly distributed spins in a ferromagnet.
When the system cools down and reaches the critical temperature, the
unique equilibrium state branches off into a larger set KMSβ and the
symmetry is broken by the choice of an extremal state in Eβ .
We will see that the case of L1 /R∗+ gives rise to a system with a
single phase transition ([8]), while in the case of L2 /C∗ the system has
multiple phase transitions.
A very important point is that we need to consider both symmetries
by automorphisms and by endomorphisms.
Automorphisms: A subgroup G ⊂ Aut(A) is compatible with σt if
for all g ∈ G and for all t ∈ R we have gσt = σt g. There is then an
induced action of G on KMS states and in particular on the set Eβ . If
u is a unitary, acting on A by
Adu : a 7→ uau∗
and satisfying σt (u) = u, then we say that Adu is an inner automor-
phism of (A, σt ). Inner automorphisms act trivially on KMS states.
58 3. QUANTUM STATISTICAL MECHANICS AND GALOIS THEORY
...
...
as a state
m
|θm,N i = e · vN ,
N +1
where vN is a superposition of occupation states
N
1 X
vN = |ni.
(N + 1)1/2 n=0
One needs then to ensure that the results are consistent over changes
of scale.
The other operators that generate the Bost–Connes algebra can be
thought of as implementing the changes of scales in the optical phases
in a consistent way. These operators are isometries µn parameterized
by positive integers n ∈ N× = Z>0 . The changes of scale are described
by the action of the µn on the e(r) by
1 X
(3.22) µn e(r)µ∗n = e(s).
n ns=r
In addition to this compatibility condition, the operators e(r) and µn
satisfy other simple relations.
These give a presentation of the algebra of the BC system of the
form ([8], [55]):
• µ∗n µn = 1, for all n ∈ N× ,
• µk µn = µkn , for all k, n ∈ N× ,
• e(0) = 1, e(r)∗ = e(−r), and e(r)e(s) = e(r + s), for all
r, s ∈ Q/Z,
• For all n ∈ N× and all r ∈ Q/Z, the relation (3.22) holds.
This shows that the Bost–Connes algebra is isomorphic to the algebra
A1 of (3.19).
In terms of this explicit presentation, the time evolution is of the
form
(3.23) σt (µn ) = nit µn , σt (e(r)) = e(r).
The space (3.16) can be compactified by replacing A· by A, as in
[22]. This gives the quotient
(3.24) Sh(nc) (GL1 , {±1}) = GL1 (Q)\A/R∗+ .
This compactification consists of adding the trivial lattice (with a pos-
sibly nontrivial Q-structure).
The dual space to (3.24), under the duality of type II and type
III factors introduced in Connes’ thesis, is a principal R∗+ -bundle over
(3.24), whose noncommutative algebra of coordinates is obtained from
62 3. QUANTUM STATISTICAL MECHANICS AND GALOIS THEORY
the algebra of (3.24) by taking the crossed product by the time evolu-
tion σt . The space obtained this way is the space of adèle classes
(3.25) L1 = GL1 (Q)\A → GL1 (Q)\A/R∗+
that gives the spectral realization of zeros of the Riemann ζ function
in [22]. Passing to this dual space corresponds to considering commen-
surability classes of 1-dimensional Q-lattices (no longer up to scaling).
2.1. Structure of KMS states. The Bost–Connes algebra A1
has irreducible representations on the Hilbert space H = `2 (N× ). These
are parameterized by elements α ∈ Ẑ∗ = GL1 (Ẑ). Any such element
defines an embedding α : Qcycl ,→ C and the corresponding represen-
tation is of the form
πα (e(r)) k = α(ζrk ) k
(3.26)
πα (µn ) k = nk
The Hamiltonian implementing the time evolution (3.18) on H is
(3.27) H k = log k k
Thus, the partition function of the Bost–Connes system is the Riemann
zeta function
∞
−βH
X
(3.28) Z(β) = Tr e = k −β = ζ(β).
k=1
Bost and Connes showed in [8] that KMS states have the following
structure, with a phase transition at β = 1.
• In the range β ≤ 1 there is a unique KMSβ state. Its restriction
to Q[Q/Z] is of the form
Y 1 − pβ−1
ϕβ (e(a/b)) = b−β .
