Two-variable polynomials with dynamical Mahler measure zero
Abstract.
We discuss several aspects of the dynamical Mahler measure for multivariate polynomials. We prove a weak dynamical version of Boyd–Lawton formula and we characterize the polynomials with integer coefficients having dynamical Mahler measure zero both for the case of one variable (Kronecker’s lemma) and for the case of two variables, under the assumption that the dynamical version of Lehmer’s question is true.
Key words and phrases:
Mahler measure, dynamical Mahler measure, polynomial, preperiodic points, equidistribution2020 Mathematics Subject Classification:
Primary 11R06; Secondary 11G50, 37P15, 37P301. Introduction
This paper considers a generalized version of Mahler measure attached to a discrete dynamical system. In [ST12, PST05], the authors consider a single-variable generalized Mahler measure, which we extend here to multivariate polynomials. We begin with some background on both Mahler measure and discrete dynamical systems as well as the connections between them.
1.1. Mahler measure
The (logarithmic) Mahler measure of a non-zero rational function is defined by
where is the unit torus.
In the particular case of a single variable polynomial, Jensen’s formula implies that if
then
(1.1) |
The single variable polynomial version was first introduced by Lehmer [Leh33] as a useful tool in a method for producing large prime numbers by generalizing Mersenne’s sequences.
It is natural to ask the question about the range of possible values of Mahler measure. For , Kronecker’s lemma [Kro57] implies that if and only if is (up to a sign) monic and a product of cyclotomic polynomials and a monomial.
From the context of finding large prime numbers, Lehmer posed the following question:
Given , can we find a polynomial such that ?
Lehmer showed that
and declared that this was the polynomial with the smallest positive measure that he was able to find. To this day, Lehmer’s question remains unanswered and Lehmer’s polynomial remains the one with the smallest known positive Mahler measure. Some key progress on this question can be found, for example, in the works of Breusch [Bre51] and Smyth [Smy71], who proved that the Mahler measures of non-reciprocal polynomials have a lower bound; Dobrowolski [Dob79], who gave the first lower bound that approaches zero when the degree goes to infinity; and the recent proof of Dimitrov [Dim] of the Schinzel–Zassenhaus conjecture.
Mahler measure for multivariate polynomials was first considered by Mahler [Mah62] in connection to heights and their applications in transcendence theory. Later Smyth [Smy81, Boy81b] obtained the first exact formulas involving Dirichlet -functions and the Riemann zeta function, such as
More recently, various formulas have been proven or conjectured involving -functions associated to more complicated arithmetic-geometric objects such as elliptic curves. For example, the following formula was conjectured by Deninger [Den97] and Boyd [Boy98] and proven by Rogers and Zudilin [RZ14]
where is an elliptic curve of conductor 15.
We finish the discussion on classical Mahler measure by recalling a result due to Boyd [Boy81b] and Lawton [Law83]. If , then
(1.2) |
where
where . Intituively, the above limit is taken while go to infinity independently from each other.
This result expresses the Mahler measure of a multivariate polynomial in terms of the Mahler measure of a single variable polynomial. It sheds light into the Lehmer conjecture for multivariate polynomials, since the existence of any such polynomial with small Mahler measure would immediately imply the existence of infinitely many single variable polynomials with small Mahler measure.
1.2. Arithmetic dynamics
In classical (holomorphic) dynamics, one studies topological and analytic properties of orbits of points under iteration of self-maps of or, more generally, of a complex manifold. In the younger field of arithmetic dynamics, the self-maps act on number theoretic objects (for example, algebraic varieties) and the objects and maps are defined over fields of number theoretic interest (number fields, function fields, finite fields, non-archimedean fields, etc.).
We fix the following notation: Let be a set, possibly with additional structure. For and , we write
(1.3) |
This conjugation is a natural dynamical equivalence relation in that it respects iteration; that is, . In the sequel, we are concerned with polynomial maps , so we consider equivalence up to conjugation by affine maps with .
Let . The forward orbit of under is the set . Questions in arithmetic dynamics are often motivated by an analogy between arithmetic geometry and dynamical systems in which, for example, rational and integral points on varieties correspond to rational and integral points in orbits, and torsion points on abelian varieties correspond to preperiodic points (points with finite orbit). See [Sil07] for more comprehensive background and motivation and [BIJ+19] for a survey of recent progress in the field of arithmetic dynamics.
Two complementary sets play an important role in holomorphic dynamics: The Fatou set of is the maximal open set on which the family of iterates is equicontinuous (considered as maps on ); its complement is the Julia set of , which we denote . Informally, the Julia set is where the interesting dynamics happens. For a polynomial , we may define the Julia set a bit more concretely.
