A composition law and refined notions of convergence
for periodic continued fractions
Bradley W. Brock
Center for Communications Research, 805 Bunn Drive, Princeton,
NJ 08540-1966, USA
[email protected], Bruce W. Jordan
Department of Mathematics, Baruch College, The City University
of New York, One Bernard Baruch Way, New York, NY 10010-5526, USA
[email protected] and Lawren Smithline
Center for Communications Research, 805 Bunn Drive, Princeton,
NJ 08540-1966, USA
[email protected]
Abstract.
We define an equivalence relation on periodic continued fractions with
partial quotients in a ring , a group law on these
equivalence classes, and a map from these equivalence classes to matrices
in with determinant .
We prove this group of equivalence classes is isomorphic to and study
certain of its one- and two-dimensional representations.
For a periodic continued fraction with period , we give a refined description
of the limits of the different -decimations of its sequence of convergents.
We show that for a periodic continued fraction associated to a matrix with eigenvalues of
different magnitudes, all of these limits exist in and a strict majority
of them are equal.
Key words and phrases:
elementary matrices,
2020 Mathematics Subject Classification:
Primary 11J70; Secondary 11G30
1. Introduction
Continued fractions with positive integer partial quotients
are a staple of classical number theory, with
an intimate connection to Euclid’s algorithm.
Here, we consider the more general case of continued fractions with partial
quotients in a subring of the complex numbers ,
as in [BEJ21].
We define an equivalence relation on the set
of periodic continued fractions
over and a group law on the resulting equivalence classes
.
The theory of periodic continued fractions, unlike continued fractions in
general, has connections to both arithmetic groups and diophantine equations.
The work [BEJ21] relates the theory of to the group
, the subgroup of consisting of matrices with
determinant , and to diophantine problems on associated varieties.
In this paper we further explore the relationship between and the group
.
A major influence in our study is the 1966 paper [Coh66] by Cohn.
Section 2 sets out preliminaries related to
actions of matrices.
We consider the action of by linear fractional transformations
on the projective line .
We also name subgroups of determined
by common eigenspaces.
In Section 3, we introduce the
concatentation binary operation
on the semigroup of finite continued fractions over .
We develop an equivalence relation on and show that
the equivalence classes form a group under .
We exhibit a map in
Proposition 3.12 arising
from the algorithm for computing the value of a
finite continued fraction.
We show that is well defined on equivalence classes,
giving a homomophism .
We review
results of Cohn [Coh66] that give information on
for certain rings .
Using this, we show how
the standard amalgam presentation
arises from
presenting as an explicit
quotient of
Section 4 extends the structure
on finite continued fractions to periodic continued fractions (PCFs).
In Theorem 4.9, we show that this yields a
natural equivalence relation
on periodic continued fractions
yielding a group of equivalence classes isomorphic to .
Section 5 relates the group to quantities defined
in [BEJ21].
We review the fundamental matrix
associated to a periodic continued fraction and
apply the isomorphism of Theorem 4.9 to show that is well-defined
on , giving a homomophism .
For we define subgroups of
by requiring
to be a linear combination of and the identity matrix .
We exhibit characters
corresponding to each eigenvalue of .
These characters have values that are units in a quadratic extension of .
In Section 6, we give examples where is a number ring.
In Section 7, we reformulate the PCF-convergence
criteria of [BEJ21, Thm. 4.3]
in parallel with the cases of convergence in Section 2.1.
We show that when has eigenvalues of different magnitudes
(which implies quasiconvergence),
the different -decimations of its sequence of convergents all converge in ,
and a strict majority agree. Unanimity is equivalent to convergence.
In Section 8, we demonstrate the sharpness of the result of Section
7. We give examples showing
that the proportion of agreement of
the different -decimations of the sequence of convergents of a
quasiconvergent periodic continued fraction can be arbitrarily close to or arbitrarily close to and different from . We show
that the equivalence relation on respects quasiconvergence
but does not respect convergence.
2. Actions of matrices
Let be a ring with .
A finite continued fraction with partial quotients in
is an iterated quotient.
It can be interpreted in terms of the action of
on the projective line by linear
fractional transformations which we review in this section.
We also name certain subgroups of determined by common eigenspaces.