1 − p−1
p prime, p|b
The tower of modular curves (cf. [31]) has base V = P1 over Q and
Vn = X(n) the modular curve corresponding to the principal congru-
ence subgroup Γ(n). The automorphisms of the projection Vn → V are
given by GL2 (Z/nZ)/{±1}, so that the group of deck transformations
of the tower V is in this case of the form
and Λ/nΛ = E[n] the nth torsion of E. Since the Q-lattices need not
be invertible, we do not require that η be an Af -isomorphism (cf. [78]).
The commensurability relation between Q-lattices corresponds to
the equivalence (E, η) ∼ (E 0 , η 0 ) given by an isogeny g : E → E 0 and
η 0 = (g ⊗ 1) ◦ η. Namely, the equivalence classes can be identified
with the quotient of M2 (Af ) × H± by the action of GL2 (Q), (ρ, z) 7→
(gρ, g(z)).
To associate a quantum statistical mechanical system to the space
of commensurability classes of 2-dimensional Q-lattices up to scaling,
it is convenient to use the description (3.35). Then one can consider
as noncommutative algebra of coordinates the convolution algebra of
continuous compactly supported functions on the quotient of the space
(3.38) U := {(g, ρ, z) ∈ GL+
2 (Q) × M2 (Ẑ) × H|gρ ∈ M2 (Ẑ)}
by the action of Γ × Γ,
(3.39) (g, ρ, z) 7→ (γ1 gγ2−1 , γ2 z).
One endows this algebra with the convolution product
X
(3.40) (f1 ∗ f2 )(g, ρ, z) = f1 (gs−1 , sρ, s(z))f2 (s, ρ, z)
s∈Γ\GL+
2 (Q):sρ∈M2 (Ẑ)
(3.42) Gρ := {g ∈ GL+
2 (Q) : gρ ∈ M2 (Ẑ)}
4.1. KMS states of the GL2 -system. The main result of [29]
on the structure of KMS states is the following.
Theorem 4.2. The KMSβ states of the GL2 -system have the fol-
lowing properties:
(1) In the range β ≤ 1 there are no KMS states.
(2) In the range β > 2 the set of extremal KMS states is given by
the classical Shimura variety
(3.48) Eβ ∼
= GL2 (Q)\GL2 (A)/C∗ .
This shows that the extremal KMS states at sufficiently low tem-
perature are parameterized by the invertible Q-lattices. The explicit
expression for these extremal KMSβ states is obtained as
1 X
(3.49) ϕβ,L (f ) = f (1, mρ, m(z)) det(m)−β
Z(β) +
m∈Γ\M2 (Z)
Here cycl : Ẑ∗ → Aut(F ) denotes the action of the Galois group Ẑ∗ '
Gal(Qab /Q) on the coefficients of the q-expansion in powers of q 1/N =
exp(2πiτ /N ).
If we took only the first two conditions, this would allow the algebra
A2,Q to contain the cyclotomic field Qab ⊂ C, but this would prevent
the existence of states satisfying the desired Galois property. In fact, if
the subalgebra contains scalar elements in Qcycl , the sought for Galois
property would not be compatible with the C-linearity of states. The
cyclotomic condition then forces the spectrum of the elements of A2,Q
to contain all Galois conjugates of any root of unity that appears in the
coefficients of the q-series, so that elements of A2,Q cannot be scalars.
The algebra A2,Q defined by the properties above is a subalgebra
of unbounded multipliers of A2 , which is globally invariant under the
group of symmetries Q∗ \GL2 (Af ).
5. Quadratic fields
The first essential step in the direction of generalizing the BC sys-
tem to other number fields and exploring a possible approach to the
explicit class field theory problem via noncommutative geometry is the
construction of a system that recovers√the explicit class field theory
for imaginary quadratic fields K = Q( −d), for d a positive integer.
This case is being investigated as part of ongoing joint work of Alain
Connes, Niranjan Ramachandran, and the author [31].
A system for imaginary quadratic fields can be constructed by con-
sidering 1-dimensional K-lattices. These are lattices in K with a homo-
morphism φ : K/O → QΛ/Λ, where O is the ring of integers of K. A
choice of a generator τ ∈ H such that K = Q(τ ) realizes 1-dimensional
K-lattices as a particular set of 2-dimensional Q-lattices. The com-
mensurability relation for 1-dimensional K-lattices can be defined as a
commensurability of the underlying 2-dimensional Q-lattices where the
isomorphism is realized by multiplication by an element in K∗ , viewed
as embedded in GL+ 2 (Q).
