Continued Fractions, Discrete Groups and Complex Dynamics
Continued Fractions, Discrete Groups and Complex Dynamics
Continued Fractions, Discrete Groups and Complex Dynamics
Alan F. Beardon
Contents
1. Introduction 536
2. Continued fractions 540
3. Complex Möbius maps 542
4. The stability of continued fractions 553
5. Möbius maps and hyperbolic geometry 559
6. General convergence 564
7. Strong divergence 568
8. Möbius maps in higher dimensions 573
9. Backward orbits 581
10. Forward orbits 585
11. Self-maps of a domain 586
12. Concluding remarks 590
References 591
Received April 29, 2002.
c 2001 Heldermann Verlag
ISSN 1617-9447/$ 2.50 °
536 A. F. Beardon CMFT
1. Introduction
This paper is about the connections between continued fractions, Möbius maps,
discrete groups, complex dynamics, hyperbolic and inversive geometries.
Traditionally, an infinite continued fraction is considered to be an expression of
the form
a1
(1.1) K(an |bn ) = ,
a2
b1 +
a3
b2 +
b3 + · · ·
where the ai and bj are infinite sequences of complex numbers with each ai 6= 0.
The continued fraction (1.1) is said to converge to, or to have value, K if the
sequence of truncated continued fractions
a1 a1 a1
(1.2) , , ,...
b1 a2 a2
b1 + b1 +
b2 a3
b2 +
b3
converges to K. More generally, one can consider continued fractions a0 +
K(an |bn ), but there is no harm in taking a0 = 0 as we shall do throughout
this paper.
From an abstract point of view, the continued fraction (1.1) is simply the pair of
complex sequences an and bn in which no ai is zero. Given two such sequences
we can form the sequence sn of Möbius maps defined by
an
(1.3) sn (z) = ,
bn + z
and sn (∞) = 0. Conversely, as any Möbius map s with s(∞) = 0 can be
expressed uniquely in the form s(z) = a/(b + z), we can identify the class of con-
tinued fractions with the class of sequences sn of Möbius maps with sn (∞) = 0
for all n. With this identification the truncated continued fractions in (1.2) are
s1 (0), s1 s2 (0), . . ., where, in general, f g denotes the map z 7→ f (g(z)), so that
the convergence of the continued fraction is equivalent to the convergence of
the complex sequence s1 · · · sn (0). As Möbius maps exist in all dimensions, the
identification of continued fractions with sequences of Möbius maps immediately
provides us with the concept of continued fractions in all dimensions, namely
sequences sn of Möbius maps with sn (∞) = 0. In higher dimensions the op-
eration of division is essentially replaced by inversion in the unit sphere, and
the arguments become geometric rather than algebraic. In this sense, continued
fraction theory may be regarded as a part of inversive geometry in n dimen-
sions. As Möbius maps acting in a higher dimension satisfy most of the familiar
properties enjoyed by complex Möbius maps, we might hope that many results
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 537
the reader to [4], [5], [12], [23], [38], [62], [64] and [70]. Finally, for the general
theory of complex iteration, we recommend [14], [25] and [74].
2. Continued fractions
Continued fractions have been studied since at least 1572 when R. Bombelli,
and
√ later P.√ Cataldi, used finite continued fractions to obtain approximations to
13 and 18, respectively. The first infinite continued fraction expansion (of
4/π) was given by Lord Brouckner around 1659, who also gave (in 1657) the first
systematic method for solving Pell’s equation by using continued fractions (see
[30, p. 108]). In 1695 John Wallis published Opera Mathematica, a text which
contained the first systematic study of the general theory of continued fractions
and in which the term continued fraction was first introduced. Euler gave a sys-
tematic exposition in De Fractionlous Continious (1737), and Jacobi, Hermite,
Gauss, Cauchy, Stieltjes, Ramanujan, and many others have all contributed to
the subject.
We have already seen that the continued fraction (1.1) can be viewed as the
sequence of Möbius maps sn in (1.3). We shall often refer to the continued
fraction (1.1) as the continued fraction generated by the sn , or simply as the
continued fraction sn . We shall also use the notation [s1 , s2 , . . .] for this continued
fraction; thus if sn (z) = an /(bn + z), then K(an |bn ) and [s1 , s2 , . . .] are different
notations for the same continued fraction. In particular, [s, s, . . .] denotes a
periodic continued fraction (of period one). We write Sn = s1 · · · sn , and we also
say that the sj generate the sequence Sj . We shall reserve the notation an , bn , sn
and Sn as used here throughout this paper.
In general, given a sequence of maps fj of a set into itself, we can construct the
sequences Fn = f1 · · · fn and Gn = fn · · · f1 . We say that Fn is the inner compo-
sition sequence, and that Gn is the outer composition sequence, of the fn . For a
continued fraction, Sn is the inner composition sequence of its generators sn , and
many convergence theorems on continued fractions are special cases of general
results on the convergence of inner composition sequences of analytic maps. We
shall not consider outer composition sequences in this paper; usually, these are
easier to handle than inner composition sequences.
The sequence of truncated continued fraction in (1.2) is the sequence Sn (0);
thus the continued fraction (1.1) converges if and only if the sequence of maps
S1 , S2 , . . . converges at 0. The convergence of a continued fraction is unaffected by
the addition or removal of a finite number of the generators sj for s1 · · · sn (0) → α
if and only if s2 · · · sn (0) → s−1 1 (α). It is easy to see that a given sequence
z1 , z2 , . . . in C∞ is the sequence Sn (0) for some continued fraction (1.1) if and
only if z1 6= 0 and zn 6= zn+1 (see, for example, [51, p. 34] or [55, p. 71]). This
implies that virtually every complex sequence zn is the sequence of approximants
Sn (0), and as these approximants may ‘wander’ over the complex plane in an
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 541
arbitrary manner, we can certainly see the need for conditions that guarantee
convergence. Despite the historical precedent, it makes no sense to to study
the convergence of a sequence of maps at just one point, and we must concern
ourselves with the possible convergence (pointwise, and uniform) of the Sn on
subsets of C∞ .
Originally, continued fractions were considered in terms of recurrence relations.
Given the continued fraction (1.1), we define sequences An and Bn by
µ ¶ µ ¶
A−1 A0 1 0
= ,
B−1 B0 0 1
µ ¶ µ ¶µ ¶
An+1 An+2 An An+1 0 an+2
= ,
Bn+1 Bn+2 Bn Bn+1 1 bn+2
where n ≥ −1, or, equivalently, by the recurrence relations
An = bn An−1 + an An−2 ,
(2.1)
Bn = bn Bn−1 + an Bn−2 .
As
µ ¶ µ ¶ µ ¶
An−1 An 0 a1 0 an
(2.2) = ···
Bn−1 Bn 1 b1 1 bn
we can take determinants of both sides and obtain
µ ¶
An−1 An
(2.3) det = (−1)n a1 · · · an 6= 0.
Bn−1 Bn
Now consider the homomorphism
µ ¶
a b az + b
(2.4) Φ : A 7→ gA , A= , gA (z) = ,
c d cz + d
from the group GL(2, C) of 2 × 2 non-singular matrices onto the group M of
Möbius maps (this is the first of several occasions on which the reader will be
urged, with good reason, to distinguish clearly between Möbius maps and ma-
trices). Applying Φ to both sides of (2.2) we see that
An−1 z + An
(2.5) Sn (z) = .
Bn−1 z + Bn
It follows that the sequence of truncated continued fractions in (1.2) or, equiva-
lently, the sequence of approximants Sn (0), is An /Bn . The recurrence relations
(2.1) will be very familiar to those readers who have studied continued fractions
before. In fact, we shall not use these at all in this paper.
Many results on continued fractions are proved by considering the behaviour of
solutions of the recurrence relations (2.1). In general, one cannot solve these
equations; thus progress depends on obtaining estimates of solutions, and there
are very many papers in which this is the central issue. The following classical
theorem is proved in this way, and here the required estimates are easily obtained.
542 A. F. Beardon CMFT
P
Theorem (The Stern-Stolz Theorem). If n |bn | converges then K(1|bn ) di-
verges.
The motivation for this result is clear, for if s(z) = 1/z, then the continued
fraction [s, s, . . .] has approximants ∞, 0, ∞, 0, . . . and so diverges. In the Stern-
Stolz Theorem, sn (z) = 1/(bn + z), and bn → 0, so that sn → s, and we might
expect that [s1 , s2 , . . .] diverges, at least if bn → 0 sufficiently quickly.
This brief discussion raises the question of what is meant by the convergence
gn → g of Möbius maps, and how one should measure the rate of this convergence.
The classical way to measure the rate at which sn → s, where sn (z) = an /(bn +z)
and s(z) = a/(b + z), is in terms of the quantities |an − a| and |bn − b|. This is
equivalent to using the norm k(z, w)k = |z|+|w| on C×C, and it is not at all clear
that this is the best measure to use. Indeed, and as we shall see later, there are
other norms that are much more closely tied to the geometric action of Möbius
maps and which generally provide much shorter and more elegant proofs of this
type of result. For example, the Möbius group M carries a natural metric σ0
with respect to which it is a topological group, and it is surely more natural to
measure the rate at which sn → s directly in terms of σ0 . Note that once again,
we prefer to distinguish clearly between Möbius maps and matrices. In any
event, if we are to measure the rate in terms of the coefficients then, because the
matrix norm is geometrically significant (whereas the individual terms |a n − a|
and |bn − b| are not), it must surely be better to pass from Möbius maps to
matrices and then use the usual matrix norm kAk on the vector space of 2 × 2
matrices. To add support to this view, we end this section with the remark that
we shall see that if sn (z) = 1/(bn + z) (as in the Stern-Stolz Theorem), then |bn |
is essentially the distance that sn moves the distinguished point in our model of
hyperbolic 3-space. The proof of the Stern-Stolz Theorem is then nothing more
than an application of the triangle inequality between hyperbolic distances, and
this proof is valid in all dimensions, and to other continued fractions. From the
matrix point of view, sn is repesented by the unimodular matrix
µ ¶
0 i
An = ,
i ibn
and the crucial fact is that kAn k is close to kIk, where I is the 2 × 2 identity
matrix. In fact, we shall see that an analogue of the Stern-Stolz Theorem is
true in all dimensions for any sequence sn of Möbius maps sn , providing only
that their matrix norms approach kIk sufficiently rapidly. It is not necessary for
the sn themselves to converge.
|az + b|2 + |cz + d|2 ≤ (|a|2 + |b|2 + |c|2 + |d|2 )(|z|2 + 1).