2.1. Linear fractional transformations
Suppose
(1)
and .
The matrix acts on by
linear fractional transformation
where we interpret .
For , let
(2)
The vector is an eigenvector of if and only if .
For
nonzero polynomial ,
let be the multiset of zeros of in ,
with the convention that if , then is a simple root of ;
and if , then has a double root at .
When , say .
For a matrix as above, set
(3)
We use to denote .
When , is an eigenvector of with
eigenvalue
(4)
In particular, and .
Proposition 2.1.
Let , be matrices over , be the identity matrix.
Let be not both zero.
There exists such that
if and only if
.
Proof.
Suppose with
such that
.
The polynomial
is independent of .
We have
Conversely, suppose
for , .
Since is linear map, .
Thus, there is such that .
∎
For a matrix , and the eigenvalues of
govern the convergence behavior of the sequence , ,
for .
Definition 2.2.
For any ,
let denote the diagonal matrix
Proposition 2.3.
For , one of four mutually exclusive possibilities holds:
(a)
has one root of multiplicity 2
and for all .
(b)
. For all , and so
.
(c)
has distinct roots , and has corresponding eigenvalues
,
as in (4), with .
For all , .
For , .
(d)
has distinct roots with having
distinct eigenvalues of the same magnitude. For ,
. Otherwise, the sequence diverges.
Proof.
Every is conjugate to its Jordan normal form, so the result follows if it is
so for in Jordan normal form.
Let .
The design of cases (a)–(d) makes evident
that exactly one of the following holds, where , are the
eigenvalues of . In cases (a) and (b) below there is one eigenvalue
of multiplicity .
(a)
is defective, is nilpotent, and
(b)
, and
(c)
, and
(d)
, , and
∎
2.2. Subgroups of \excepttoc\fortoc determined by eigenspaces
The subsets of with particular eigenvectors form subgroups of .
Proposition 2.4.
Let .
Every satisfies
if and only if
is a scalar multiple of the identity matrix .
Proof.
Write as in (1).
If , then every satisfies .
Conversely, if every satisfies , then and
are eigenvectors of and is diagonal.
Since is also an eigenvector,
must be a scalar multiple of the identity.
∎
Proposition 2.5.
Let , ,
with root . Set
.
(a)
The set is a subgroup of
conjugate to the Borel subgroup of
consisting of upper triangular matrices.
(b)
For every , divides .
If , means .
(c)
The group has a multiplicative character
mapping to its -eigenvalue.
Proof.
(a): The stabilizer of
consists of the upper triangular matrices. Since
acts transitively on , is a subgroup of
conjugate to for every .
(b): Suppose . Then
and divides .
(c): Suppose . Then and
so is a multiplicative character on .
∎
Proposition 2.6.
Let with .
The set
is a group.
When is not a square,
and is conjugate to the group of diagonal matrices in .
When is a square in , is conjugate to the
the subgroup of upper triangular matrices in with equal diagonal entries.
Proof.
When is not a square, , .
In this case,
is equivalent to having eigenvectors
and .
Hence, . A change of basis mapping to
and to conjugates to , the group of diagonal matrices.
When is a square, has one element of multiplicity ..
Every is either a scalar multiple of or defective.
A change of basis mapping to
conjugates to . A matrix in is either a scalar multiple of or
upper triangular with equal diagonal entries and not diagonal.
∎
3. Finite continued fractions
Let be a ring containing 1.
We define the set of finite continued fractions
with partial quotients in ,
and an operation
making a semigroup.
We give an equivalence relation on and show that the equivalence classes
form a group.
Let , .
Definition 3.1.
A finite continued fraction (FCF) is the formal expression
(5)
We say that the FCF has length
.
The elements are called the partial quotients.
For , the convergent
of is the evaluation of the finite continued fraction ,
where we interpret as .
We distinguish between the
formal object and its value
.
There is a simple rule to iteratively compute as a ratio . Let
Let be the identity matrix and
be a finite continued fraction as in (5).
Define by the matrix equation
Recursively define , for by
The numerators and denominators
with
can be written in terms of continuant polynomials (see, e.g.,
[BEJ21, p. 385]).
An -FCF is a FCF
having all partial quotients .
3.1. An equivalence relation on finite continued fractions
Definition 3.2.