Let . The filled Julia set of is
The Julia set of is the boundary of the filled Julia set. That is, . It follows from these definitions that for a polynomial , both and are compact. In dynamical Mahler measure, the Julia set for a general polynomial will play the role of the unit torus .
Two families of polynomials play an important role in this work: The pure power maps and the Chebyshev polynomials. Both of these families have special dynamical properties that arise because they are related to endomorphisms of algebraic groups.
Consider the multiplicative group where for a field , the -valued points are . The endomorphism ring of is :
Iteration of pure power maps is particularly easy to understand:
For and we have three cases: If , then with . If , then with . If , then for all . So for pure power maps, we can understand the Julia sets completely: is the unit disc, and is the unit circle.
To give another example, since the automorphism of given by commutes with the power map , the power map descends to an endomorphism of , which we denote , the Chebyshev polynomial.111Classically the Chebyshev polynomials are normalized so that . The two normalizations satisfy .
Taking as a definition the fact that satisfies
(1.4) |
one may prove existence and uniqueness of the Chebyshev polynomials along with a recursive formula:
(1.5) |
The following rule for composition of Chebyshev polynomials follows directly from the definition in (1.4):
and it gives a simple form of iteration
(1.6) |
For , we have two cases for : If , then . If , then with . Thus we understand the Julia set in this case as well: .
See [Sil07, Chapter 6] for more on the dynamics of pure power maps, Chebyshev polynomials, and other maps arising from algebraic groups, including proofs of some of the statements above.
1.3. Dynamical Mahler measure and statement of results
An important tool in the study of arithmetic dynamics is the canonical height of a point in relative to a morphism , defined by Call and Silverman [CS93] and modeled after Tate’s construction of the canonical (Néron-Tate) height on abelian varieties. The standard (Weil) height of a point (written in lowest terms) can be computed as , and, observing that and are (up to sign) the coefficients of the minimal polynomial of over , this definition extends easily to algebraic numbers . If is a polynomial and , we then define:
where represents the -fold composition of and . In analogy with the canonical height on abelian varieties, we find that this limit exists and that
-
•
, and
-
•
if and only if is a preperiodic point for (that is, for some natural numbers ).
Let with minimal polynomial . The theory of heights and Mahler measure are connected via the formula
(1.7) |
Let be compact. For any Borel probability measure on the energy of is given by [Ran95, Definition 3.2.1]:
An equilibrium measure on is a measure satisfying
where the supremum is taken over all Borel probability measures. An equilibrium measure always exists and is often unique [Ran95, Theorem 3.2.2].
Let
where and .
Brolin [Bro65], Lyubich [Lyu83], and Freire, Lopes, and Mañe [FLM83] construct a unique equilibrium measure , supported on the Julia set , which is invariant under . That is, , where the push-forward is defined by for any Borel set . We also have
(1.8) |
and in particular, when is monic.
Suppose that is monic, and that (written in lowest terms). Then one can show that
More generally, if is an algebraic integer with monic minimal polynomial , then
(1.9) |
Notice that we are associating the canonical height for the rational map to a generalized Mahler measure for polynomials . Taking recovers the familiar Mahler measure from equation (1.7).
Piñeiro, Szpiro, and Tucker [PST05] have shown that, like heights, this generalized Mahler measure may be computed via an adelic formula:
(1.10) |
where is a number field, , is a minimal polynomial for over , and denotes the Julia set of . For finite , unless has bad reduction at or all the coefficients of have nonzero -adic valuation. Thus, equation (1.9) follows from (1.10) by specializing to and algebraic integer.
The contributions at the finite places in equation (1.10) were shown to have the appropriate integral forms by Favre and Rivera-Letelier [FRL06, Proposition 1.3]. Thus, equation (1.9) can also be extended to the case where is not necessarily monic if the integral is replaced by a sum of integrals at different places.
Szpiro and Tucker [ST12] used diophantine approximation to show that this generalized Mahler measure of a polynomial at a place can be computed by averaging over the periodic points of .
It is not easy to give precise examples of . In the classical case when with , then for . For the Chebyshev polynomial with , normalized so that , one can show that . Because of the underlying group structure discussed above, these examples are straightforward to compute (see [CLMM]). For general polynomials, exact computation is much more difficult, and in fact these two families are the exceptions to many theorems stated in the sequel.
One can compute approximations to dynamical Mahler measure in some cases. For example, Ingram [Ing13] gave an asymptotic formula for for as
Similarly for , one can find
It is natural to ask about higher-dimensional (multivariate) forms of this generalized Mahler measure.
Definition 1.1.
Let monic of degree . The -dynamical Mahler measure of is given by
(1.11) |
We will later see that the above integral always converges and that in fact, if is nonzero.