The set of commensurability classes of 1-dimensional K-lattices up
to scale is then described by the quotient
(3.61) AK,f /K∗ ,
where AK,f = Af ⊗K are the finite adèles of K, while the set of invertible
1-dimensional K-lattices up to scale can be identified with the idèle
class group
(3.62) CK /DK = A∗K,f /K∗ .
5. QUADRATIC FIELDS 73
forms” identified by Zagier, point to the fact that there should exist
a rich class of objects replacing modular forms, when “pushed to the
boundary”.
Regarding the role of modular forms, notice that, in the case of
(3.74), we can again consider the dual system. This is a C∗ -bundle
(3.75) L2 = GL2 (Q)\M2 (A).
On this dual space modular forms appear naturally instead of modular
functions and the algebra of coordinates contains the modular Hecke
algebra of Connes–Moscovici ([33], [34]) as arithmetic subalgebra.
Thus, the noncommutative geometry of the space of Q-lattices mod-
ulo the equivalence relation of commensurability provides a setting that
unifies several phenomena involving the interaction of noncommutative
geometry and number theory. These include the Bost–Connes system,
the noncommutative space underlying the construction of the spec-
tral realization of the zeros of the Riemann zeta function in [22], the
modular Hecke algebras of [33] [34], and the noncommutative modular
curves of [70].
CHAPTER 4
This chapter is based on joint work of Katia Consani and the au-
thor ([36], [37], [38], [39], [40]) that proposes a model for the dual
graph of the maximally degenerate fibers at the archimedean places
of an arithmetic surface in terms of a noncommutative space (spec-
tral triple) related to the action of a Schottky group on its limit set.
This description of ∞-adic geometry provides a compatible setting that
combines Manin’s result [66] on the Arakelov Green function for arith-
metic surfaces in terms of hyperbolic geometry and for a cohomologi-
cal construction of Consani [35] associated to the archimedean fibers
of arithmetic varieties, related to Deninger’s calculation of the local
L-factors as regularized determinants (cf. [43], [44]).
1. Schottky uniformization
Topologically a compact Riemann surface X of genus g is obtained
by gluing the sides of a 4g-gon. Correspondingly, the fundamental
group has a presentation
Y
π1 (X) = ha1 , . . . , ag , b1 , . . . , bg | [ai , bi ] = 1i,
i
is the set of doubly infinite words in the generators and their inverses,
with the admissibility condition that no cancellations occur. The map
T is the invertible shift
Then we can pass from the discrete dynamical system (S, T ) to its
suspension flow and obtain the mapping torus
(4.7) C(S) oT Z
S ×Z R = S T .
2. DYNAMICS AND NONCOMMUTATIVE GEOMETRY 89
K1 (C(S)) o K1 (C(S)) o K0 (A)
The H 1 (ST ) can be identified with the Čech cohomology group given
by the homotopy classes [ST , U (1)], by mapping f 7→ [exp(2πitf (x))],
for f ∈ C(S, Z)/δ.
over Spec(Z),
XZ ⊗Z Spec(Q) = X,
where the closed fiber of XZ at a prime p ∈ Spec(Z) is the reduction
X mod p. Thus, viewed as an arithmetic varities, an algebraic curve
becomes a 2-dimensional fibration over the affine line Spec(Z).
One can also consider reductions of X (defined over Z) modulo
n
p for some prime p ∈ Spec(Z). The limit as n → ∞ defines a p–
adic completion of XZ . This can be thought of as an “infinitesimal
neighborhood” of the fiber at p.
The picture is more complicated at arithmetic infinity, since one
does not have a suitable notion of “reduction mod ∞” available to
define the closed fiber. On the other hand, one does have the analog of
the p–adic completion at hand. This is given by the Riemann surface
(smooth projective algebraic curve over C) determined by the equation
of the algebraic curve X over Q, under the embedding of Q ⊂ C,
X(C) = X ⊗Q Spec(C)
with the absolute value | · | at the infinite prime replacing the p-adic
valuations.
Similarly, for K a number field and OK its ring of integers, a smooth
proper algebraic curve X over K determines a smooth minimal model
XOK , which defines an arithmetic surface XOK over Spec(OK ). The
closed fiber X℘ of XOK over a prime ℘ ∈ OK is given by the reduction
mod ℘.