The semigroup is the set
with the concatenation
operation :
Equivalently, is the semigroup of the
set of words in formal symbols , for , with the
two descriptions identified by
Definition 3.3.
Let be the set
modulo the equivalence relation generated by the relations
(6)
For denote by the equivalence
class containing .
For a word in the ’s denote by
the equivalence class containing .
Proposition 3.4.
(a)
If with
and ,
then
(b)
The binary operation on induces a well-defined
binary operation on , making a semigroup.
(c)
The semigroup is a group with identity .
(d)
The element has inverse
.
Proof.
(a) is immediate, and (b) is implied by (a).
(c):
Substituting into relation (6) shows that
the class of
is the identity in . The equality
exhibits the inverse .
∎
Definition 3.5.
Define for by
, ,
and .
We consider the group ,
where is viewed as an additive group and is a generator.
Let be the surjective
map of semigroups
In particular, note
that
, ,
and .
Proposition 3.6.
The morphism of semigroups
of Definition 3.5 induces an isomorphism of
groups .
Proof.
Verify that for .
Hence induces a well-defined map on equivalence
classes by the Definition 3.3
of . To see that is an isomorphism, verify that
its inverse is given by
Remark 3.7.
For , set
.
Observe that
:
This formula for also follows from
Proposition 3.4(d).
Definition 3.8.
The finite continued fraction is reduced if
it has no interior zeros:
for .
Proposition 3.9.
For , there is a unique reduced
such that
Proof.
Given , iteratively apply relation (6)
to remove interior zeros to produce a reduced .
On the other hand, suppose
are both reduced and .
Then .
Since they are reduced, both and
map under
to words in ,
each literally expressed without spurious insertions of the identity, and they are equal.
Since is a free product, these words are the same and
.
∎
Definition 3.10.
For any
, the normal form of is
the unique reduced representative
of the equivalence class .
Remark 3.11.
Two elements are equal if and only if they have
the same normal forms in :
In practice one computes for by
the algorithm of repeatedly applying relation (6)
until all interior zeros are eliminated.
For example, consider
We have
and hence with
.
3.2. Mapping finite continued fractions to matrices
Let
The kernel of the determinant map, , is normal of index 2 in .
Proposition 3.12.
The map
given by
(7)
induces a well-defined homomorphism of groups
.
We therefore have a homomorphism of groups
.
Proof.
We have specified the map on . We must show that respects
equivalence under (6).
By direct computation,
Proposition 3.13.
Suppose and .
(a)
The values
of and are equal: .
(b)
We have , so
there is a well-defined function
Proof.
(a): Since ,
by Proposition 3.12
We have that , because
and likewise for by (7).
(b): Reformation of a word by relation (6)
maintains the parity of the word length.
∎
Proposition 3.14.
Let
be
.
Let
be the group homomorpism to that maps to
and is trivial on the second factor.
For , we have
and .
Proof.
The maps and are
determined by their action on the generators
of , namely
and .
Hence and
for ,
.
∎
Definition 3.15.
Set .
Proposition 3.16.
Let
be as in Proposition 3.14. There is an isomorphism
given by sending in the first factor to
and in the second factor to .
Hence maps isomorphically onto .
Proof.
The definition of the free product implies that is
injective.
A word in with an even number of ’s is of the form
with possibly , , or both 0,
which is in the image of .
∎
Let for be as in
Definition 3.5.
We have in the first factor of
and in the second factor.
In particular generates
.
Proof.
The fact that the equivalence relation does not change the parity of the number
of ’s implies (a).
(b) follows from Proposition 3.16 after unwinding
the definitions.
∎
The group is the subgroup of generated by elementary matrices in .
(d)
The group is the subgroup of generated by elementary matrices
in .
(e)
Set .
Proposition 3.21.
The image of
is
and .
Proof.
By Remark 3.19,
maps generators of
to generators of
and generators of to generators of .
∎
Corollary 3.22.
We have
and
Proof.
Since is generated by , the equality
is an
immeditate consequence of Proposition 3.21.
The kernel of the determinant map applied to is .
The group is the kernel of determinant composed with
by Proposition 3.14 and Definition 3.15.
Thus, the equality
follows.
∎
Remark 3.23.
Note that if and only if
.