It is an interesting to ask whether multivariable dynamical Mahler measure can also be interpreted as a height. We do not answer the question in this paper, but we suspect that the answer is yes. Zhang [Zha95], in the more general setting of a variety with a map and a line bundle such that , defined a canonical height of an arbitrary subvariety . Chambert-Loir and Thuillier [CLT09] showed that in the case where , and , that, for a homogeneous polynomial and the hypersurface , is the usual multivariable Mahler measure of , which is also equal to the multivariable Mahler measure of the inhomogeneous polynomial .
We conjecture that the multivariable dynamical Mahler measure can be recovered in a similar way, using instead the map , which acts as on each of the components.
In this note, we prove the following results.
Lemma 1.2 (Dynamical Kronecker’s Lemma).
Let be monic of degree and let . If , then where each is a preperiodic point of .
Proposition 1.3 (Weak Dynamical Boyd–Lawton).
Let monic of degree and let . Then
The above statement is expected to be true with an equality. This would correspond to a dynamical version of the Boyd–Lawton limit (1.2).
For the main result, we will assume the following conjecture.
Conjecture 1.4 (Dynamical Lehmer’s conjecture).
[Sil07, Conjecture 3.25] There is some such that any single variable polynomial with satisfies .
Theorem 1.5.
Assume the Dynamical Lehmer’s conjecture.
Let be a monic polynomial of degree which is not conjugate to or to , where is the -th Chebyshev polynomial. Then any polynomial which is irreducible in (but not necessarily irreducible in ) with , and which contains both variables and , divides a product of complex polynomials of the following form:
where are integers, is a linear polynomial commuting with an iterate of , and is a nonlinear polynomial of minimal degree commuting with an iterate of (with possibly different choices of , , , and for each factor).
As a partial converse, suppose there exists a product of complex polynomials such that
-
(1)
each has the form , where and are as above (with possibly different choices of , , , and for each );
-
(2)
; and
-
(3)
divides in .
Then .
Note that the second half of Theorem 1.5 is only a partial converse. We can guarantee that by enlarging the set to contain all the Galois conjugates of all its members. However, as we do not know whether the polynomials have algebraic integer coefficients, we cannot guarantee . We conjecture however that it is always the case that the have algebraic integer coefficients, in which case we could strengthen this converse.
We remark that Theorem 1.5 complements a statement for classical Mahler measure. Indeed, it is shown in [EW99, Theorem 3.10] that for any primitive polynomial , is zero if and only if is a monomial times a product of cyclotomic polynomials evaluated on monomials. This case is fundamentally different from what is considered in Theorem 1.5, as the classical Mahler measure contains more structure given by the multiplicative nature of the function .
It is also interesting to compare our results with [Zha95, Conjecture 2.5], which says that a subvariety whose dynamical canonical height is must be preperiodic. If Zhang’s height agrees with the dynamical Mahler measure, as we expect to be the case, then Lemma 1.2 and Theorem 1.5 can be seen as special cases of Zhang’s Conjecture 2.5 for the dynamical maps and , respectively. (Note that Zhang’s conjecture is known not to hold in general: see [GTZ11] for a counterexample. However, it may be possible to use the arguments in [GNY19] to obtain a proof for the case of .)
This article is organized as follows. Section 2 covers some background on complex dynamics. Section 3 is devoted to showing that the dynamical Mahler measure is well-defined in general, and non-negative for polynomials with integral coefficients, while Section 4 discusses the dynamical version of Kronecker’s Lemma 1.2. The Weak Dynamical Boyd–Lawton Proposition 1.3 is discussed and proven in Section 5. It serves as a preparation for the proof of our main result, Theorem 1.5, covered in Sections 6 and 7. Finally, in Section 8, we give a precise description of the polynomials and that appear in the statement of this theorem.
Acknowledgements
We are grateful to Patrick Ingram for proposing that we study the dynamical Mahler measure of multivariate polynomials and for many early discussions, and to the anonymous referees for their careful reading of the article and several helpful suggestions. This project was initiated as part of the BIRS workshop “Women in Numbers 5”, held virtually in 2020. We thank the workshop organizers, Alina Bucur, Wei Ho, and Renate Scheidler for their leadership and encouragement that extended for the whole duration of this project. This work has been partially supported by the Natural Sciences and Engineering Research Council of Canada (Discovery Grant 355412-2013 to ML), the Fonds de recherche du Québec - Nature et technologies (Projets de recherche en équipe 256442 and 300951 to ML), the Simons Foundation (grant number 359721 to MM), and the National Science Foundation (grant DMS-1844206 supporting AC and grant DMS-1902772 to LM). This material is based upon work supported by and while the third author served at the National Science Foundation. Any opinion, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation.
Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.