(when ad − bc = 1) so that
¡ ¢
σ g(z), g(w) 1
≥ .
σ(z, w) kgk2
544 A. F. Beardon CMFT
We now write u = g(z), v = g(w); then, noting that kgk = kg −1 k (as ad−bc = 1),
we obtain
¡ ¢
σ g −1 (u), g −1 (v) σ(z, w)
= ¡ ¢ ≤ kgk2 = kg −1 k2 .
σ(u, v) σ g(z), g(w)
As this holds for all u and v, and all Möbius g −1 , we see (after changing the
variables again) that for all z, w and g,
¡ ¢
σ g(z), g(w)
≤ kgk2 .
σ(z, w)
The best Lipschitz constant of a Möbius map g is L(g), where
¡ ¢
σ g(z), g(w)
L(g) = sup ;
z6=w σ(z, w)
thus
(3.2) L(g) ≤ kgk2 = |a|2 + |b|2 + |c|2 + |d|2 .
In Section 5 we shall give an explicit formula for L(g) in terms of a, b, c and d
which is slightly better than (3.2). However, (3.2) is asymptotically correct as
kgk → +∞.
The chordal metric σ enables us to introduce a metric σ0 on the group M of
complex Möbius maps, namely the metric of uniform convergence on C∞ . This
metric is defined by
¡ ¢
(3.3) σ0 (f, g) = sup σ f (z), g(z) ,
z∈C∞
and by gn → g we shall always mean σ0 (gn , g) → 0; that is, the uniform con-
vergence of gn to g on C∞ . We note that σ0 satisfies the two elementary, but
extremely useful, properties
(3.4) σ0 (f h, gh) = σ0 (f, g), σ0 (hf, hg) ≤ L(h) σ0 (f, g).
The first of these says that σ0 is a right-invariant metric. The second property
is a direct consequence of the inequalities
¡ ¢ ¡ ¢
σ hf (z), hg(z) ≤ L(h) σ f (z), g(z) ≤ L(h) σ0 (f, g)
that hold for all z. We recall the following basic facts about the metric space
(M, σ0 ).
Theorem 3.1. The space (M, σ0 ) is a complete metric space, and also a topo-
logical group.
To say that (M, σ0 ) is a topological group means that the maps f 7→ f −1 and
(f, g) 7→ f g are continuous maps of M and M × M into M, respectively, and
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 545
this is easily proved. First, as we are considering a metric space, we can use
sequences to establish continuity. For any f and g we have
σ0 (f −1 , g −1 ) = σ0 (f −1 g, I) = σ0 (f −1 g, f −1 f ) ≤ L(f −1 ) σ0 (g, f ),
The importance of these results is that they provide a link between neighbour-
hoods of the identities in the groups M and SL(2, C). Theorem 3.3 occurs in
[12, p. 79], but, by using a more detailed argument, Gehring and Martin ([40,
p. 48]) have shown that
σ0 (g, I) ≤ kG − G−1 k ≤ 2kG − Ik,
the second inequality holding because for any A in SL(2, C),
kA − A−1 k ≤ kA − Ik + kI − A−1 k = 2kA − Ik.
Theorem 3.4 is [39, Theorem 4.1].
Now in any topological group, we can transfer results about a neighbourhood
of the identity to results about a neighbourhood of any other point; thus we
can extend Theorems 3.3 and 3.4 as follows. First, if F and G are unimodular
matrices, and f = Φ(F ) and g = Φ(G), then
σ0 (f, g) = σ0 (I, gf −1 ) ≤ 2kI − GF −1 k
= 2k(F − G)F −1 k ≤ 2kF − Gk kF −1 k
= 2kF − Gk kF k.
Moreover, as we can interchange F and G in this last inequality, we have proved
that if F and G are unimodular matrices, and f = Φ(F ) and g = Φ(G), then
¡ ¢
(3.5) σ0 f, g ≤ 2 min{kF k, kGk} kF − Gk.
This prompts the question: what is the smallest constant µ such that for all
unimodular matrices F and G,
(3.6) σ0 (f, g) ≤ µ min{kF k, kGk} kF − Gk ?
We have just seen that µ ≤ 2, and we shall now show that µ ≥ 1.
Let ε be any positive number, and let
1
1 1ε
µ ¶ µ ¶
1+ε ε
F = , G= 1 .
0 1 0 1+ε
Sketch of the proof of Theorem 3.2. First, (3.5) shows √ that Φ is continu-
ous. The fact that Φ is injective on any open ball of radius 2 follows directly
from the Cauchy-Schwarz inequality (see [12, p. 78]). It remains to prove that Φ
is an open map. Let U be any non-empty open subset of SL(2, C), take any
matrix F in U and let f = Φ(F ). We want to show that f is an interior point of
Φ(U ). As U is an open neighbourhood of F , there is some positive δ such that
{X ∈ SL(2, C) : kX − F k < δ} ⊂ U . It is clear from (3.7) that there is some
positive η such that if σ0 (g, f ) < η, then g can be represented by a unimodular
matrix G such that kG−F k < δ. Thus Φ(U ) contains the open ball with centre f
and radius η; hence Φ is an open map.
The key step in the proof of Theorems 3.3 and 3.4 is that the unitary matrices act
(via Möbius transformations) as isometries with respect to the chordal metric,
and also (by conjugation) as isometries with respect to the metric kA − Bk. Let
us expand on this statement. A 2 × 2 complex matrix A is unitary if and only if
AĀt = I or, equivalently, if and only if kAk2 = 2 (see [12, p. 17]), and we have
the following result.
Theorem 3.5. Suppose that A is a 2 × 2 unitary matrix. Then the conjuga-
tion X 7→ AXA−1 is an isometry of the vector space of 2 × 2 complex matrices
with respect to the metric kX − Y k. Now let f be the Möbius transformation
Φ(A); then f is a chordal isometry. In particular, for all Möbius g and h,
σ0 (f gf −1 , f hf −1 ) = σ0 (g, h), so that the conjugation g 7→ f gf −1 is an isome-
try of (M, σ0 ). Finally, every chordal isometry is of the form Φ(A) for some
unitary A.
For example, the map z 7→ 1/z is represented by the unimodular matrix A with
entries 0, i, i, 0, respectively, and this is unitary as kAk2 = 2. We remark that
the crucial hypothesis in the Stern-Stolz Theorem is not that the maps sn are of
the form an /(bn + z), nor that they are close to the map z 7→ 1/z; the conclusion
depends on the fact that the sn are Möbius maps (without restriction) that are
close to some unitary map.
This is an appropriate point to mention the following familiar result that is
closely related to Theorem 3.4.
Theorem 3.6. If a sequence gn of Möbius maps converges at three distinct points
of C∞ to three distinct values, then the sequence gn converges uniformly on C∞
to some Möbius map g.
This result is not as transparent as it might seem at first sight for it is, in fact,
the combination of two separate results. The assumption that gn converges at
three distinct points is sufficient to guarantee in all dimensions that the family
{g1 , g2 , . . .} is a normal (equicontinuous) family. Now any limit function arising
from any such family is necessarily Möbius, but the fact that the entire sequence
converges uniformly to some Möbius map is only true in dimension two because it
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 549
is only here that the values at three points are sufficient to completely determine
the Möbius map. We shall consider the higher dimensional case later. A proof
of Theorem 3.6 is given in, for example, [31], but the following proof (which
illustrates the use of the metric σ0 ) is much neater.
In [31] DePree and Thron proved various results about the convergence of Möbius
maps and matrices; in particular, they obtained the following theorems (which
are expressed here in our notation).
P
Theorem 3.7. Let sn be a sequence of Möbius maps. If n σ0 (sn , I) converges,
then s1 · · · sn converges uniformly on C∞ to some Möbius map.
Theorem 3.8. Let An be matrices in SL(2, C) with kAn − Ik ≤ k(−An ) − Ik,
and let fn be the Möbius map corresponding to An . Then the three series
X X X
(3.8) σ0 (fn , I), kAn − Ik, kPn − Pn−1 k
n n n
Again we illustrate the value of the metric σ0 by giving proofs that are much
shorter than those in [31].
Proof of Theorem 3.7. As σ0 (sn , I) = σ0 (I, s−1 n ), our assumption implies that
−1
P
n σ0 (sn , I) converges. The right-invariance of σ0 now implies that
X
σ0 (s−1 −1 −1 −1
n · · · s1 , sn−1 · · · s1 )
n
550 A. F. Beardon CMFT
converges, and hence that (s1 · · · sn )−1 is a Cauchy sequence. Thus this sequence,
and hence the sequence of its inverses, namely s1 · · · sn , converges in M. This
proof is valid in all dimensions.
Proof of Theorem 3.8. Theorems 3.3 and 3.4 implyPthat the first two series in
(3.8) converge or diverge together. Suppose now that n kPn − Pn−1 k converges.
Then Pn is a Cauchy sequence and hence bounded (as SL(2, C) is complete). As
−1
kAn − Ik = kPn−1 (Pn − Pn−1 )k ≤ kPn−1 k kPn − Pn−1 k,
P
we see that n kAn − Ik converges.
P P
Finally, suppose that n kAn − Ik converges. Then n σ0 (fn , I) converges. Let
pn = f1 · · · fn . Then, by right invariance,
σ0 (fn , I) = σ0 (fn−1 , I) = σ0 (p−1 −1
n , pn−1 );
−1
thus n σ0 (p−1
P
n , pn−1 ) converges. As (M, σ0 ) is complete, the Cauchy sequence
p−1
n converges, and as M is a topological group, pn → q, say, for some Möbius q.
Let Q be a unimodular matrix representing q; then there is a sequence εn , with
εn = ±1, such that εn Pn → Q. We deduce that kPn k ≤ M , say, and as
kPn − Pn−1 k = kPn−1 (An − I)k ≤ kPn−1 k kAn − Ik,
P
we see that n kPn − Pn−1 k converges.
One of the most important techniques in studying Möbius maps is that of taking
conjugates. Conjugation by g (that is, the map f 7→ gf g −1 ) is a homeomor-
phism of M onto itself, and most of the important properties of Möbius maps
are invariant under conjugation. For example, in many proofs it is sufficient to
replace a given Möbius map by a (much simpler) conjugate map and then com-
plete the proof for this map; specifically, it is often sufficient to only consider the
maps z 7→ z + 1 and z 7→ kz. This technique is used constantly in the literature
on discrete groups; it is acknowledged, but rarely exploited, in the literature on
continued fractions. Consider, for example, the classification of Möbius maps.