For a number field with integers , it is known
that unless
where and is squarefree.
Vaseršteĭn [Vas72] proved that
if is not imaginary quadratic (see also [Lie81]), and Cohn
[Coh66, Thm. 6.1] had previously settled the imaginary quadratic case.
Nica [Nic11] provides a short constructive proof that for
the ring of integers of an imaginary quadratic field ,
either or
is an infinite index non-normal subgroup of .
3.3. The kernel of \excepttoc\fortoc and presentations of
\excepttoc\fortoc
The map
of Proposition 3.12
has image by
Proposition 3.21. It is a difficult
problem to find the kernel of . We begin with the
observation that is not injective.
Remark 3.24.
Let
Suppose , and there exists such that
, .
Note is an integral domain since , so
implies .
We also have because
.
This is sufficient to compute that
Hence, for example we have the following nontrivial elements in the kernel of
:
(8)
(9)
(10)
provided , respectively.
A ring has a norm given by the usual absolute value.
Definition 3.25.
A ring is discretely normed [Coh66, p. 16]
when every unit has norm 1, and
every nonzero, nonunit has norm at least .
An imaginary quadratic number ring is discretely normed unless it has
discriminant
The exceptions arise from elements such that when
or when , namely
up to signs.
These all give nontrivial kernel elements when substituted into
expressions (8) and (9), respectively,
such that all partial quotients are nonzero nonunits.
If is any discretely normed ring
and ,
then some is a unit or zero as shown by
a modification of
[Coh66, Lemma 5.1].
In parallel with [Coh66, p. 27], we define the following subgroup of
Definition 3.26.
(a)
For , let
(b)
Let be the normal closure in of the group generated by
(11)
(12)
(13)
Calculating that
in the notation of Definition 2.2,
it is routinely verified that (11) – (13)
map to under and hence
is a normal subgroup of .
Relations (11) and (12)
give
the same commutative group structure as
.
In particular,
Since , induction
shows that for all
, establishing (16).
We have since each of its generators
in (15) has even length, satisfying Definition 3.15.
Since is a normal subgroup of index with the
nontrivial coset of containing
, to show
that the normal closure of
in
is equal to its normal closure in , it suffices
to observe that
Theorem 3.30(Cohn).
The map is surjective with kernel
. Likewise
is surjective with kernel .
Proof.
Cohn [Coh66, Thm. 5.2] shows that discretely normed rings are universal for
as in Definition 3.27, and is discretely normed.
The statement for
implies that for .
Of course, it follows from the Euclidean algorithm that ; see, for example,[Art91, Chap. 12, Sect. 4].
∎
Knowing the kernel of gives a presentation of ,
and a presentation of if
is surjective. More generally, information
on gives information on the generation of
if .
It is an interesting question whether information on ,
such as knowing , gives
results on bounded generation in case is infinite
as in [MRS18] and [JZ19].
We show that the presentation of given by Theorem 3.30 gives
the familiar amalgamated product presentation.
We now consider the group . A presentation of
this group is the free group on two generators
and modulo the obvious relations:
(18)
where is the normal closure
of
in .
We now state and prove a series of claims:
Claim 1.Set , the normal closure
of in .
Then is the normal closure of
in . So by Proposition 3.29,
.
Proof.
:
Using (17), verify that .
Hence by Theorem 3.30.
: It suffices to note the following:
proving Claim 1.
Claim 2.The elements and generate
.
Proof.
By Proposition 3.18, is a free group
on the two generators and .
But
proving Claim 2.
We can now conclude the proof of Corollary 3.31. By Theorem
3.30 we have
When is a discretely normed quadratic imaginary number ring,
Cohn shows
and, in [Coh66, Thm. 6.1],
is a proper containment.
4. Periodic continued fractions
An infinite continued fraction is a formal expression
where the sequence of partial quotients does not terminate.
The infinite continued fraction may
also be expressed as a nonterminating version of (5).
As in the finite case, the convergent is the evaluation of
For as in (7),
set
with so that .
The value (or limit) of is
when the limit exists in , in which case
converges.
A periodic continued fraction (PCF) is an
infinite continued fraction together with a type
with , such that for .
We denote the sequence of partial quotients of the PCF by
(19)
The natural number is the period of ; the
initial part of in (19)
is the FCF ; and the repeated part
of is the FCF .