2. Some background on complex dynamics
In this section we recall some results from complex dynamics that will be useful for the rest of this article. The main reference for this is the book of Ransford [Ran95].
Let be a finite Borel measure on with compact support . One can define the potential function [Ran95, Definition 3.1.1]
by
(2.1) |
The potential is subharmonic in by [Ran95, Theorem 3.1.2].
Thus we can write the energy as
Then Frostman’s Theorem [Ran95, Theorem 3.3.4] implies that if is the equilibrium measure of , then
-
(1)
on ;
-
(2)
on , where .
By (1.8), and is a finite number.
Let a proper sub-domain of the Riemann sphere. A Green’s function [Ran95, Definition 4.4.1] for is a map such that for each :
-
(1)
is harmonic on , and bounded outside each neighbourhood of ;
-
(2)
, and as ,
-
(3)
as for outside of a Borel polar subset of .
A set is called polar if for every finite Borel measure for which is a compact subset of [Ran95, Definition 3.2.2]. Otherwise is called non-polar. If is a domain of the Riemann sphere such that is non-polar, then, there is a unique Green’s function for [Ran95, Theorem 4.4.2].
In particular, the proof of [Ran95, Theorem 4.4.2] gives us explicitly that, if is non-polar and if is the equilibrium measure on , then
(2.2) |
for .
If is the filled Julia set, we remark that [Ran95, Corollary 6.5.4] implies that the Green’s function of its complement is given by
where when and 0 otherwise.
From now on, we will write for the Green’s function of the complement of the filled Julia set.
We can summarize the results of the discussion above in the following two propositions:
Proposition 2.1.
If monic of degree , then
(2.3) |
for .
Proof.
Proposition 2.2.
If is monic of degree , then for all , and the equality holds exactly when .
3. Convergence and positivity of the dynamical Mahler measure
We will prove that the dynamical Mahler measure defined by (1.11) is well-defined. Moreover, it is non-negative when is nonzero. We will follow a structure of a proof that is similar to that of [EW99, Lemma 3.7].
We start first with the following result, which can be considered an analogue to Jensen’s formula in the dynamical universe.
Lemma 3.1.
Let monic of degree . If , then
Proof.
Now we proceed to prove the main result of this section.
Proposition 3.2.
Let monic of degree and let . Then the integral in (1.11) defining the -dynamical Mahler measure converges. Moreover, if is nonzero, then .
Proof.
First assume that is nonzero. Let , and write
Then for , we have
From this, we see that .
For the rest of the statement, we proceed by induction. The case is a consequence of what we have just discussed and Lemma 3.1, which implies that when is nonzero and when is nonzero. Assume that the result is proven for all polynomials in variables. We write
for certain algebraic functions . Then we have that
Here our final integrand is upper semicontinuous, since is upper semicontinuous and the multiset of values depends continuously on (even though the individual are not themselves continuous at the branch locus). By Proposition 2.2, . By induction, . Therefore the above integral is non-negative and the first paragraph shows that the entire right-hand side is bounded above. Therefore the integral defining the Mahler measure exists.
Now assume that . We write , with nonzero and apply the above to conclude that are well-defined. Then we have and is well-defined too. This concludes the proof of the statement. ∎
4. Dynamical Kronecker’s Lemma
The goal of this section is to characterize the single variable polynomials with integral coefficients having , which is the dynamical version of Kronecker’s Lemma:
Lemma 4.1 (Kronecker, [Kro57]).
Let . If all of the roots of satisfy , then the are zero or roots of unity.
In fact, more is true. Equation (1.1) implies the following statement.
Corollary 4.2.
Let . If , then the roots of are either zero or roots of unity. Conversely, if is primitive and its roots either zero or roots of unity, then .
An immediate consequence of Lemma 3.1 is the following.
Lemma 4.3.
Let be monic of degree and . Then we have if and only if and all roots lie in .
Proof.
Because , we have , and by Proposition 2.2, all the other summands in the dynamical Jensen formula from Lemma 3.1 are also nonnegative. Therefore, if and only if and for all . The first condition holds exactly when . The second condition is an immediate consequence of the second statement in Proposition 2.2. ∎
We are now ready to prove the Dynamical Kronecker’s Lemma (Lemma 1.2).
Proof of Lemma 1.2.
Since Lemma 4.3 implies that is (up to a sign) monic, we can write
Consider the polynomials
The coefficients of are symmetric functions in the algebraic integers and thus symmetric functions of the , so they are elements of (all the conjugates of each are present as roots of , since the coefficients are rational). Since , Lemma 4.3 implies that , and the same is true for . Since is compact, the are uniformly bounded, and the same is true for the coefficients of . Thus, the set must be finite. In other words, there are for which
That means,
Thus, there is a permutation such that
If has order , we get,
which shows that each is preperiodic. ∎
Remark 4.4.