It seems customary in continued fraction theory to classify Möbius maps by the
position of their fixed points. For example, one reads that if s(z) = a/(b + z),
then s has two fixed points, say z1 and z2 , and s is parabolic if z1 = z2 , elliptic
if z1 6= z2 but |z1 | = |z2 |, and loxodromic if |z1 | 6= |z2 |. We remark that this
classification is only valid for maps of the form a/(b + z) and, more importantly,
it is not invariant under conjugation. By contrast, the classification in terms of
the trace function is applicable to all Möbius maps, it is invariant under conju-
gation, and it is simpler in that it does not require one to find the fixed points.
If we are to extend results on continued fractions to general Möbius maps then
we must work with general tools, and there is ample evidence of the importance
of the trace function from discrete group theory. In fact, this function actually
parametrizes the conjugacy classes, and its importance extends far beyond the
classification of Möbius maps (see, for example, [81]).
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 551
Obviously, M is the disjoint union of {I} and the three classes of parabolic,
elliptic and loxodromic maps. The following well known result illustrates the
importance of the trace function.
Theorem 3.10. Let f and g be two Möbius maps neither of which is the identity.
Then f and g are conjugate in M if and only if trace2 (f ) = trace2 (g).
The class of loxodromic maps divides into two subclasses: a loxodromic map g (as
above) is hyperbolic if and only if (a + d)2 ∈ (4, +∞), and it is strictly loxodromic
if and only if (a + d)2 ∈ / [0, +∞). The set of hyperbolic maps is an important
subclass of the set of loxodromic maps because a loxodromic map g is hyperbolic
if and only if it has an invariant disc D (that is, such that g(D) = D), and also
if and only if it can be written as the composition of exactly two inversions (most
Möbius maps require four inversions). Now maps of the form s(z) = 1/(b + z),
where b 6= 0, are of interest in number theory and continued fraction theory. As
(trace)2 (s) = −b2 we see that if b is real then s is loxodromic but not hyperbolic,
so that s has no invariant disc in C∞ . However, as we shall see later, s does
have an invariant disc in hyperbolic 3-space, and that this disc is a model of the
hyperbolic plane that happens to lie in this higher dimensional space rather than
in C∞ .
Certain questions about the convergence and stability of Möbius maps can be
settled for loxodromic maps, but not for elliptic and parabolic maps, and the
reason for this is that the set L of loxodromic maps is an open subset of M. If
we make a sufficiently small perturbation of a loxodromic map the resulting map
is loxodromic; a sufficiently small perturbation of an elliptic or parabolic map
can result in an elliptic or a loxodromic map. To see that L is an open subset
of M we consider Φ : SL(2, C) → M¡ and the¢ map τ : SL(2, C) → C given by
τ (A) = [trace(A)]2 . Then L = Φτ −1 C\[0, 4] , and as τ is continuous, and Φ
is an open map, L is an open subset of M. Note also that (in some sense) the
generic Möbius map is loxodromic.
It is easy to characterize the classification given above in terms of the geometric
action of the maps, and the following is standard. A Möbius map g is
(1) parabolic if it is conjugate to z 7→ z + 1;
552 A. F. Beardon CMFT
Of course, this result is well known, but it often stated in a less concise way
(in terms of the fixed points of s). Moreover, its history provides evidence of
a lack of communication between different fields within this general area. Stolz
is sometimes given credit for first analyzing periodic continued fractions; see,
for example, [51, Chapter 3] where some of the history of this topic is given.
Also, see [53], where a reference is made to a proof by Stolz in 1886, an im-
proved proof by Pringsheim in 1900, and a shorter proof by Perron in 1905 “in
which the ideas related to the transformation point of view were virtually su-
pressed”. It is interesting to compare this with the state of discrete groups at
this time. In 1880 Poincaré submitted an eighty page paper on Fuchsian groups
and hyperbolic geometry to the French Academy, Klein’s Erlangen Program was
announced in 1872, and by 1887 Schottky was already studying the geometry
of free products of cyclic groups of Möbius maps. These works may not have
contained specific references to continued fractions, but they went far beyond the
study of cyclic groups and it is certain that the essential content of the results
about periodic continued fractions were very well known long before 1886. Even
more surprisingly, credit for a description of periodic continued fractions purely
in terms of Möbius maps is sometimes given to Schwerdtfeger ([72]) as late as
1946. In the light of the interest in complex dynamics and Julia sets in recent
years, we should perhaps note that [72] contains an incorrect definition of an
attracting fixed point, and a criticism of Julia’s definition of an indifferent fixed
point.
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 553
We end this section by analyzing Möbius maps in terms of inversions across circles
and lines in the plane. This can be done in a canonical way using isometric circles
which were introduced by L. R. Ford in 1927. These give a particularly simple
decomposition of a Möbius map g into inversions, and use only the smallest
number of inversions that are needed (two or four). Consider the Euclidean
distortion of the map
az + b
g(z) = , ad − bc = 1, c 6= 0.
cz + d
The distortion at z is |g 0 (z)| or, equivalently, |cz + d|−2 . According to Ford, the
isometric circle of g is the set of points at which this Euclidean distortion is 1:
thus the isometric circle of g is given by |cz + d| = 1 (providing that ad − bc = 1)
and this is indeed a circle. We denote this circle by I(g); its centre is g −1 (∞)
and its radius is 1/|c|. What is the image of I(g) under g? It must be a circle,
say C, and as the restriction of g to I(g) is a Euclidean isometry, the restriction
of g −1 to C must also be a Euclidean isometry. We conclude that C is the circle
I(g −1 ); this has centre g(∞) and the same radius as I(g). In short, g maps I(g)
given by |cz + d| = 1 onto I(g −1 ) which is given by |cz − a| = 1.
Now let α denote inversion in I(g), and let β denote reflection across the Eu-
clidean bisector of the points g(∞) and g −1 (∞) (henceforth, we shall use the
words ‘reflection’ and ‘inversion’ interchangeably). Then both g and βα are
Möbius maps that map I(g) onto I(g −1 ) and ∞ onto the centre of I(g −1 ). It
follows that g(βα)−1 leaves I(g −1 ) invariant, and as it fixes ∞, it also fixes the
centre of I(g −1 ) (the inverse point to ∞). Thus g(βα)−1 is some rotation γ
about g(∞). This shows that g = γβα, and hence that every Möbius map can be
expressed as the composition of at most four inversions (namely α, β, and two
reflections for γ). If the rotation γ is trivial then only two inversions are needed.
The classification of a Möbius map g is partly (but not entirely) reflected in
the relative position of the isometric circles I(g) and I(g −1 ). If g is parabolic,
then these two circles are tangent; if g is elliptic, these two circles cross; if g is
hyperbolic, these two circles are exterior to each other. In each of these cases, g
is the inversion in I(g) followed by the reflection in the Euclidean bisector of the
points g −1 (∞) and g(∞) (and the rotation γ is trivial). If g is strictly loxodromic
then there is no restriction on the location of the circles I(g) and I(g −1 ), and g is
the composition of the two inversions described above followed by a non-trivial
rotation about the point g(∞) by an angle −2 arg(a + d) (see [38, p. 28]).
case of a result about analytic functions. Despite this increased generality, our
arguments are simpler (and shorter) than those given for continued fractions,
and they enable us to describe the stability in terms of explicit inequalities. We
begin with the following result ([56, Theorem 2.2]).
Theorem 4.1. Given a continued fraction K(an |bn ), there exist positive numbers
r1 , r2 , . . . and s1 , s2 , . . . such that if |αn − an | < rn and |βn − bn | < sn for all n,
then the continued fractions K(αn |βn ) and K(an |bn ) are both convergent or both
divergent.
We shall show that this is a special case of the following result on topological
groups, and this generalization implies that the analogous result to Theorem 4.1
holds for any inner composition sequence of Möbius maps in any dimension.
Theorem 4.2. Suppose that G is a topological group whose topology is derived
from a right-invariant metric σ0 , and that (G, σ0 ) is complete. Let f1 , f2 , . . . be
any sequence of elements of G. Then, for each k, there is a neighbourhood Nk of
fk such that if, for all j, gj ∈ Nj , then (g1 · · · gn )(f1 · · · fn )−1 converges to some
element h of G.
Proof. We write Fn = f1 · · · fn , Gn = g1 · · · gn , and hn = Gn Fn−1 . For any k,
σ0 (h−1 −1 −1 −1
k , hk+1 ) = σ0 (hk Gk+1 , hk+1 Gk+1 ) = σ0 (Fk gk+1 , Fk fk+1 ),
so that if m < n,
n−1
X n−1
X
σ0 (h−1 −1
m , hn ) ≤ σ0 (h−1 −1
k , hk+1 ) = σ0 (Fk gk+1 , Fk fk+1 ).
k=m k=m
Now let θk : G → G be the map defined as left multiplication by Fk . Then θk
is continuous, so that if ∆k+1 denotes the open ball with centre θk (fk+1 ) and
radius 1/2k+1 , then ¡there is an open neighbourhood, say Nk+1 , of fk+1 such that
if g ∈ Nk+1 then σ0 θk (g), θk (fk+1 ) < 1/2k+1 , that is, σ0 (Fk g, Fk fk+1 ) < 1/2k+1 .
It follows that if gk+1 ∈ Nk+1 for all k, then h−1 n is a Cauchy sequence and so
−1
converges. As x 7→ x is continuous, the sequence hn also converges.
As the metric σ0 of uniform convergence on C∞ is a right invariant metric on the
Möbius group M, and as (M, σ0 ) is complete, we can apply Theorem 4.2 to M.
Now (by right invariance) for each z,
¡ ¢
σ Gn (z), hFn (z) ≤ σ0 (Gn , hFn ) = σ0 (hn , h),
so we obtain the next result (which contains Theorem 4.1).
Corollary 4.3. Let f1 , f2 , . . . be any sequence of Möbius maps. Then, for each k,
there is a neighbourhood Nk of fk such that if¡ gj ∈ Nj , j = 1, 2, . . ., then
¢ there
is some Möbius map h such that for all z, σ g1 · · · gn (z), hf1 · · · fn (z) → 0 as
n → ∞. In particular, for each point z, limn g1 · · · gn (z) exists if and only if
limn f1 · · · fn (z) exists.