The PCF as in (19) has Galois dual
(20)
see, e.g., [BEJ21, p. 384].
For a ring containing 1,
say
is an - if
for , .
Let denote the set of all -PCFs. If , then
.
A PCF with type is purely periodic.
For brevity, we abbreviate purely periodic continued fraction as RCF,
after the phrase repeating continued fraction.
The set of RCFs which are -PCFs is denoted .
An untyped periodic continued fraction (UCF)
is an infinite continued fraction which is periodic of some type .
Notice that the type of a UCF is not uniquely determined:
A UCF with type also has type for every
and every multiple of .
Definition 4.1.
There are three equality relationships between a PCF of type
and a PCF of type .
(a)
and are CF-equal
(written )
when and are the same as UCFs.
(b)
and are k-equal (written ) when and are the same as UCFs
and .
(c)
and are equal (written ) when and are the same as PCFs, that is, and .
For example, ,
but
.
And , but
.
4.1. An equivalence relation and group law on PCFs
We give an equivalence relation on and a group law
on the equivalence classes by means of the group law on .
Definition 4.2.
The map maps the
PCF with initial part and repeating part to
(21)
where is as in Definition 3.5 and
is as in Remark 3.7.
The type is essential data attached to for computing
the map .
We also have maps
and
defined by
(22)
(23)
Note that
is the identity on and
the map factors as
(24)
The maps and
give inverse bijections:
(25)
The equivalence relation on
defined in Section 3.1
induces an equivalence relation, also denoted .
on .
The proposition then follows from this by iteration.
∎
Definition 4.6.
An RCF
is reduced when
is reduced as in Definition 3.8, i.e.,
for .
Remark 4.7.
Note that is reduced if and only
if is reduced since .
Proposition 4.8.
For ,
let be as in Proposition
3.9. Then
the element is
the unique reduced element of such that
Proof.
For we have that
is reduced, as stated in Remark 4.7. We claim that
. Note that
implies that
as in Remark
4.4. By Definition 4.3, implies
. Thus,
.
Suppose are both reduced and .
In this case, are both reduced and
by Definition 4.3. Proposition 3.9 shows
that and hence by (25).
∎
Write .
For ,
let be the
equivalence class containing . By Definition 4.3 and
Remark 4.4,
the maps
and induce inverse bijections
Via the bijections ,
the binary operation of
Proposition 3.4 on induces a binary
operation, also denoted , on :
for we have
(27)
Theorem 4.9.
The semigroup
with its binary operation
is a group isomorphic to .
The map
of Definition 4.2 induces an
isomorphism of groups .
For the convenience of the reader we give in Figure 1 the maps defined
thus far and the relations between them.
Figure 1. The commutative diagrams formed by the maps in Sections
3
and 4.1. The maps and can
be added to the diagrams.
Remark 4.10.
If with , then
Proposition 4.8
shows that the equivalence class
contains a unique reduced RCF . We call the
the normal form of . Two elements
are equal if and only if they have
the same normal forms:
For , there are PCFs with not all partial quotients in
with normal form ; most simply,
for any .
We test the equivalence of via their
normal forms .
It is never necessary to make an excursion outside to determine
equivalence of PCFs with partial quotients in .
We give an example of how to compute using
Proposition 4.8 and Remark 3.11.
Suppose .
Then
We have
Hence and
with
In practice, is computed
using the normal forms ,:
if and
, then
Remark 4.11.
Suppose
are of types and respectively. Then the product takes
a particularly simple form with a representative of type :
(28)
A similar exercise shows that the product is simple for
with .
Suppose
Then
(29)
Remark 4.12.
For every PCF of the form ,
is an involution, but, in
general, is not equivalent to .
Every PCF of the form is in the class of the identity in
, that is, .
5. The subgroup and its representations
For , [BEJ21, Defn. 2.4]
defines a matrix which plays a large role
in the study of . We review the definition, using
the notation of this paper.
We denote the quadratic polynomial
as in (3) by
and the multiset of its roots by .
Proposition 5.2.
The map induces a homomorphism with image
as in Section 3.
Proof.
From Definition 4.3 and Proposition 3.12
we see that is well-defined on equivalence classes in .