When Spec(OK ) is compactified by adding the archimedean primes,
one also obtains n Riemann surfaces Xα (C), obtained from the equation
defining X over K under the embeddings α : K ,→ C. Of these Riemann
surfaces, r1 carry a real involution.
Thus, the picture of an arithmetic surface over Spec(OK ) is as fol-
lows:
/ XSpec(O ) / XSpec(O ) o
X℘ K K
? _ ???
℘ / Spec(OK ) / Spec(OK ) o ?_ α
where we do not have an explicit geometric description of the closed
fibers over the archimedean primes (Figure 7).
Formally, one can enlarge the group of divisors on the arithmetic
surface by adding formal real
P linear combinations of irreducible “closed
vertical fibers at infinity” α λα Fα . Here the fibers Fα are only treated
as formal symbols, and no geometric model of such fibers is provided.
3. ARITHMETIC INFINITY: ARCHIMEDEAN PRIMES 95
2 3 5 7 11 oo
where the first term counts the contribution from the finite places (i.e.
what happens over Spec(OK )) and the second term is the contribution
of the archimedean primes, i.e. the part of the intersection that happens
over arithmetic infinity.
96 4. NONCOMMUTATIVE GEOMETRY AT ARITHMETIC INFINITY
at points
α
{Pi,β | β = 1, . . . , [K(Di ) : K]} ⊂ Xα (C),
for a finite extension K(Di ) of K determined by Di . Here α = 1 for
real embeddings and = 2 for complex embeddings.
For a detailed account of these notions of Arakelov geometry, one
can refer to [41] [59].
Further evidence for the similarity between the archimedean and
the totally degenerate fibers came from an explicit computation of the
Green function at the archimedean places derived by Manin [66] in
terms of a Schottky uniformization of the Riemann surface Xα (C).
Such uniformization has an analog at a finite prime, in terms of p-adic
Schottky groups, only in the totally degenerate case. Another source
of evidence comes from a cohomological theory of the local factors at
the archimedean primes, developed by Consani [35], which shows that
the resulting description of the local factor as regularized determinant
at the archimedean primes resembles mostly the case of the totally
degenerate reduction at a finite prime.
We will present both results in the light of the noncommutative
space (OA , H, D) introduced in the previous section. As showed by
Consani and the author in [36] [37] [38] [39] the noncommutative
geometry of this space is naturally related to both Manin’s result on
the Arakelov Green function and the cohomological construction of
Consani.
b*{c,d}
a {c,d}
b
a*{c,d}
gives the lift to ΩΓ of a differential of the third kind on the Riemann sur-
face X(C), endowed with the choice of Schottky uniformization. These
differentials have residues ±1 at the images of a and b in X(C), and
they have vanishing ak periods, where {ak , bk }k=1...g are the generators
of the homology of X(C).
Similarly, we obtain lifts of differentials of the first kind on X(C),
by considering the series
X
(4.24) ωγ = d loghhz + (γ), hz − (γ), z, z0 i,
h∈C(|γ)
Thus, one obtains ([66], cf. also [93]) from (4.20) and (4.26) that
the Arakelov Green function for X(C) with Schottky uniformization
can be computed as
P
g((a) − (b), (c) − (d)) = h∈Γ log |ha, b, hc, hdi|
(4.28)
− g`=1 X` (a, b) h∈S(g` ) log |hz + (h), z − (h), c, di|.
P P
100 4. NONCOMMUTATIVE GEOMETRY AT ARITHMETIC INFINITY
3
H
X
q
-
1
-
x
x -
1 1 -
x
0
1
x
3 Xq
Η
δ0
-
0
-
xn
γ
0
-x
-x
0 1 - -x 1
1 0 2 x
n
q x
Here we are using the notation x̄ = x ∗ {0, ∞}; z̄n = q n z ∗ {1, ∞},
z̃n = q n z −1 ∗ {1, ∞} as in [66], as illustrated in Figure 10 and 11.
These terms describe gravitational properties of the Euclidean BTZ
black hole. For instance, `(γ0 ) measures the black hole entropy. The
whole expression is a combination of geodesic propagators.
. . . . .. .. . .
.. .
. .
. .
. .
. .
. .
.
.
.. . .
. .
. .
. .
. .
. .
. .
. .
. .