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 555
We remark that the argument given here extends to Möbius maps acting in any
dimension. Also, even though it is not needed for our application to Möbius
maps, the following remark may be of interest. A fundamental system of neigh-
bourhoods of the identity e in a topological group G is a sequence of neighbour-
hoods U1 , U2 , . . . of e such that
(i) ∩n Un = {e}, and
(ii) for any neighbourhood V of e, Un ⊂ V for some n.
If such a system exists, then the topology on G is derived from a right invariant
metric on G (see [32, p. 39]).
Theorem 4.1 was used in [56] to prove that if Ω is a simple convergence set for
continued fractions, then so is its closure Ω, and it is clear that Corollary 4.3 can
also be used in this way. Also, Theorem 4.1 extends earlier results in [50], where
estimates of rn and sn were given for certain classes of continued fractions. No
explicit estimates were obtained in [56]. However, despite the generality of our
ideas, we can give explicit examples of the neighbourhoods Nk in the case of all
complex Möbius maps. Using (3.2) and (3.5) (and temporarily using norms of
Möbius maps when we really mean norms of their matrix representations) we
have
σ0 (Fk v, Fk fk+1 ) ≤ L(Fk )σ0 (v, fk+1 )
≤ L(f1 ) · · · L(fk ) 2 kfk+1 k kv − fk+1 k
< 2kf1 k2 · · · kfk k2 kfk+1 k kv − fk+1 k.
Thus, in the case of complex Möbius maps, we may take Nk to be the set of
Möbius g that satisfy
1
kg − fk k < k+2 .
2 kf1 k · · · kfk−1 k2 kfk k
2
It is possible to obtain estimates of the rn and sn in Theorem 4.1 from this, but
we suggest that the Nk are the more natural neighbourhoods to use.
We turn now to discuss the second result on the stability of continued fractions.
By Theorem 3.11, the periodic continued fraction [s, s, . . .] converges when s
is loxodromic. The continued fraction [s1 , s2 , . . .] is a limit periodic continued
fraction of loxodromic type if sn → s for some loxodromic s, and it is natural
to expect that any such continued fraction should also converge. This is true
and for a discussion of this, and of Perron’s contribution to this topic, see [51,
p. 108]. However, this result has nothing to do with continued fractions, or
even with Möbius maps (as we shall see later). First, Mandell and Magnus [61]
took this result out of the context of continued fractions by showing that if s
is loxodromic with repelling fixed point β, and if sn → s uniformly on C∞ , then
s1 · · · sn converges to a constant on C∞ \{β}. It is not necessary to assume that
the sn are loxodromic here (as is done in [61]) because the set L of loxodromics
is an open subset of M, so that sn must be loxodromic for all sufficiently large n.
556 A. F. Beardon CMFT
In fact, because L is open, the real issue here is one of stability, not convergence,
and we have the following result.
Theorem 4.4. Suppose that s is loxodromic with repelling fixed point β. Then
there is a neighbourhood N of s such that if sn ∈ N for all n, then s1 · · · sn
converges to a constant on C∞ \{β}.
Example 4.5. It is easy to see that in Theorem 4.4, the sequence s1 · · · sn (β)
may, but need not, converge. It will certainly converge if we take sn = s for all
n, and we now give an example in which this sequence is dense in C∞ . First,
we write the set of numbers (−1)m /m + (−1)n i/n, n, m = 1, 2, . . ., as a sequence
w1 , w2 , . . . in such a way that the partial sums of the series w1 + w2 + · · · is dense
in C∞ . Now let sn (z) = 2z + wn /2n−1 , and s(z) = 2z. Then sn → s uniformly
on C∞ (because wn → 0), and it is easy to see (by induction) that
√ ¡
s1 · · · sn (z) = 2n z + 2 w1 + · · · + wn ).
This implies that the sequence s1 · · · sn (0) is dense in C∞ .
Theorem 4.4 has probably been known for a long time, but in any event it is a
straightforward corollary of a result in function theory. The crucial fact is that
if α and β are the attracting and repelling fixed points of s, respectively, and
if D = {z : |z − α| < k|z − β|}, where k > 0, then s(D) is a compact subset
of D. Now let Ω be a subdomain of D such that s(D) ⊂ Ω ⊂ Ω ⊂ D. If g is any
Möbius map such that σ0 (g, s) is less than the chordal distance between s(D)
and ∂Ω, then g(D) ⊂ Ω ⊂ Ω and Theorem 4.4 follows directly from the next
(known) result.
Theorem 4.6. Suppose that K is a compact subset of a simply connected subdo-
main D of C∞ , and that f1 , f2 , . . . are analytic maps of D into K. Then f1 · · · fn
converges locally uniformly on D to some constant.
The essential feature here is the existence of a compact subset K of a domain D,
and a family of analytic maps of D into K. As we have already mentioned,
this implies the existence of a relatively compact subdomain Ω of D that con-
tains K, and this is precisely the situation that is used in the construction of
the compact-open topology on function spaces. This observation leads to the
following very general description of the phenomenon that is illustrated by limit
periodic loxodromic continued fractions.
Let X and Y be topological spaces, and let C(X, Y ) be the space of continuous
maps f : X → Y . Then the compact-open topology on C(X, Y ) (which is also
known as the topology of locally uniform convergence on X) is the topology
generated by the sets
[E, U ] = {g ∈ C(X, Y ) : g(E) ⊂ U },
where E is a compact subset of X, and U is an open subset of Y . Now let D be
a domain and consider any analytic f : D → D with the property that f (D) lies
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 557
The proof of Theorem 4.7 depends on removing the hypothesis of simple connec-
tivity from Theorem 4.6, so we need to examine the proof of Theorem 4.6. Now
proofs of Theorem 4.6 are given in, for example, [10], [46] and [54], but these do
not seem to generalize to give Theorem 4.7. In fact, the real reason why Theo-
rem 4.6 is true is that D supports a hyperbolic metric and as this has nothing to
with the connectivity of D, a proof based on this should yield Theorem 4.7. To
illustrate the main ideas here we shall give a proof of Theorem 4.6 in the case
when D is a disc; this proof uses only the hyperbolic metric and the standard
Contraction Mapping Theorem.
and w in D, and let γ be the geodesic that joins f (z) to f (w) in the metric space
(Ω, ρΩ ). Then
Z Z
¡ ¢ ¡ ¢
ρD f (z), f (w) ≤ λD (z) |dz| ≤ ` λΩ (z) |dz| = ` ρΩ f (z), f (w) .
γ γ
Now the general form of the Schwarz-Pick Lemma says that as f : (D, ρD ) →
(Ω, ρΩ ) is analytic, it is contracting in the two hyperbolic metrics; thus, for any
analytic map f : D → Ω, we have
¡ ¢ ¡ ¢
ρD f (z), f (w) ≤ ` ρΩ f (z), f (w) ≤ ` ρD (z, w).
As 0 < ` < 1, this shows that f acts on (D, ρ) as a strict contraction in the sense
of the familiar Contraction Mapping Theorem. Note that the scaling factor `
depends on D and Ω, but not on f ; thus the set of analytic maps f : D → Ω
satisfy a uniform contraction property with respect to the hyperbolic metric ρ D .
Now take any sequence f1 , f2 , . . . of analytic maps from D to Ω. Then, exactly
as in the standard proof of the Contraction Mapping Theorem, we have
ρD f1 · · · fn (z), f1 · · · fn fn+1 (z) ≤ `n ρD z, fn+1 (z) .
¡ ¢ ¡ ¢
As fn+1 (z) lies in K, and K is compact, the term ρD (z, fn+1 (z)) is (for each z)
bounded as a function of n, and the usual argument shows that f1 · · · fn (z) is a
Cauchy sequence and so converges to some value, say g(z). Finally, g is constant
because
ρD f1 · · · fn (z1 ), f1 · · · fn (z2 ) ≤ k n ρD (z1 , z2 ) → 0
¡ ¢
We have given Theorem 4.6 in the form that it is usually quoted, but this ar-
gument suggests that it would be more revealing to state it in the following
equivalent form: if Ω is a relatively compact subdomain of D, and if each f n
is an analytic map of D into Ω, then f1 · · · fn converges to a constant on D.
It is clear that essentially the same proof will work if we now assume that we
are dealing with a family of analytic maps f : D → Ω, where the subdomain Ω
has the property that all such maps f satisfy a uniform Lipschitz condition with
respect to the hyperbolic metric on D (notice that there is no compact set in this
formulation). In order to apply this general proof we need to characterize those
domains Ω that have this uniform Lipschitz property. This is done in [17], where
the generalization of Theorem 4.6 that is used above is proved.
We end this section with an example that illustrates how the ideas described
above can be used to give an explicit construction of the neighbourhood N of
a loxodromic Möbius map s that occurs in Theorem 4.4. It is not necessary to
insist that s(∞) = 0 here but in order to illustrate the effectiveness of argu-
ments involving conjugation even when considering the special case of continued
fractions we shall take s(∞) = 0.
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 559
√
Example 4.8. Let s(z) = 1/(z + 3i/ 2). Then s is loxodromic √ with attracting
√
fixed point α and repelling fixed point β, where α = −i/ 2 and β = −i 2.
Thus
s(z) − α z−α
=k ,
s(z) − β z−β
where k = s0 (α) = 1/2. Now let h(z) = (z − α)/(z − β), and let s̃ = hsh−1 . Thus
s̃(z) = 1/2z.
Now s̃ maps D into {z : |z| ≤ 1/2}. As σ(1/2, 1) = 2/5 > 1/2, we can define
p
Combining (5.3) and (5.4) we obtain (as promised earlier) a formula for L(g) in
terms of the coefficients of g, namely
1¡ p
L(g) = kgk2 + kgk4 − 4 ;
¢
(5.5)
2
in particular,
1
(5.6) kgk2 ≤ L(g) ≤ kgk2 ,
2
2 −1
¡ ¢
¡ −1L(g) ¢∼ kgk as kgk → +∞. Next, as g is an isometry, ρ j, g(j) =
and
ρ g (j), j , so we see from (5.4) that, for all g,
(5.7) L(g −1 ) = L(g) ≥ 1.
¡ ¢
The situation in which the terms kgk, L(g) and ρ j, g(j) simultaneously take
their minimum values is of interest.
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 563
Proof. It is easy to see that (5.3) and (5.4) show that (a), (b) and (c) are
equivalent. Further it is clear that (d) implies (b). Now suppose that (b) holds.