A more general approach to dynamical Mahler measure, for example as suggested in [FRL06], may allow to be defined for a non-monic polynomial or indeed a rational function. The following argument shows that Lemma 1.2 continues to hold in this case: Noting that , we see from equation (1.9) that vanishes with the -canonical height of the roots of . Over a global field, the canonical height of a rational function vanishes precisely at preperiodic points for (see, for example, [Sil07, Theorem 3.22]).
5. Weak Dynamical Boyd–Lawton Theorem
A result of Boyd [Boy81b] and Lawton [Law83] given by equation (1.2) allows one to compute, in the classical setting, the higher dimensional Mahler measure by a series of approximations by one-dimensional Mahler measures. Here we exhibit the dynamical version of a weaker result for two variables, Proposition 1.3. Our result is analogous to Lemma 2 in [Boy81a].
First we need to establish the following lemma, which is analogous to Lemma 1 in [Boy81a].
Lemma 5.1.
If is continuous, then
Proof.
By making the change of variables , we have
(5.1) |
where . We use this to expand our left hand side, interchanging the limit and integral by the bounded convergence theorem.
By [Ran95, Theorem 6.5.8], we have that the limit inside the parentheses is equal to , as desired. ∎
We are now ready to prove the Weak Dynamical Boyd–Lawton Theorem.
6. A family of polynomials having dynamical Mahler measure zero
In Section 4 we discussed necessary and sufficient conditions for a single variable polynomial with integer coefficients to have dynamical Mahler measure zero. In this section we will show that the dynamical Mahler measure of is zero independently of as long as is monic, and we will also use this fact to construct a more general class of polynomials whose dynamical Mahler measure is zero. This is the beginning of the proof of Theorem 1.5, which will be finished in Section 7.
Recall that the multivariate dynamical Mahler measure
is precisely the energy of the equilibrium measure. By (1.8), when is monic, , and therefore .
Notice that the strong Boyd–Lawton Conjecture, which claims that , applies in this case. Indeed, for , the polynomial has roots equal to the -periodic points of , so
where denotes the Green’s function of the filled Julia set . However if is periodic, then is since the log term is bounded. Hence , and this coincides with , which is what one would predict from the strong Boyd–Lawton Conjecture.
Now we intend to prove a more general result.
Lemma 6.1.
Let be a linear polynomial in commuting with an iterate of , and be a nonlinear polynomial in commuting with an iterate of . Then for any nonnegative integers ,
To prove Lemma 6.1, we first need a lemma on dynamical systems.
Lemma 6.2.
Let be any polynomial commuting with a power of (including the case of linear). Then sends the Julia set to itself. Furthermore, preserves the invariant measure . Explicitly, this means that for any test function , .
Proof.
Since and have the same Julia set and invariant measure, we can replace with to reduce to the case where commutes with .
Since commutes with , the measure on defined by is also invariant under and has total measure . By uniqueness of the invariant measure, we have as desired. Since is the support of , it too is preserved by . ∎
Proof of Lemma 6.1.
We apply the invariance of measure in Lemma 6.2 to both variables in this integral, and obtain
which we know to be . ∎
Remark 6.3.
As a corollary, we obtain a result proving part of Theorem 1.5.
Corollary 6.4.
Let be monic of degree , and let be nonzero. Suppose there exists a product of complex polynomials such that
-
(1)
each has the form , where is a linear polynomial commuting with an iterate of and is a nonlinear polynomial of minimal degree commuting with an iterate of (with possibly different choices of , , , and for each );
-
(2)
; and
-
(3)
divides in .
Then .
Proof.
We have
for some . Taking Mahler measures of both sides, we get:
and both summands are nonnegative (since and have integer coefficients), so we must have that . ∎
7. Two-variable Dynamical Kronecker
Our goal in this section is to finish the proof of our main result, Theorem 1.5. More precisely, in Corollary 6.4 we proved that certain polynomials have dynamical Mahler measure zero. It is now time to prove that those are the only polynomials that satisfy this condition. In other words we will prove the following result.
Theorem 7.1.
Assume the Dynamical Lehmer’s Conjecture 1.4.
Let be a monic polynomial of degree which is not conjugate to or to , where is the -th Chebyshev polynomial. Let be such that , is irreducible in , and uses both variables. Then every irreducible factor of in divides a polynomial of the following form:
where are integers, is a linear polynomial commuting with an iterate of , and is a nonlinear polynomial of minimal degree commuting with an iterate of .
In order to prove Theorem 7.1 we will need to assume the Dynamical Lehmer’s Conjecture 1.4. In addition, we will use Proposition 1.3. Finally, we will also need the following unlikely intersections result due to Ghioca, Nguyen and Ye [GNY19].