.
. . ..
. .
. .
. .
. .
. .
. . . .
. . . . .
Figure 12. Mumford curve of genus one: Jacobi–Tate
uniformized elliptic curve
XΓ := ΩΓ /Γ
a
b b
a b
a
a a
b b b b
a a
a
b b
a
b b
c
a a c
b
b
a c c a
c b b
c b c
a a a a
b a c b
b c a c
c b a b c b c a
a
a a a
b c b a c
a
a b c ab
c b c a
c b a
c
6.2. Model of the dual graph. The dictionary between the case
of Mumford curves and the case at arithmetic infinity is then summa-
rized as follows:
108 4. NONCOMMUTATIVE GEOMETRY AT ARITHMETIC INFINITY
equivalently as
Y m (X I℘
(4.39) L℘ (H m (X), s) = (1 − λq −s )− dim H Γ )λ
,
∗)
λ∈Spec(F r℘
I
where H m (XΓ )λ℘ is the eigenspace of the Frobenius with eigenvalue λ.
For our purposes, what is most important to retain from the dis-
cussion above is that the local L-factor (4.35) depends upon the data
H ∗ (X̄, Q` )I℘ , F r℘∗
(4.40)
of a vector space, which has a cohomological interpretation, together
with a linear operator.
7.1. Archimedean L-factors. Since étale cohomology satisfies
the compatibility (4.36), if we again resort to the general philosophy,
according to which we can work with the smooth complex manifold
X(C) and gain information on the “closed fiber” at airhtmetic infin-
ity, we are lead to expect that the contribution of the archimedean
primes to the L-function may be expressed in terms of the cohomology
H ∗ (X(C), C), or equivalently in terms of de Rham cohomology.
In fact, Serre showed ([87]) that the expected contribution of the
archimedean primes depends upon the Hodge structure
(4.41) H m (X(C)) = ⊕p+q=m H p,q (X(C))
and is again expressed in terms of Gamma functions, as in the case of
(4.33). Namely, one has a product of Gamma functions according to
the Hodge numbers hp,q ,
(4.42)
hp,q
( Q
∗ p,q ΓC (s − min(p, q))
L(H , s) = Q hp,q hp+ p−
ΓR (s − p + 1)h
Q
p<q ΓC (s − p) p ΓR (s − p)
where the two cases correspond, respectively, to the complex and the
real embeddings. Here hp,± is the dimension of the ±(−1)p -eigenspace
of the involution on H p,p induced by the real structure and
(4.43) ΓC (s) := (2π)−s Γ(s), ΓR (s) := 2−1/2 π −s/2 Γ(s/2).
One of the general ideas in arithmetic geometry is that one should
always seek a unified picture of what happens at the finite and at the
infinite primes. In particular, there should be a suitable reformula-
tion of the local factors (4.35) and (4.42) where both formulae can be
expressed in the same way.
Seeking a unified description of local L-factors at finite and infinite
primes, Deninger in [43], [44], [45] expressed both (4.35) and (4.42) as
infinite determinants.
112 4. NONCOMMUTATIVE GEOMETRY AT ARITHMETIC INFINITY
Vn ⊂ P n dim Vn = 2g
We’ll see in the next section that, in fact, the right hand side of (4.55)
is a particular case of a more general construction that works for arith-
metic varieties in any dimensions and that gives a cohomological inter-
pretation of (4.49).
8. ARCHIMEDEAN COHOMOLOGY 115
8. Archimedean cohomology
Consani gave in [35], for general arithmetic varieties (in any dimen-
sion), a cohomological interpretation of the pair (Hm , Φ) in Deninger’s
calculation of the archimedean L-factors as regularized determinants.
Her construction was motivated by the analogy between geometry
at arithmetic infinity and the classical geometry of a degeneration over
a disk. She introduced a double complex of differential forms with
an endomorphism N representing the “logarithm of the monodromy”
around the special fiber at arithmetic infinity, which is modelled on
(a resolution of) the complex of nearby cycles in the geometric case.
The definition of the complex of nearby cycles and of its resolution,
on which the following construction is modelled is rather technical.
What is easier to visualize geometrically is the related complex of the
vanishing cycles of a geometric degeneration (Figure 15).
We describe here the construction of [35] using the notation of [40].
We construct the cohomology theory underlying (4.49) in several steps.