Then, equality holds in (5.7)) and we have
σ g(z), g(w) ≤ L(g) σ(z, w) = σ(z, w) = σ g −1 g(z), g −1 g(w)
¡ ¢ ¡ ¢
Proof. We recall that kgk = kGk, where G is any unimodular matrix represen-
tation of g. Now (3.7) implies that (for suitable choices of the signs of F and G)
564 A. F. Beardon CMFT
√
if σ0 (f, g) < 2, then kG − F k ≤ 2kF k, so that kGk ≤ kG − F k + kF k ≤ 3kF k.
The inequality (5.6) now gives
L(g) ≤ kgk2 = kGk2 ≤ 9kF k2 = 9kf k2 ≤ 18L(f )
as required.
We end this section with a brief sketch of a beautiful application of these ideas;
this is related to our generalization of continued fractions in as far as it describes
all finite groups of unitary maps. In this paragraph, a Möbius transformation
shall be any composition sequence of inversions in spheres (or planes) in R3 ∪{∞}.
Consider any finite Möbius group G. The orbit of the point j is a finite set E
in H3 , and it is not difficult to see that there is therefore a unique smallest closed
hyperbolic ball B that contains E. As E is invariant under G, so is B, and
hence so is the centre ζ of B. We can now map H3 onto B3 by a Möbius map h,
say, with h(ζ) = 0. Thus hGh−1 is a group of Möbius maps each element of
which fixes the origin and leaves B3 invariant. However, the only Möbius maps
that do this are the orthogonal maps (given by orthogonal matrices); thus we
have shown that every finite complex Möbius group is conjugate (in the group
of Möbius maps acting on R3 ∪ {∞}) to a finite group of rotations of R3 , and
hence to the symmetry group of one of the five Platonic solids, or to a cyclic or
dihedral group.
6. General convergence
The value of the continued fraction (1.1) generated by the sj is limn→∞ Sn (0).
We know that Sn (∞) = Sn−1 (0) and, presumably with this in mind, in 1957
Piranian and Thron proved the following result [66] and expressed the hope that
this might be useful in continued fraction theory.
Theorem 6.1. Suppose that a sequence gn of Möbius transformations converges
to the same value, say w, at two distinct points of C∞ . Then, with the possible
exception of one particular value of z, if gn (z) converges then it converges to w.
The sequence gn (z) = nz shows that the exceptional value of z in Theorem 6.1
can exist. We shall give a proof of Theorem 6.1 that is based on hyperbolic
geometry; this proof is short, free of coefficients, and valid in all dimensions. We
shall discuss a generalization of the results in [66] to all dimensions in Section
10.
A geometric proof of Theorem 6.1. It is sufficient to assume that there are
four distinct points z1 , z2 , z3 , z4 and three distinct values w1 , w2 , w3 such that
gn (z1 ) and gn (z2 ) tend to w1 , while gn (z3 ) → w2 and gn (z4 ) → w3 , and reach a
contradiction. By considering a conjugate sequence ggn g −1 for a suitable g, we
may suppose that no wj is ∞. Let γ12 be the geodesic in H3 with endpoints
z1 and z2 , let γ34 be the geodesic with endpoints z3 and z4 , and let d be the
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 565
hyperbolic distance between the geodesics γ12 and γ34 (we remark that given
any two hyperbolic geodesics with no end-points in common, there is a unique
hyperbolic segment that is orthogonal to both, and the length of this segment is
the hyperbolic distance between the two geodesics). As each gn is an isometry
of H3 , the distance between gn (γ12 ) and gn (γ34 ) is also d. However for large n,
gn (γ12 ) lies in small Euclidean neighbourhood of w1 , and gn (γ34 ) is approximately
the geodesic with endpoints w2 and w3 (when w2 6= w3 ) or it lies in a small
Euclidean neighbourhood of w2 (when w2 = w3 ). In both cases, the distance
between gn (γ12 ) and gn (γ34 ) tends to +∞ as n → ∞ and this is our contradiction.
At first sight Theorem 6.1 seems to justify the use of the point 0 in the definition
of convergence of a continued fraction, for if Sn (0) converges then so does Sn (∞)
and so, with possibly one exceptional value of z, limn→∞ Sn (z) (when it exists) is
independent of z. Despite this, there is a very good reason why the definition of
convergence of continued fractions should be modified. The problem is that Sn (z)
may converge (to a constant value) at many z, but perhaps not when z is 0 or ∞.
For example, consider any continued fraction in which the sequence s1 , s2 , . . . has
period three, and in which s1 s2 s3 is a loxodromic transformation g with attracting
and repelling fixed points α and β, respectively. Then the sequence g n of iterates
of g converges to α on C∞ \{β}, and it follows from this that Sn → α locally
uniformly on the complement of the three points β, s−1 −1
1 (β) and (s1 s2 ) (β). As
convergence is a topological rather than an arithmetic concept, it seems beyond
question that in this case one should regard the continued fraction as being
convergent, and assign the value α to it, regardless of whether or not Sn converges
at 0 (or at ∞). We now give an explicit example of this behaviour (this is a trivial
simplification of that given in [48]).
Example 6.2. Consider the continued fraction in which the Möbius sequence
s1 , s2 , s3 , . . . is s, t, s, s, t, s, s, t, s, . . . (of period three), where s(z) = 2/(2 + z)
and t(z) = 2/(−1 + z). A calculation shows that sts(z) = −1/2z; thus
µ ¶m
m 1
(sts) (z) = − z,
2
µ ¶m µ ¶
m 1 2
(sts) s(z) = − ,
2 2+z
µ ¶m µ
z−1
¶
m 1
(sts) st(z) = − .
2 z
Clearly, this continued fraction diverges. However, given any compact subset K
of C∞ \{0, −2, ∞} there is a constant k such that each of the functions (sts)m ,
(sts)m s and (sts)m st are bounded by k/2m on K; thus Sn → 0 locally uni-
formly on C∞ \{0, −2, ∞}. Surely, then, this continued fraction should (in any
reasonable definition of convergence) converge to 0?
566 A. F. Beardon CMFT
The role of (ii) will become clearer in Section 9. Finally, in 1986, Lisa Jacobsen
(now L. Lorentzen) proposed the following definition ([48, p. 480]).
Definition 6.4. The continued fraction K(an |bn ) in (1.1) converges generally,
or is generally convergent, to a value α in C∞ if there exist sequences un and vn
in C∞ such that
lim Sn (un ) = lim Sn (vn ) = α, lim inf σ(un , vn ) > 0.
n→∞ n→∞ n→∞
Möbius maps, nor in the higher dimensional situation) and by using cross-ratios.
However, the uniqueness of the limit is an immediate consequence of the following
geometric result which fully justifies the introduction of hyperbolic space, and
which may now be taken as the definition of general convergence.
Theorem 6.6. A sequence gn of Möbius maps converges generally to α in C∞
if and only if gn → α pointwise on H3 .
It is well known, and easy to see, that a sequence of Möbius maps converges
pointwise on H3 to a point α in C∞ if and only if it converges to α uniformly
on compact subsets of H3 . This is a direct consequence of the fact that each
Möbius map is an isometry of H3 and that the Euclidean length of a ‘ruler’ of
fixed hyperbolic length shrinks to zero as the ruler approaches the boundary
of hyperbolic space (as is beautifully illustrated in two dimensions by Escher’s
well-known tesselations of the unit disc). Our proof of Theorem 6.6 is based on
the following geometric lemma.
Lemma 6.7. Let γ be a geodesic in H3 whose endpoints u and v in C∞ are at a
chordal distance δ apart. Then γ passes within a hyperbolic distance cosh −1 (2/δ)
of the point j, where j = (0, 0, 1).
Proof. We map H3 onto the open unit ball B3 in R3 by a Möbius map that
is an extension of the stereographic projection of C∞ onto ∂B3 . This map is a
hyperbolic isometry, and it maps j to 0 in B3 . Let u0 and v 0 be the points on ∂B3
that correspond under this map to u and v in C∞ . Then (by the definition of the
chordal metric) σ(u, v) = |u0 −v 0 |. We now apply the angle of parallelism formula
(5.1) to Figure 6.1 and obtain cosh d sin θ/2 = 1. As σ(u, v) = |u0 −v 0 | = 2 sin θ/2,
we see that the distance d between j and γ is given by cosh d = 2/σ(u, v) as
required.
0r
v0
@ θ/2
@
d@
u0
Figure 6.1
Despite the fact that we have not yet discussed Möbius maps acting in higher
dimensions (see Section 8), we end this section by giving a formal definition of a
continued fraction in RN .
Definition 6.8. A continued fraction in RN is a sequence of Möbius maps
s1 , s2 , . . ., that act on RN ∪ {∞}, and that satisfy sn (∞) = 0 for all n. The
continued fraction [s1 , s2 , . . .] converges to α if the sequence gn converges gener-
ally to α.
With some minor modifications, Theorem 6.6 and its proof remain valid in all
dimensions, and so we see that the general convergence of a continued fraction
in RN is equivalent to the pointwise convergence of the sequence Sn on HN +1 .
7. Strong divergence
As convergence (whether classical or general) includes convergence to ∞, the
divergence of a sequence of Möbius maps implies an oscillatory behaviour. In
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 569
Presumably there are examples in which the sequence gn (j) accumulates both
at points of C∞ and points in H3 , and also examples in which gn (j) accumu-
lates at different points in C∞ but at no point of H3 . Such cases are likely to
be difficult to analyze, and they are closely related to the idea of a restrained
sequence of Möbius maps. A sequence gn of Möbius maps is restrained if and
only if kgn k → +∞ (although this was not the original definition of a restrained
sequence given in [49]); equivalently, by (5.4), if and only if the sequence gn (j)
acumulates only on the boundary of H3 . Such sequences have been examined
in detail by Beardon and Lorentzen (see [18]). By contrast, strong divergence is
easily analyzed. Because
1 ¡
kgn k2 = kIk2 + 4 sinh2 ρ j, gn (j) ,
¢
(7.1)
2
we have the following result.
Theorem 7.2. A sequence gn of complex Möbius maps is strongly divergent if
and only if the sequence kgn k is bounded.
We can now give a simple criterion for the inner composition sequence g1 · · · gn
to be strongly divergent, and as this is directly in terms of the generators g n of
the composition sequence this is more useful than Theorem 7.2 or Corollary 7.3.
The result is as follows.
P ¡ ¢1/2
Theorem 7.4. Let g1 , g2 , . . . be Möbius maps. If n kgn k2 − 2 converges,
then the sequence g1 · · · gn is strongly divergent.