By Proposition 3.21, the image of is .
∎
Definition 5.3.
Let Set
Recall from Proposition 2.5
that is the subgroup
and has a multiplicative character
mapping to its -eigenvalue.
The preimage of under is ,
so is a subgroup of . The
multiplicative character maps an element
to the -eigenvalue of .
Definition 5.4.
(a)
For with set
where is the subgroup of defined in
Proposition 2.6.
(b)
For , set
(c)
Recall from Proposition 5.2
that factors through :
For , set
For with , the preimage of under is ,
so is a group. For , the group
has a multiplicative character
inherited from the containment .
Theorem 5.5.
Let .
(a)
If , ,
and has corresponding eigenvectors , then
and the values of the characters and
on are
quadratic integral over and in .
The product of these characters is .
(b)
If has one element of multiplicity ,
has a character
whose square is .
(c)
Otherwise, is a scalar multiple of the identity and has a
character
equal to
for every Its square is .
Proof.
Let and .
(a):
If is not zero and not a square, then
has distinct elements and has corresponding eigenvalues
that solve the monic quadratic equation .
It follows from Propostion 2.6 that
Relation (4) shows
the eigenvalue .
The sum is in , so .
The product is , so both
and are units.
(b):
When is a nonzero square, has a single eigenvector .
For as above,
the eigenvalue solves a monic linear equation over and
.
(c):
When is zero, every is , a scalar
multiple of . In this case, for every .
∎
Remark 5.6.
Let be a PCF of type .
The convergent
relates and the eigenvalues of .
When
(30)
we have .
For ,
we have
Thus, is the th convergent .
If has two distinct elements and has associated eigenvalues
,
If has one element of multiplicity two, is the eigenvalue of . If is a scalar multiple of , its eigenvalue is .
Remark 5.7.
The maps defined in this section form a commutative diagram as given in
Figure 2.
Figure 2. The commutative diagram formed by the maps in Section
5
for with and .
6. Examples of Theorem 5.5 arising from number rings
6.1. \excepttoc\fortoc
The -PCF is of type and
so , , and the first convergent
.
We have , with
as in Definition 5.4(c).
The characters
We have
From Example 6.1 and from
Example 6.3 .
We verify using (32) and (33)
6.5. \excepttoc\fortoc
The -PCF from [BEJ21, Cor. 8.19]
is of type . We have
,
and the convergent
,
which is a horrible approximation to in keeping with
[BEJ21, p. 415], which remarks on and quantifies the
extremely slow convergence of .
The values of the characters
on the class
are
which are both units in the ring .
In fact
7. Refined notions of convergence for periodic continued fractions
For a PCF , the multiset contains the limit
of when that limit exists; see, e.g.,
[BEJ21, Appendix].
Theorem 7.1.
Let
and
be PCFs of types and that are -equal.
(a)
If so that and are -equal, then
, ,
and .
(b)
If , then , and
and are linearly dependent.
If and are both nonzero, then
.
(c)
The PCF converges if and only if converges and in this case
their limits are equal: .
Proof.
(a): Suppose . If , then equals .
In the alternative , we may assume with .
In this case,
where we take the subscripts of the ’s to put them in the range 1 to
and for .
So
using the fact that and times.
The rest of (a) follows trivially.
(b): Since the choice of does not affect ,
we may assume .
The equality follows from the observation that the concatenation of
copies of equals the concatenation of copies of
.
The Cayley–Hamilton theorem implies that any power of a
matrix is a linear combination of and .
This implies , , and are
linearly dependent, which implies and are linearly dependent,
which implies they have the same roots if both are nonzero.
For the PCF
of type , we introduce notation
for the matrix and its entries:
and for the limits, if they exist,
(34)
Definition 7.2.
The matrix is heavy when
Rephrased with this terminology, the condition
[BEJ21, Thm. 4.3(b), INEQ 4.1] says that is heavy for some
.
The matrix is . The matrices are all conjugate to .
From the definition, we see .
Proposition 7.3.
Let be a PCF. Suppose is heavy.
(a)
The limit exists
and is an -eigenvector of .
(b)
is not heavy.
Proof.
(a):
The heavy condition implies has an eigenvector
with eigenvector . Therefore,
has eigenvector .