In the following we let X = X(C) be a complex compact Kähler
manifold.
k
m, .
Tp,q
2r−2k+ |p−q|+m=0
r
2r+m−k=0
the Z-graded real vector space T · , since the inner product (4.58) is real
on real forms and induces an inner product on T · . We write δ = d0 + d00
for the total differential.
We can describe the real vector spaces T · in terms of certain cutoffs
on the indices of the complex C · . Namely, for
(4.66) Λp,q = {(r, k) ∈ Z2 : k ≥ κ(p, q, r)}
with
|p − q| + 2r + m
(4.67) κ(p, q, r) := max 0, 2r + m,
2
(cf. Figure 16), we identify T · as a real vector space with the span
(4.68) T · = Rhα ⊗ U r ⊗ ~k i,
¯ with ξ ∈ Ωp,q (X).
where (r, k) ∈ Λp,q for α = ξ + ξ,
the form
∂
(4.69) N = U~ Φ = −U
∂U
and they satisfy [N, d ] = [N, d ] = 0 and [Φ, d0 ] = [Φ, d00 ] = 0, hence
0 00
while the action of the subgroup ν(s), s ∈ R∗ of SL(2, R) via the rep-
resentation σ R on this Hilbert space is by unbounded densely defined
operators.
8.3. Renormalization group and monodromy. This very gen-
eral structure, which exists for varieties in any dimension, also has in-
teresting connections to noncommutative geometry. For instance, we
can see that in fact the map N does play the role of the “logarithm
of the monodromy” using an analog in our context of the theory of
renormalization à la Connes–Kreimer [28].
In the classical case of a geometric degeneration on a disk, the
monodromy around the special fiber is defined as the map
√
(4.75) T = exp(−2π −1 Res0 (∇))
where
(4.76) N = Res0 (∇)
is the residue at zero of the connection, acting as an endomorphism of
the cohomology.
In our setting, we consider loops φµ with values in the group G =
Aut(TC· , δ), depending on a “mass parameter” µ ∈ C∗ . Here TC· is the
complexification of the real vector space T · . The Birkhoff decomposi-
tion of a loop φµ consists in the multiplicative decomposition
(4.77) φµ (z) = φ−
µ (z)
−1 +
φµ (z),
for z ∈ ∂∆ ⊂ P1 (C), where ∆ is a small disk centered at zero. Of the
two terms in the right hand side of (4.77), φ+
µ extends to a holomorphic
function on ∆ and φµ to a holomorphic functions on P1 (C) r ∆ with
−
with
Z
(4.84) dk = θ−s1 (β) · · · θ−sk (β) ds1 · · · dsk .
s1 ≥···≥sk ≥0
for λ = et ∈ R∗+ .
Thus, we only need to specify the residue in order to have the
corresponding renormalization theory associated to (TC· , δ). By analogy
to the case of the geometric degeneration (4.76) it is natural to require
that Res φ = N . We then have ([40]):
Proposition 8.1. A loop φµ in G = Aut(TC· , δ) with Res φµ = N ,
subject to (4.79) and with φ− independent of µ, satisfies
z
µ
φµ (z) = exp N
z
with Birkhoff decomposition
µz − 1
φµ (z) = exp(−N/z) exp N .
z
In fact, by (4.83) and Res φµ = N we have β = [N, Φ] = N and
θt (N ) = et N , hence the scattering formula (4.84) gives
φ− (z) = exp(−N/z)
8. ARCHIMEDEAN COHOMOLOGY 121
where the KN are free abelian of rank (2g − 1)N + 1 for N even and
(2g − 1)N + (2g − 1) for N odd. The Z-module KN is generated by
the closed geodesics represented by periodic sequences in S of period
N + 1. These need not be primitive closed geodesics. In terms of
primitive closed geodesics we can write equivalently
H1 (ST , Z) = ⊕∞
N =0 RN
125
126 BIBLIOGRAPHY
[64] Yu.I. Manin, Theta functions, quantum tori, and Heisenberg groups, Lett.
in Math. Phys. 56 (2001) 295–320.
[65] Yu.I. Manin, Lectures of zeta functions and motives (according to Deninger
and Kurokawa). Columbia University Number Theory Seminar (New York,
1992). Astérisque No. 228 (1995), 4, 121–163.
[66] Yu.I. Manin, Three–dimensional hyperbolic geometry as ∞–adic Arakelov
geometry, Invent. Math. 104 (1991) 223–244.