570 A. F. Beardon CMFT
Proof. It is clear from (7.1), and the fact that sinh x ∼ x as x → 0, that the
three series
X¡ ¢1/2 X 1 X ¡
kgn k2 − 2
¢
, sinh ρ(j, gn (j), ρ j, gn (j) ,
n n
2 n
P ¡ ¢
converge or diverge together. Thus, by assumption, n ρ j, gn (j) = M , say,
where M < +∞. Let G0 = I, and Gn = g1 · · · gn . As the gj are isometries of H3 ,
we have
n−1
X n−1
¡ ¢ ¡ ¢ X ¡ ¢
ρ j, Gn (j) ≤ ρ Gm (j), Gm+1 (j) = ρ j, gm+1 (j) ≤ M.
m=0 m=0
This means that the points Gn (j) lie in some compact subset of H3 and so the
sequence Gn is strongly divergent.
Theorem 7.4 contains the Stern-Stolz Theorem and much more. If we apply it
to the sequence sn , where sn (z) = 1/(z + bn ) then (after representing sn by a
unimodular matrix) we find that ksn k2 = 2 + |bn |2 , and the Stern-Stolz Theorem
follows immediately. Note that in this case we find that
1 ¡ ¢
|bn | = 2 sinh ρ j, sn (j) .
2
We stress that this is an exact result (and not an estimate of |bn |), and that (as
far as I know) no such comparable interpretation is available if we confine our
attention to C∞ . In fact, we have obtained more than the classical Stern-Stolz
Theorem for we have shown that K(an |bn ) is strongly divergent. In addition,
we have used an argument that is valid in all dimensions. At this point it is
appropriate to mention a problem raised by Wall. In the footnote on [83, p. 33],
Wall raises the Pquestion of whether in the Stern-Stolz Theorem the conditional
convergence of n bn is sufficient to guarantee the divergence of the continued
fraction, As far as I know, this question has not yet been resolved. It may be that
the geometric approach taken here can shed more light on this matter; indeed, it
seems plausible that that the answer to Wall’s question may lie in understanding
the way that the orbit Sn (j) lies in H3 .
Let us now apply Theorem 7.4 to the general continued fraction generated by the
sequence sn (z) = an /(bn + z). In this case sn is represented by the unimodular
matrix à √ !
0 i an
i ibn
,
√ √
an an
so that
2 (1 − |an |)2 |bn |2
ksn k − 2 = + .
|an | |an |
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 571
P ¡ ¢1/2
Now¯ it is easy¯ to see that n ksn k2 −2 converges if and only if the two series
P P
¯1 − |an |¯ and
n n |bn | converge; thus Theorem 7.4 contains the following
result.
P ¯ ¯ P
Corollary 7.5. If n ¯1 − |an |¯ and n |bn | converge, then K(an |bn ) is strongly
divergent.
It is worthwhile to look at Theorem 7.4 in a different light. The Möbius maps
that fix j are the unitary maps; that is, g(j) = j if and only if kgk2 = 2. Thus
Theorem 7.4 says that the inner composition sequence g1 · · · gn is strongly diver-
gent if and only if (in some sense) the gn approach the subgroup of unitary maps
sufficiently rapidly. Thus, in this result we do not require that the gn converge,
but merely that each gn looks increasingly like some unitary map. With contin-
ued fractions in mind we note that the map s(z) = a/(b + z) is unitary if and
only if |an | = 1 and bn = 0, and this is the special case given in Corollary 7.5.
Corollary 7.5 can be derived from the Stern-Stolz Theorem (from the theory of
equivalent continued fractions), and it occurs in [57], but (as we shall see later)
Theorem 7.4 applies to all Möbius maps in all dimensions.
We shall now discuss the following two results which we express in terms of
matrix norms (rather than sequences of coefficients as is done in [57] and [61]).
Theorem 7.6. Suppose
P that s given by s(z) = a/(b + z) is elliptic, that sn (z) =
an /(bn + z) and n ksn − sk converges. Then K(an |bn ) diverges. Further, if
z0 is a fixed point of s, then Sn (z0 ) converges to some limit, say w, and then
Sn−1 (w) → z0 .
P
Theorem 7.7. Suppose that tn → t, where each tn and t is elliptic. If n ktn −tk
converges then t1 · · · tn converges to distinct values at the two fixed points of t,
and diverges elsewhere.
The following result is equivalent to Theorem 7.7 in the case of complex Möbius
maps, but it (and the proof) generalizes to higher dimensions. Moreover, it is
easier to prove if we use the metric σ0 rather than the matrix norm (or the
sequences of coefficients).
P
Theorem 7.8. Let sn and s be complex Möbius maps. Suppose that n σ0 (sn , s)
converges, and that s lies in a compact subgroup G of M. Then s1 · · · sn converges
at the two fixed points of s, and diverges elsewhere.
The crucial step in the proof of Theorem 7.8 depends directly on the compactness
of G rather than the nature of the elliptic maps. Although it is true that every
elliptic map lies in a compact subgroup of M, and that every element of a
compact subgroup of M is elliptic, it is compactness that is the decisive feature
for the following reason. By (5.5), the map g 7→ L(g) is continuous and positive,
so that if G is a compact subgroup of M, then
(7.2) L(G) = sup{L(g) : g ∈ G} < +∞.
572 A. F. Beardon CMFT
Of course, we can apply this to sk , sk+1 , . . . and thereby obtain the inequality
We know that s is elliptic. Suppose now that z is not a fixed point of s; then the
sequence sn (z) diverges so that there is a positive d such that for every m the
set {sn (z) : n ≥ m} has chordal diameter at least d. Now choose k such that
∞
X d
L(G) σ0 (sk , s) < .
n=k
4
Then, from (7.3), we have σ0 (sk sk+1 · · · sk+n−1 , sn ) < d/4. It is now obvious
that the sequence sk · · · sn (z), n = k, k + 1, . . ., diverges and hence so does the
sequence s1 · · · sn (z), n = 1, 2, . . .. This shows that s1 · · · sn diverges at every z
that is not fixed by s.
Now let w be a fixed point of s. We shall show that s1 · · · sn (w) is a Cauchy
sequence and so converges. First, from (7.3), we see that for k ≥ K, say,
√
σ0 (sk sk+1 · · · sk+n−1 , sn ) ≤ 2
for all n. This combined with Lemma 5.5 then shows that for all m,
It is now clear that there is some constant M such that for all n, L(s1 · · · sn ) ≤ M ;
in other words, the family of s1 · · · sn is uniformly Lipschitz.
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 573
≤ M σ0 sn , sm+1 · · · sm+n
¡ ¢
£ ¤
≤ L(G) M σ0 (sm+1 , s) + · · · + σ0 (sm+n , s) ,
P
the last inequality here following from (7.3) with k = m + 1. As n σ0 (sn , s)
converges we conclude that s1 · · · sn (w) is a Cauchy sequence and this completes
the proof.
Essentially the same argument (in which we replace the uniform Lipschitz con-
dition by the inequality L(sn ) ≤ L(s)n ) yields the following result.
Theorem 7.9. Let sn and s be complex Möbius maps. Suppose that the series
n n
P
n σ0 (sn , s)L(s) converges. If the sequence s (z) diverges then so does the
sequence s1 · · · sn (z). Further, s1 · · · sn converges at the two fixed points of s.
There are other results of this type to be found in, for example, [41] and [61],
and it may be that these and other results of a similar nature can be proved by
geometric means. For example, Thron and Waadeland have proved the following
result ([77, Proposition 1, p. 652]).
Theorem 7.10. Let s(z) = a/(b + z) have fixed points x and y, and suppose
that r satisfies 0 < r < |x/y| < 1/r. Then there is a positive A such that if
|an − a| ≤ Ar n and |bn − b| ≤ Ar n , then the sequence s1 · · · sn (x) converges,
where sn (z) = an /(bn + z).
The following similar result follows directly from Theorem 7.9 (and this too could
be phrased in terms of matrix norms or coefficients).
Theorem 7.11. Let sn and s be Möbius maps, and suppose that µ < L(s). If
there is some A such that for all n, σ0 (sn , s) ≤ Aµn , then the sequence s1 · · · sn (x)
converges for each fixed point x of s.
valid for the corresponding action on all of R3∞ . As there are many arguments
in continued fraction theory that involve cross-ratios (see, for example, [79] and
[82]) this remark shows how some of these might be generalised to R3∞ . This idea
will be exploited in our discussion of tail sequences in Section 9.
Our first model of hyperbolic space is
HN +1 = {(x1 , . . . , xn+1 ) : xn+1 > 0},
where this is equipped with the hyperbolic metric derived from the line element
ds = |dx|/xN +1 . Observe that the boundary of HN +1 is RN ∞ . Our second model
of hyperbolic space is the unit ball B N +1
in R N +1
with the hyperbolic metric
given by
2|dx|
ds = .
1 − |x|2
For brevity, we shall denote the hyperbolic metric in both of these models, and
in all dimensions, by ρ; it should be clear from the context which space ρ refers
to (and in any case the two models are isometrically equivalent). Finally, we
note that both models of hyperbolic space are complete metric spaces. Shortly,
we shall discuss a third model of hyperbolic space.
We shall now describe how Möbius maps act on these spaces, and, in particular,
we shall show how stereographic projection ties together the three spaces that
we have discussed so far. A Möbius map acting in RN ∞ is a finite composition
γ1 · · · γk , where each γj is an inversion (or a reflection) in some (n−1)-dimensional
hyperplane or hypersphere in RN ∞ . The Möbius group acting on R∞ is the group
N
√
Note that Σ has centre eN +1 , and radius 2, and that SN ⊂ Σ. Now α satisfies
¡ ¢
α(y) = eN +1 + t y − eN +1 ,
where t is chosen so that |α(y) − eN +1 | |y − eN +1 | = 2, and this gives
|y − eN +1 |2 − 2
µ ¶ µ ¶
2
(8.5) α(y) = y+ eN +1 .
|y − eN +1 |2 |y − eN +1 |2
If y ∈ RN , then y is orthogonal to eN +1 , so that |y − eN +1 |2 = |y|2 + 1. In this
case we see from (8.1) and (8.5) that ϕ(y) = α(y), and so we have proved that
the stereographic projection ϕ is just the restriction of α to RN . This confirms
that the chordal metric is inextricably linked with inversions, and hence with the
Möbius group.
As α(RN ∞ ) = S , it follows that α must map H
N N +1
bijectively onto either the
unit ball B N +1
, or the exterior E N +1
of the closure of this ball. Now it is evident
(either from (8.5), or geometrically) that α interchanges 0 and −eN +1 ; thus we
conclude that α(HN +1 ) = EN +1 . By considering the lower half space of RN +1 ,
the following result is now clear.