Proposition 7.2.
[GNY19, Theorem 1.5] Let be a polynomial of degree which is not conjugate to or to , and let be defined by Let be an irreducible curve defined over which projects dominantly onto both coordinates. Then contains infinitely many preperiodic points under the action of if and only if is an irreducible component of the locus of an equation of the form , where are integers, is a linear polynomial commuting with an iterate of , and is a nonlinear polynomial of minimal degree commuting with an iterate of .
The following definition will be key to our proof:
Definition 7.3.
The triple with , , and satisfies the bounded orders property (BOP) if the set of nonnegative integers
is bounded. (In the above sequence we ignore any terms where is the zero polynomial, which can happen for only finitely many .)
We say that satisfies the preperiodic bounded orders property (preperiodic BOP) if has BOP for every such that is a preperiodic point of .
The following general reduction properties follow directly from the definition. We omit their proof.
Lemma 7.4.
Let and .
-
(1)
(Translation) Let and . If has BOP, then has BOP.
-
(2)
(Iteration) For let be given by . If has BOP for , then has BOP.
-
(3)
(Divisibility) If has BOP and , then has BOP.
Now we consider some less immediate reduction properties.
Lemma 7.5 (Composition/Pushforward reduction).
Suppose that and .
-
(4)
(Composition) Let be the polynomial given by . If has BOP, then has BOP.
-
(5)
(Pushforward) For irreducible in , we define to be an irreducible polynomial in which vanishes on the curve ( is defined up to multiplication by a non-zero scalar222The fact that the pushforward is only defined up to multiplication by a scalar is not significant, since we only use this construction in the context of order of vanishing.). More generally, we can extend this definition multiplicatively: for any with irreducible factorization , define .
Then for any in , if has BOP, , also has BOP.
Proof.
(4) Composition: For and ,
Hence the sequence is bounded if is.
(5) Pushforward: Assume has BOP. Since BOP is preserved by taking products, we can without loss of generality assume that is irreducible. By the composition property, also has BOP. However, by construction, the two-variable polynomial vanishes at any with , and since is irreducible, we must have by the Nullstellensatz. Hence has BOP by divisibility. ∎
The motivation for the definition of BOP is the following result.
Proposition 7.6.
Theorem 7.1 holds under the extra hypothesis that satisfies preperiodic BOP.
Proof.
Suppose that is irreducible and . We consider the family of polynomials .
By Weak Dynamical Boyd–Lawton (Proposition 1.3), we have that . We now apply the Dynamical Lehmer’s Conjecture 1.4 to obtain that for . Then if is any root of , then, by the single variable Kronecker’s Lemma 1.2, is preperiodic for .
We now argue that the set is infinite. Indeed,
goes to as , but preperiodic BOP for says that the individual summands are bounded, so the sum must be infinite. Notice that is the union of sets , as ranges through the irreducible factors of in , and at least one of these sets must be infinite. However, because is irreducible in , the irreducible factors of form a set of Galois conjugates, and so the sets all have the same cardinality, and are all infinite. Hence any irreducible factor of has the property that the set is infinite.
We are now in a position to apply Proposition 7.2 to . Indeed, if is a root of some , then the curve passes through the point , whose coordinates are both preperiodic for . Also, the dominance condition in Proposition 7.2 follows from the fact that uses both variables and .
Thus, we conclude that each complex factor of divides a polynomial of the form . ∎
It now remains to show the following result.
Proposition 7.7.
If of degree and is such that every irreducible factor involves both variables, then satisfies preperiodic BOP.
Proof.
We must show that if and satisfy the above conditions, then for any preperiodic point of , the triple satisfies BOP. Our strategy will be to show this in the case where is a fixed point of , and so we first show that the general case can be reduced to this case.
Assume that we have proved that all such triples with a fixed point of have BOP. Take a triple with preperiodic for . Then there exists an such that , so that is a fixed point for . By hypothesis, has BOP for each (note that does not vanish on any horizontal or vertical line because does not, so all its irreducible factors involve both variables). By pushforward, has BOP for each , and by iteration does also.
We have thus reduced the argument to showing that has BOP whenever is a fixed point for . By translation, we can further assume that . Additionally, if is a root of unity of order , then for , we have that . Using the iteration reduction to replace by as necessary, we can assume that if is a root of unity, then .
So suppose that is a monic polynomial such that and is either or not a root of unity, and is such that every irreducible factor involves both variables. We show that the sequence is bounded.
We can factor , where is a nonzero polynomial in and are non-constant Puiseux series in (formal power series with possibly fractional exponents, and also possibly a finite number of negative exponents, which are obtained by dividing out by ). Then
Since is independent of , it suffices to show that each summand is bounded. Note that if has a term with a negative or non-integer exponent, since is a polynomial, this bounds , so we can restrict to the case where .