We compute the limit
We are now done because and are both
elements of the equivalence class .
(b): The calculation, with argument suppressed and , for the value of
such that ,
shows that and cannot both be heavy.
If they are, , yielding , a contradiction.
∎
We recall a theorem proved in the appendix of [BEJ21], with the cases aligned to
match Proposition 2.3.
Theorem 7.4.
([BEJ21, Thm. 4.3])
Let be a PCF, let be the eigenvalues of chosen
so that , and let be the
corresponding eigenvectors.
If converges, its limit .
Exactly one of the following holds.
(a)
has one root of multiplicity ,
and
(b)
, , and diverges.
(c)
has roots with .
(i)
For some , is heavy, ,
,
and diverges.
(ii)
For all , is not heavy, and converges to .
(d)
has distinct roots with ,
and diverges. In fact, does not exist for some .
Proof.
In case (a), by Proposition 2.3(a),
for all , , so converges to .
In case (b),
for all so
.
No two consecutive ones are equal: by the
more general
rule that no two consecutive convergents can be equal, so diverges.
Now assume that cases (a) and (b) do not hold.
has two distinct roots, , . The matrix
has corresponding distinct eigenvalues
, .
In case (c), and
Proposition
2.3(c) shows all exist.
Suppose, as in the first subcase, there exists such that is heavy,
Proposition 7.3 shows and is not heavy.
Proposition 2.3(c)
shows so diverges.
In the alternative subcase of (c), where there
is no such that is heavy, we have for all ,
and
converges to .
In case (d) of divergence, we have
and . By Proposition 2.3(d),
the limit exists if and only if .
We claim there exists such that .
Assume, to the contrary, that for all .
To deduce a contradiction, we show implies .
From the equality
(35)
we see that implies .
Similarly, implies , so and are
diagonal with distinct diagonal entries .
Equation (35) implies
.
Since for all , we have that for all .
If is odd then for every j, a contradiction.
If is even then we are in case (b), also a contradiction,
thus proving the claim.
Hence, for some , does not exist, and diverges.
∎
We name the convergence behaviors of a PCF identified in Theorem 7.4.
Definition 7.5.
In cases 7.4(a) and
7.4(c), is quasiconvergent.
In cases 7.4(b) and
7.4(d), is strictly divergent.
In the divergent subcase of 7.4(ci),
is strictly quasiconvergent.
Divergent means strictly divergent or strictly quasiconvergent.
Quasiconvergent means convergent or strictly quasiconvergent.
Theorem 7.6.
Let PCFs and be equivalent.
is quasiconvergent if and only if is quasiconvergent.
If and both converge, then
.
Proof.
PCFs and are quasiconvergent together,
because .
For the same reason, if and both converge, they converge to the same value.
∎
The following lemma and theorem characterize strict quasiconvergence.
Lemma 7.7.
Let be a PCF. If
and are both heavy, then for the value of such that
.
Proof.
Inspect the relation, with argument suppressed, , and :
Observe that implies
If and , then , contradicting
.
Hence, implies
∎
Theorem 7.8.
The PCF is strictly quasiconvergent if and only if
for each , ,
exists, a strict majority, but not all, of which are the same.
Proof.
Let the eigenvalues of be and , with .
If both eigenvalues have magnitude , arbitrarily assign one to be .
Define and such that ,
are eigenvectors of with corresponding eigenvalues and .
If is strictly quasiconvergent, then there exists such
that is heavy and
Proposition 7.3 shows that
for every such that is heavy,
and is not heavy.
For those such that is not heavy,
If is heavy for every even or every odd ,
then either is even and, by Lemma 7.7, every second
partial quotient of is zero, or is odd, and all partial quotients of
are zero.
In the even case, is conjugate to an elementary matrix and satisfies
7.4(a) or 7.4(b), a contradiction.
In the odd case, is conjugate to the matrix
of Remark 3.19 and satisfies 7.4(d).
Hence, for only a minority of , ,
can be heavy and .
Conversely, suppose
for each , ,
exists, a strict majority, but not all, of which are the same.
If satisfies Theorem 7.4(b), then
.
So a majority of the , , are not equal to a
particular value.
The PCF
satisfies Theorem 7.4(d) if and only if
has eigenvalues with .