Theorem 8.1. Let α be the inversion in the sphere given in (8.4), and let β be
the reflection across the plane xN +1 = 0. Then αβ is in MN +1 , αβ = ϕ on RN
∞,
αβ(eN +1 ) = 0, and αβ(H N +1
)=B N +1
.
Next, we observe that (as in the complex case) the chordal metric induces the
metric
σ0 (f, g) = sup{σ f (x), g(x) : x ∈ RN
¡ ¢
∞}
on the space of continuous maps of R∞ into itself. The restriction to the space of
N
With this, the same proof as in the complex case yields the following result.
Theorem 8.3. The group MN with the metric σ0 is both a complete metric
space, and a topological group.
We end this section with a discussion of a generalization of the way that the
action of complex Möbius maps on H3 can be expressed through the algebra of
quaternions. The Clifford algebra CN is the associative algebra over R generated
by the N elements i1 , i2 , . . . , iN subject to the relations ij 2 = −1 and ij ik =
−ik ij when j 6= k. Explicitly, Pthis means that each element of CN has a unique
representation in the form I aI I, where the aI are real, and where in this
sum I runs over all products iν1 iν2 · · · iνp , where 1 ≤ ν1 < · · · < νp ≤ N . The
empty product is allowed, and it is interpreted as the real number 1. The cases
N = 0, 1, 2 give the real numbers, the complex numbers, and the quaternions,
respectively; for quaternions we have i1 = i, i2 = j, and k = i1 i2 .
Clearly, CN is a real normed vector space of dimension 2N , with the norm |a| and
scalar product a·b defined in the natural way by
³ X ´1/2 X
|a| = aI 2 , a·b= aI b I .
I I
N +1
Let V be the subspace of elements in CN of the form a0 + a1 i1 + · · · + aN iN .
We refer to these elements as vectors, and we identify V N +1 with RN +1 in the
obvious way. We also adjoin ∞ to V N +1 to form V∞N +1 , and this is identified
with RN +1
∞ .
There are three involutions that play a central role in the theory: we shall
define each of these on the basis elements iν1 iν2 · · · iνp , and each is then ex-
tended to act on CN by linearity. The first of these is a 7→ a0 : this is the map
iν1 iν2 · · · iνp 7→ (−iν1 )(−iν2 ) · · · (−iνp ). The second involution is a 7→ a∗ : this
is the map ir1 · · · irs 7→ irs · · · ir1 . Finally, as these two involutions commute,
they induce a third involution defined by x 7→ x̄, where x̄ = (x0 )∗ = (x∗ )0 .
These involutions interact with multiplication in the following way: (ab) 0 = a0 b0 ,
(ab)∗ = b∗ a∗ , and (ab) = ba. Next, all three involutions map V N +1 into itself,
580 A. F. Beardon CMFT
The proof of Theorem 8.10 is geometric, but does an algebraic proof exist?
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 581
9. Backward orbits
Theorem 6.1 shows that if a continued fraction converges strongly to α then,
apart from possibly one exceptional value of z, Sn (z) → α whenever this limit
exists. Moreover, the following result ([78, Theorem 2.1]) tells us where this
limit exists, and it also illuminates the role of (ii) in Definition 6.3 of strong
convergence.
Theorem 9.1. Suppose that K(an |bn ) converges strongly to a complex number α.
Then Sn → α on C∞ \L, where L is the derived set of the sequence Sn−1 (∞).
Theorem 9.1 suggests that it might be profitable to study the sequence Sn−1 (w) of
inverse images of a point w for different w but before we do this, let us consider
how the same concept plays a crucial role in complex dynamics and in discrete
Möbius groups. Historically, the theory of discrete groups was developed before
iteration theory, but as most readers will probably be more familiar with the
latter, we shall discuss iteration first.
Let R be a rational function. Then R induces a partition of the complex sphere
C∞ into two sets, namely the Julia set J (R) and the Fatou set F(R). The Julia
set is compact, and the Fatou set is open (but not necessarily non-empty). The
two basic properties of these sets are
(a) apart from possibly one exceptional value of w, the Julia set J (R) is the
derived set of the backward orbit of w, that is of the set
∞
[
{z : Rn (z) = w},
n=1
continued fractions) special and, once again, we see the value of extending the
action of Möbius maps from Euclidean space to hyperbolic space. We begin our
discussion of inverse orbits with the following basic lemma.
Lemma 9.3. Suppose that g1 , g2 , . . . is a sequence of Möbius maps acting in RN∞
that converges generally to α. If w ∈ RN ∞
+1
and w 6
= α then g n → α locally
uniformly on RN +1
∞ \Λ(w).
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 583
Thus we see that there is a positive M (which depends on w and K, but not on
n) such that for all sufficiently large n,
σ(gn x, gn bn ) ≤ M σ(gn an , gn bn ).
This was proved under the assumption that x 6= bn , but it clearly holds when
x = bn ; thus it holds for all x in K. As (9.1) holds, we see that gn → α uniformly
on K, and hence locally uniformly on RN +1
∞ \Λ(w).
Proof. We may assume that u 6= v. Let us suppose that Λ(u) 6= Λ(v). Then
(by interchanging u and v if necessary) there is some ζ in Λ(u) but not in Λ(v).
Now find a neighbourhood Nα of α that does not contain u, and also a neigh-
bourhood Nζ of ζ whose closure is disjoint from the closed set Λ(v). As v 6= α,
Lemma 9.3 implies that gn → α uniformly on Nζ , so that for all sufficiently
large n, gn (Nζ ) ⊂ Nα . For these n, gn−1 (u) ∈
/ Nζ so that ζ cannot be in Λ(u),
contrary to our assumption.
Lemma 9.4 enables us to make the following definition.
Definition 9.5. Suppose that g1 , g2 , . . . is a sequence of Möbius maps acting in
RN∞ that converges generally to α. Then the limit set Λ of the sequence g n is the
set Λ(w) for all w in RN +1
∞ \{α}.
Lemma 9.4, contains infinitely many points gn−1 (u) for each u other than α.
Clearly, this prevents gn from converging uniformly to α on any neighbourhood
of x.
In addition to the result in Theorem 9.6, it is also true that RN +1
∞ \Λ(w) is
the largest open set in R∞ on which {g1 , g2 , . . .} is a normal family (see [2]).
N +1
Because of this result, and the strong similarity it has with the earlier theorems,
we might also call Λ the Julia set of the sequence g1 , g2 , . . .. As the generally
convergent sequences include convergent continued fractions as a special case,
we should now ask whether any other theorems about Julia sets or limit sets (in
discrete group theory) have counterparts in the theory of continued fractions.
Moreover, we might ask whether Λ is (always) fractal, and does it have any
self-similarity properties? Someone should provides pictures of the sets Λ for
different continued fractions.
There is one more topic in this section, and this alone fully justifies (even in
the context of classical continued fractions) our consideration of the extension
of Möbius maps from RN ∞ to H∞ . Let gn be a sequence of complex Möbius
N +1
is the n-th iterate of g, then gn converges on A ∪ {∞} but nowhere else, and this
case is not covered by Theorem 10.1 in its present form. The following result is
the desired generalization of Theorem 10.1 (see [22]).
Theorem 10.2. Let gn be any sequence of Möbius maps acting on RN ∞ , and
suppose that the gn converge pointwise on the non-empty set C (and nowhere
else) to the function g. Then one of the following occurs:
(a) g is the restriction of some Möbius map to C;
(b) C = RN ∞ , and g takes exactly two values, one of which is taken exactly once;
(c) g is constant on C.
Suppose that we are in case (b); thus there is some x0 such that, say, gn → α
pointwise on RN ∞ \{x0 }, and gn (x0 ) → β, where α 6= β. In these circumstances
one can show that gn → α locally uniformly on RN ∞ \{x0 }. Now consider the
sets C in (a). If N = 2 then, by Theorem 10.1 C is either C∞ or some double-
ton {z1 , z2 }. However, in higher dimensions other possibilities can occur and a
complete description of the possibilities is given in our next result.
Theorem 10.3. Suppose that g and g1 , g2 , . . . are Möbius maps acting on RN
∞,
and that gn converges pointwise to g on, and only on, the non-empty set C. Then
there is a subspace V of RN of dimension d, where d 6= N − 1, and a Möbius
map h, such that C = h(V ∪ {∞}). Further, every set C of this form arises as
the set C for some sequence gn .
These results give insight into the two-dimensional case. If N = 2, then V in
Theorem 10.3 is {0} or C and we obtain the following corollary which explains
part of the conclusion in Theorem 10.1. For other results of this type in all
dimensions, see also [2].
Corollary 10.4. Suppose that g and g1 , g2 , . . . are complex Möbius maps, and
that gn → g on, and only on, the non-empty set C. Then either C = C∞ or C
contains exactly two points.
References
1. W. Abikoff, The bounded model for hyperbolic 3-space and a quaternionic Uniformization
Theorem, Math. Scand. 54 (1984), 5–16.
2. B. Aebischer, The limiting behaviour of sequences of Möbius transformations, Math. Zeit.
205 (1990), 49–59.
3. B. Aebischer, Stable convergence of sequences of Möbius transformations, in: Computa-
tional Methods and Function Theory (CMFT ’94), eds. R. M. Ali, St. Ruscheweyh and E.
B. Saff, World Scientific Pub. Co., 1995, 1–21.
4. L. V. Ahlfors, Hyperbolic motions, Nagoya Math. J. 29 (1967), 163–166.
5. L. V. Ahlfors, Möbius Transformations in Several Dimensions, Univ. Minnesota Lecture
Notes, Minnesota, 1981.
6. L. V. Ahlfors, On the fixed points of Möbius transformations in Rn , Ann. Acad. Sci. Fenn.
Ser. A. 1. Math. 10 (1985), 15–27.
7. L. V. Ahlfors, Möbius Transformations and Clifford Numbers, in Differential Geometry
and Complex Analysis, eds. I. Chavel, and H.M. Farkas, Springer-Verlag, Berlin (1985),
65–73.
8. L. V. Ahlfors, Clifford numbers and Möbius transformations in Rn , Clifford algebras and
their applications in mathematical physics, (Canterbury, 1985), 167–175, NATO Adv. Sci.
Inst. Ser. C Math. Phy. Sci., 183, Reidel, Dordrecht, 1986.
9. L. V. Ahlfors, Möbius transformations in Rn expressed through 2 × 2 matrices of Clifford
numbers, Complex Variables Theory Appl. 5 (1986), 215–224.