We first deal with the case of separately. In this case, as , so for sufficiently large and the sequence is bounded.
Hence we are reduced to the case , and either is not a root of unity or . We will use the following notation: if is a power series, then denotes the coefficient of in . We have that is the smallest such that . The claim states that the sequence of such is bounded as varies, which will follow from the following sublemma:
Lemma 7.8.
Let be a polynomial of degree with and , and either is not a root of unity, or . There exists some such that the set of coefficients contains any number at most once.
Proof of Lemma 7.8.
Since , and , we can write . If is not a root of unity, then
and works.
On the other hand, if , we can write , where is the first nonzero coefficient after the linear one. We can calculate
so the value works. ∎
Now, for the value of given by the above lemma, there can be at most one such that , and for all other we have . Hence the values of are bounded. This finishes the proof in the case that is a fixed point of , and so by our reductions, also the proof of Proposition 7.7. ∎
8. A detailed study of and
For a fixed nonlinear, monic polynomial of degree which is not conjugate to a power function or a Chebyshev polynomial, Theorem 1.5 characterizes the irreducible polynomials with in terms of linear polynomials and nonlinear polynomials of minimal degree commuting with an iterate of . In this section we describe the polynomials and which arise in this manner. Having integral coefficients plays no special role here, so we will work throughout with polynomials over .
For a general given polynomial , we define . In our case is monic, so . The constant is invariant under replacing by one of its iterates. Conjugating by gives a polynomial of degree whose degree- coefficient is 0. Define to be the greatest common divisor of those for which . The key result is the following, due to Boyce:
Theorem 8.1.
[Boy72, Theorem 4] If is a degree- polynomial that commutes with , then the degree- polynomials commuting with are exactly those of the form for a -th root of unity.
This will allow us to completely understand the choices for both and , modulo the calculation of , as soon as we have identified suitable seed polynomials . In the linear case, we may take . The following result, a form of Ritt’s theorem [Rit23] as it appears in Boyce [Boy72], tells us how to proceed in the nonlinear case:
Theorem 8.2.
[Jul22, Rit23] If two polynomials and commute under composition, then up to conjugation with the same linear polynomial, either both are power functions, both are Chebyshev polynomials, or both are roots of unity times iterates of the same polynomial (in which case the roots of unity themselves will commute with the latter polynomial).
Since we are assuming that is not conjugate to either or , in the nonlinear case we may then take to be a polynomial of minimal degree, some iterate of which is equal to or to times a root of unity that commutes with . Note that such a polynomial will commute with all iterates of , and will have minimal degree among all nonlinear polynomials which commute with any iterate of .
We now wish to determine, for a fixed monic polynomial , what values are taken by the quantity . It will simplify the notation to index the coefficients of in order of decreasing degree, so we denote by the degree- coefficient of ; note that and for all . Let denote the greatest common divisor of those indices for which . As for any , we have .
Although is defined in terms of the full set of coefficients of , it turns out that it is in fact completely determined by its largest-degree coefficients . We elaborate on this below.
Lemma 8.3.
For , the degree- coefficient of is equal to the degree- coefficient of . In particular,
Proof.
Notice that when computing as for , the contribution to the first coefficients comes exclusively from the term , since has no degree- term, and the highest coefficient of has degree .
Further, the first coefficients of are determined by the first coefficients of , since any contribution of the coefficient of degree or below will have at most degree .
Thus the first coefficients of are determined by the first coefficients of , from the term of the form . Proceeding by induction, we conclude that the first coefficients of are determined by the first coefficients of (i.e. all the coefficients of ), from a term of the form .
In other words, the first coefficients of are given by the first coefficients of . ∎
Lemma 8.4.
Let , as above. Then unless .
Proof.
We have unless by construction. Suppose towards induction that the statement holds for some fixed . If one multiplies monomials of degree equivalent to modulo , the product has degree equivalent to . The expression is thus a sum of monomials whose degrees are equivalent to modulo , and it follows that is also. ∎
Lemma 8.5.
Enumerate the indices for which ; note that , and since is not conjugate to a power function, there is at least one nonzero coefficient with . Let
this set always contains at least and . Then for all and for all .
Proof.
For , let , so that .
As is the smallest positive index such that , it is also the smallest positive index such that , with in this case.
Suppose that . It follows that for with . The coefficient is the sum of terms of the form
where the product runs over the entries of a -tuple of indices summing to . Since , the integer does not divide , so if for each index , then at least one index has , and the corresponding coefficient is thus zero. So the only nonzero terms have one index equal to and the others equal to 0; in other words,
Corollary 8.6.
We have .
Proof.
Proposition 8.7.