[67] Yu.I. Manin, p-adic automorphic functions. Journ. of Soviet Math., 5 (1976)
279-333.
[68] Yu.I. Manin, Parabolic points and zeta functions of modular curves, Math.
USSR Izvestija, vol. 6 N. 1 (1972) 19–64. Selected Papers, World Scientific,
1996, 202–247.
[69] Yu.I. Manin, M. Marcolli, Holography principle and arithmetic of algebraic
curves, Adv. Theor. Math. Phys. Vol.3 (2001) N.5, 617–650.
[70] Yu.I. Manin, M. Marcolli, Continued fractions, modular symbols, and non-
commutative geometry, Selecta Mathematica (New Series) Vol.8 N.3 (2002)
475-520.
[71] M. Marcolli, Limiting modular symbols and the Lyapunov spectrum, Journal
of Number Theory, Vol.98 N.2 (2003) 348-376.
[72] M. Marcolli, Modular curves, C ∗ -algebras and chaotic cosmology, preprint.
[73] K. Matsuzaki, M. Taniguchi, Hyperbolic manifolds and Kleinian groups, Ox-
ford Univ. Press, 1998.
[74] D.H. Mayer, Relaxation properties of the mixmaster universe, Phys. Lett. A
121 (1987), no. 8-9, 390–394.
[75] D.H. Mayer, Continued fractions and related transformations, 1991.
[76] B. Mazur, Courbes elliptiques et symboles modulaires, Séminaire Bourbaki
vol.1971/72, Exposés 400–417, LNM 317, Springer Verlag 1973, pp. 277–294.
[77] L. Merel, Intersections sur les courbes modulaires, Manuscripta Math., 80
(1993) 283–289.
[78] J.S. Milne, Canonical models of Shimura curves, manuscript, 2003
(www.jmilne.org)
[79] D. Mumford (with M.Nori and P.Norman), Tata lectures on theta, III,
Progress in Mathematics Vol. 97, Birkhäuser 1991.
[80] D. Mumford, An analytic construction of degenerating curves over complete
local rings, Compositio Math. 24 (1972) 129–174.
[81] D. Mumford, C. Series, D. Wright, Indra’s Pearls. The Vision of Felix Klein.
Cambridge University Press, New York, 2002.
[82] W. Parry, S. Tuncel, Classification problems in ergodic theory, London Math.
Soc. Lecture Notes Series 67, 1982.
[83] M. Pollicott, H. Weiss, Multifractal analysis of Lyapunov exponent for con-
tinued fraction and Manneville-Pomeau transformations and applications to
Diophantine approximation, Comm. Math. Phys. 207 (1999), no. 1, 145–171.
[84] D.B. Ray, I.M. Singer, Analytic torsion for complex manifolds. Ann. of Math.
(2) 98 (1973), 154–177.
[85] M.A. Rieffel, C∗ -algebras associated to irrational rotations, Pacific J. Math.
93 (1981) 415-429.
[86] A.L. Rosenberg, Noncommutative schemes. Compositio Math. 112 (1998)
N.1, 93–125.
BIBLIOGRAPHY 129
[87] J. P. Serre, Facteurs locaux des fonctions zêta des variétés algébriques
(définitions et conjectures). Sém. Delange-Pisot-Poitou, exp. 19, 1969/70.
[88] G. Shimura, Arithmetic Theory of Automorphic Functions, Iwanami Shoten
and Princeton 1971.
[89] H.M. Stark, L-functions at s = 1. IV. First derivatives at s = 0, Adv. Math.
35 (1980) 197–235.
[90] P. Stevenhagen, Hilbert’s 12th problem, complex multiplication and Shimura
reciprocity, Advanced Studies in Pure Math. 30 (2001) “Class Field Theory
– its centenary and prospect” pp. 161–176.
[91] A. Weil, Basic Number Theory, Springer 1974.
[92] R.O. Wells, Differential analysis on complex manifolds. Springer–Verlag,
1980.
[93] A. Werner, Local heights on Mumford curves. Math. Ann. 306 (1996), no. 4,
819–831.
[94] D. Zagier, Modular forms and differential operators, in K. G. Ramanathan
memorial issue. Proc. Indian Acad. Sci. Math. Sci. 104 (1994), no. 1, 57–75.