10. I. N. Baker and P. J. Rippon, Towers of exponents and other composite maps, Complex
Variables 12 (1989), 181–200.
11. A. F. Beardon and B. Maskit, Limit points of Kleinian groups and finite-sided fundamental
polyhedra, Acta Math. 132 (1974), 1–12.
12. A. F. Beardon, The Geometry of Discrete Groups, Springer-Verlag, 1983.
13. A. F. Beardon and J. B. Wilker, The norm of a Mobius transformation, Math. Proc. Camb.
Phil. Soc. 96 (1984), 301–308.
14. A. F. Beardon, Iteration of Rational Functions, Springer-Verlag, 1991.
15. A. F. Beardon, The geometry of Pringsheim’s continued fractions, Geometriae Dedicata
84 (2001), 125–134.
16. A. F. Beardon and D. Minda, Sphere-preserving maps in inversive geometry, Proc. Amer.
Math. Soc. 130 (2002), 987–998.
17. A. F. Beardon, T. K. Carne, D. Minda, and T. W. Ng, Random iteration of analytic maps,
to appear in Ergodic Th. & Dyn. Systems.
18. A. F. Beardon and L. Lorentzen, Continued fractions and restrained sequences of Möbius
maps, to appear in Rocky Mount. J. Math..
19. A. F. Beardon, Continued fractions, Möbius transformations and Clifford algebras, to
appear in Bull. London Math. Soc..
20. A. F. Beardon, The Hillam-Thron Theorem in higher dimensions, to appear in Geom.
Dedicata.
21. A. F. Beardon, Repeated compositions of analytic maps, to appear in Comp. Methods and
Function Theory.
22. A. F. Beardon, The pointwise convergence of Möbius maps, preprint, 2002.
23. R. Benedetti and C. Petronio, Lectures on hyperbolic geometry, Universitext, Springer-
Verlag, 1992.
24. C. Cao and P. L. Waterman, Conjugacy invariants of Möbius groups, Quasiconformal
mappings and analysis (Ann Arbor, MI, 1995), Springer, New York, 1998, 109–139.
25. L. Carleson and T. W. Gamelin, Complex Dynamics, Springer-Verlag, 1993.
592 A. F. Beardon CMFT
26. S. S. Chen and L. Greenberg, Hyperbolic spaces, in: Contributions to Analysis, eds L. V.
Ahlfors, I. Kra, B. Maskit, and L. Nirenberg, Academic Press, 1974.
27. H. S. M. Coxeter, Inversive distance, Ann. Mat. Pura Appl. 71 (1966), 73–83.
28. H. S. M. Coxeter, The inversive plane and hyperbolic space, Hamburger Math. Abh. 29
(1966), 217–242.
29. H. S. M. Coxeter, The Lorentz group and the group of homographies, Proc. Intern. Conf.
Theory of Groups (Canberra 1965), Gordon and Breach, New York, 1967.
30. H. Davenport, The Higher Arithmetic, Sixth Edn, Cambridge Univ. Press, 1995.
31. J. D. de Pree and W. J. Thron, On sequences of Moebius transformations, Math. Zeit. 80
(1962), 184–193.
32. J. Dieudonné, Treatise on Analysis, Vol II, Academic Press, 1970.
33. P. Du Val, Homographies, Quaternions and Rotations, Clarendon Press, Oxford, 1964.
34. W. Fenchel, Elementary Geometry in Hyperbolic Space, De Gruyter Studies in Mathemat-
ics 11, de Gruyter, 1989.
35. L. R. Ford, A geometrical proof of a theorem of Hurwitz, Proc. Edinburgh Math. Soc. 35
(1917), 59–65.
36. L. R. Ford, Rational approximations to irrational complex numbers, Trans. Amer. Math.
Soc. 19 (1918), 1–42.
37. L. R. Ford, Fractions, Amer. Math. Monthly 45 (1938), 586–601.
38. L. R. Ford, Automorphic functions, Chelsea Pub. Co., Second Edition, 1951.
39. F. W. Gehring and G. J. Martin, The Matrix and Chordal Norms of Möbius Transforma-
tions, Complex Analysis, Birkhäuser, Basel, 1988, 51–59.
40. F. W. Gehring and G. J. Martin, Inequalities for Möbius transformations and discrete
groups, J. Reine Angew. Math. 418 (1991), 31–76.
41. J. Gill, Infinite compositions of Möbius transformations, Trans. American Math. Soc. 176
(1973), 479–487.
42. P. G. Gormley, Stereographic projection and the linear fractional group of transformations
of quaternions, Proc. Royal Irish Acad. Sect A 51 6 (1947), 67–85.
43. L. Greenberg, Discrete subgroups of the Lorentz group, Math. Scand. 10 (1962), 85–107.
44. W. R. Hamilton, On continued fractions in quaternions, Phil. Mag., iii (1852), 371–373; iv
(1852), 303; v (1853), 117–118, 236–238, 321–326.
45. W. R. Hamilton, On the connexion of quaternions with continued fractions and quadratic
equations, Proc. Royal Irish Acad. 5 (1853), 219–221, 299–301.
46. P. Henrici, Applied and Computational Complex Analysis, Vol. 2, Special Functions, In-
tegral Transforms, Asymptotics and Continued Fractions, Wiley, New York, 1977.
47. K. L. Hillam and W. J. Thron, A general convergence criterion for continued fractions
K(an /bn ), Proc. Amer. Math. Soc. 16 (1965), 1256–1262.
48. L. Jacobsen, General convergence of continued fractions, Trans. Amer. Math. Soc. 294
(1986), 477–485.
49. L. Jacobsen and W. J. Thron, Limiting structures for sequences of linear fractional trans-
formations, Proc. Amer. Math. Soc. 99 (1987), 141–146.
50. L. Jacobsen, Nearness of continued fractions, Math. Scand. 60 (1987), 129–147.
51. W. B. Jones and W. J. Thron, Continued Fractions: Analytic Theory and Applications,
Encyclopedia of Mathematics and its Applications, 11, Addison-Wesley, Reading, Mass,
(1980). Now distributed by Cambridge Univ. Press, New York.
52. I. L. Kantor and A. S. Solodovnikov, Hypercomplex Numbers, Springer-Verlag, 1989.
53. R. E. Lane, The convergence and values of periodic continued fractions, Bull. Amer. Math.
Soc. 51 (1945) 246–250.
54. L. Lorentzen, Compositions of contractions, J. Comput. Appl. Math. 32 (1990), 169–178.
1 (2001), No. 2 Continued Fractions, Discrete Groups and Complex Dynamics 593
55. L. Lorentzen and H. Waadeland, Continued Fractions and Some of its Applications, North-
Holland, 1992.
56. L. Lorentzen, The closure of convergence sets for continued fractions are convergence sets,
Proc. Ednburgh Math. Soc., 37 (1993), 39–46.
57. L. Lorentzen, Divergence of continued fractions related to hypergeometric series, Math.
Comp. 62 (1994), 671–686.
58. L. Lorentzen, A Convergence Property for Sequences of Linear Fractional Transformations,
Continued fractions and orthogonal functions (Loen, 1992), 281–304, Lecture Notes in Pure
and Appl. Math., 154, Dekker, New York, 1994.
59. L. Lorentzen, A convergence question inspired by Stieltjes and by value sets in continued
fraction theory, J. Comp. Appl. Math. 65 (1995), 233–251.
60. L. Lorentzen, Convergence of compositions of self-mappings, Ann. Univ. Marie Curie
Sklodowska A 53 13 (1999), 121–145.
61. M. Mandell and A. Magnus, On Convergence of Sequences of Linear Fractional Transfor-
mations, Math. Zeit. 115 (1970), 11–17.
62. J. W. Magnus, Non-Euclidean Tessselations and their Groups, Academic Press, 1974.
63. J. Milnor, Hyperbolic geometry — the first 150 years, Bull. Amer. Math. Soc. 6 (1982),
9–24.
64. P. J. Nicholls, The Ergodic Theory of Discrete Groups, London Math. Soc. Lectures Notes
143, Camb. Univ. Press, 1989.
65. J. F. Paydon and H. S. Wall, The continued fraction as a sequence of linear transformations,
Duke Math. Jour. 9 (1942), 360–372.
66. G. Piranian and W. J. Thron, Convergence properties of sequences of linear fractional
transformations, Michigan Math. J. 4 (1957), 129–135.
67. O. Perron, Die Lehre von den Kettenbrüchen, Band I, Teubner, Stuttgart, 1954.
68. O. Perron, Die Lehre von den Kettenbrüchen, Band II, Teubner, Stuttgart, 1957.
69. I. R. Porteous, Topological Geometry, Van Nostrand, 1969.
70. J. G. Ratcliffe, Foundations of hyperbolic manifolds, Graduate Texts 149, Springer-Verlag,
1994.
71. D. E. Roberts, On a representation of vector continued fractions, J. Comput. Appl. Math.
105 (1999), 453–466.
72. H. Schwerdtfeger, Moebius transformations and continued fractions, Bull. Amer. Math.
Soc. 52 (1946), 307–309.
73. T. J. Stieltjes, Recherches surles fractions continues, Ann. Fac. Sci. Toulouse, 8J (1894),
1–122; 9A (1894), 1–47; Oeuvres, 2, 402–566.
74. N. Steinmetz, Rational Iteration, De Gruyter Studies in Mathematics 16, de Gruyter, 1993.
75. W. J. Thron, Convergence regions for continued fractions and other infinite processes,
Amer. Math. Monthly 68 (1961), 734–750.
76. W. J. Thron, Convergence of Sequences of Linear Fractional transformations and of Con-
tinued Fractions, J. Indian Math. Soc. 27 (1963), 103–127.
77. W. J. Thron and H. Waadeland, Convergence questions for limit periodic continued frac-
tions, Rocky Mountain J. Math. 11 (1981), 641–657.
78. W. J. Thron and H. Waadeland, Modifications of continued fractions, a survey, Analytic
theory of continued fractions (Loen, 1981), 38–66, Lecture Notes in Math.. 932, Springer,
Berlin-New York, 1982.
79. W. J. Thron, Continued fraction identities derived from the invariance of the crossratio
under l.f.t., Analytic theory of continued fractions, III (Redstone, CO, 1988), 124–134,
Lecture Notes in Math.,1406, Springer, Berlin, 1989.
80. H. Waadeland, Tales about tails, Proc. Amer. Math. Soc. 90 (1984), 57–64.
594 A. F. Beardon CMFT