We claim that for , either does not exist, or
exists and
is a constant independent of .
We shall prove this for . The result follows for all by rotating the
periodic part of .
Write
and observe
If exists, then is an eigenvector of ,
so either or .
This proves the claim.
By assumption that the limits exist and the fact that consecutive
convergents cannot be equal, we cannot have a strict majority of agree and
satisfy Theorem 7.4(d)
The only possibility left is that satisfies
Theorem 7.4(ci),
so is strictly quasiconvergent.
∎
The strict majority hypothesis in the sufficiency part of the proof of
Theorem 7.8 is essential.
For example, has limits
and . The split is equal and
is not quasiconvergent.
It is strictly divergent because it satisfies Theorem 7.4(b).
8.2. The fraction of \excepttoc\fortoc can be made arbitrarily close to \excepttoc\fortoc
We can make the fraction of arbitrarily close to .
Let and
have odd length period .
We have and , .
If we are in case (b) of Theorem 7.4,
and .
If , , we are in case (d), and
does not exist.
If we are in case (ci) and
so has a slim majority.
A similar construction can be made for even periods by taking
where the length of the strings of 0’s have the same parity.
8.3. The fraction of \excepttoc\fortoc can be made arbitrarily close to \excepttoc\fortoc
Likewise we can make the fraction of arbitrarily
close to . Let be the
Fibonacci sequence. Let and
where there are ’s.
The continued fraction for provides an example where exactly
of the ’s are .
(If , doesn’t exist, and if , .)
If , then we would have from (8.3).
But this is
not possible for unless :
takes care of and if ,
then since .
Hence if , ,
so in this case
by Proposition 2.3(c),
with the “” there equal to and the “” there
equal to .
For the exceptional case , explicit computation of using
the Cassini/Vajda identities for Fibonacci numbers gives
(40)
Hence by (40) and
since .
Hence again by
Proposition 2.3(c),
with the “” there equal to and the “” there
equal to .
Lastly we show the computation giving and :
(41)
Calculating and entails finding
the eigenvectors of with as in (37) and
as in (8.3).
Let
Observe that since .
Using the Cassini identity yet again, verify that
So will have eigenvectors ,
with eigenvalues , , respectively.
We have
8.4. The equivalence
\excepttoc\fortoc
on
\excepttoc\fortoc
does not respect convergence
Theorem 7.6 shows quasiconvergence is a property of
PCF equivalence class.
It is possible for , both quasiconvergent,
with strictly quasiconvergent (and hence divergent)
and convergent.
For example, take and let , .
Let the convergents of be
, .
If , then satisfies Theorem 7.4(ci),
and hence is strictly quasiconvergent.
The limits for are
The PCF converges to , since the
convergents are just .
[BEJ21]
Bradley W. Brock, Noam D. Elkies, and Bruce W. Jordan.
Periodic continued fractions over -integers in number fields and
Skolem’s -adic method.
Acta Arith., 197(4):379–420, 2021.
[Coh66]
P. M. Cohn.
On the structure of the of a ring.
Inst. Hautes Études Sci. Publ. Math., (30):5–53, 1966.
[JZ19]
Bruce W. Jordan and Yevgeny Zaytman.
On the bounded generation of arithmetic .
Proc. Natl. Acad. Sci. USA, 116(38):18880–18882, 2019.
[Knu97]
Donald E. Knuth.
The art of computer programming. Vol. 1.
Addison-Wesley, Reading, MA, 1997.
Fundamental algorithms, Third edition [of MR0286317].
[Lie81]
Bernhard Liehl.
On the group over orders of arithmetic type.
J. Reine Angew. Math., 323:153–171, 1981.
[MRS18]
Aleksander V. Morgan, Andrei S. Rapinchuk, and Balasubramanian Sury.
Bounded generation of over rings of -integers with
infinitely many units.
Algebra Number Theory, 12(8):1949–1974, 2018.
[Nic11]
Bogdan Nica.
The unreasonable slightness of over imaginary
quadratic rings.
Amer. Math. Monthly, 118(5):455–462, 2011.
[Vas72]
L. N. Vaseršteĭn.
The group over Dedekind rings of arithmetic type.
Mat. Sb. (N.S.), 89(131):313–322, 351, 1972.