Suppose that has prime factorization . Let
Then there are exactly choices for the linear polynomial and for the nonlinear polynomial appearing in the statement of Theorem 1.5: the linear polynomials commuting with some iterate of are exactly those of the form for an -th root of unity, and the nonlinear polynomials of minimal degree commuting with some iterate of are exactly those of the form , where is a polynomial of minimal degree, some iterate of which is equal to , and is an -th root of unity.
Proof.
We have shown that . If , then , so also ; in particular, . As , is a unit in , so there exists an such that , and in this case we have . The result is now an immediate consequence of Theorem 8.1. ∎
Example 8.8.
Consider the polynomial . Then , , , , and
Thus
Thus the linear polynomials commuting with some iterate of are exactly those of the form for a third root of unity. As is not a power of any smaller integer, is not an iterate of any polynomial of smaller degree, so the nonlinear polynomials of minimal degree commuting with an iterate of are exactly those of the form for a third root of unity.
References
- [BIJ+19] Robert Benedetto, Patrick Ingram, Rafe Jones, Michelle Manes, Joseph H. Silverman, and Thomas J. Tucker, Current trends and open problems in arithmetic dynamics, Bull. Amer. Math. Soc. (N.S.) 56 (2019), no. 4, 611–685. MR 4007163
- [Boy72] William M. Boyce, On polynomials which commute with a given polynomial, Proc. Amer. Math. Soc. 33 (1972), 229–234. MR 291138
- [Boy81a] David W. Boyd, Kronecker’s theorem and Lehmer’s problem for polynomials in several variables, J. Number Theory 13 (1981), no. 1, 116–121. MR 602452
- [Boy81b] by same author, Speculations concerning the range of Mahler’s measure, Canad. Math. Bull. 24 (1981), no. 4, 453–469. MR 644535
- [Boy98] by same author, Mahler’s measure and special values of -functions, Experiment. Math. 7 (1998), no. 1, 37–82. MR 1618282
- [Bre51] Robert Breusch, On the distribution of the roots of a polynomial with integral coefficients, Proc. Amer. Math. Soc. 2 (1951), 939–941. MR 45246
- [Bro65] Hans Brolin, Invariant sets under iteration of rational functions, Ark. Mat. 6 (1965), 103–144 (1965). MR 194595
- [CLMM] A. Carter, M. Lalín, M. Manes, and A. B. Miller, On the dynamical Mahler measure, in preparation.
- [CLT09] Antoine Chambert-Loir and Amaury Thuillier, Mesures de Mahler et équidistribution logarithmique, Ann. Inst. Fourier (Grenoble) 59 (2009), no. 3, 977–1014. MR 2543659
- [CS93] Gregory S. Call and Joseph H. Silverman, Canonical heights on varieties with morphisms, Compositio Math. 89 (1993), no. 2, 163–205. MR 1255693
- [Den97] Christopher Deninger, Deligne periods of mixed motives, -theory and the entropy of certain -actions, J. Amer. Math. Soc. 10 (1997), no. 2, 259–281. MR 1415320
- [Dim] Vesselin Dimitrov, A proof of the Schinzel-Zassenhaus conjecture on polynomials, arXiv:1912.12545.
- [Dob79] E. Dobrowolski, On a question of Lehmer and the number of irreducible factors of a polynomial, Acta Arith. 34 (1979), no. 4, 391–401. MR 543210
- [EW99] Graham Everest and Thomas Ward, Heights of polynomials and entropy in algebraic dynamics, Universitext, Springer-Verlag London, Ltd., London, 1999. MR 1700272
- [FLM83] Alexandre Freire, Artur Lopes, and Ricardo Mañé, An invariant measure for rational maps, Bol. Soc. Brasil. Mat. 14 (1983), no. 1, 45–62. MR 736568
- [FRL06] Charles Favre and Juan Rivera-Letelier, Équidistribution quantitative des points de petite hauteur sur la droite projective, Math. Ann. 335 (2006), no. 2, 311–361. MR 2221116
- [GNY19] D. Ghioca, K. D. Nguyen, and H. Ye, The dynamical Manin-Mumford conjecture and the dynamical Bogomolov conjecture for split rational maps, J. Eur. Math. Soc. (JEMS) 21 (2019), no. 5, 1571–1594. MR 3941498
- [GTZ11] Dragos Ghioca, Thomas J. Tucker, and Shouwu Zhang, Towards a dynamical Manin-Mumford conjecture, Int. Math. Res. Not. IMRN (2011), no. 22, 5109–5122. MR 2854724
- [Ing13] Patrick Ingram, Variation of the canonical height for a family of polynomials, J. Reine Angew. Math. 685 (2013), 73–97. MR 3181564
- [Jul22] Gaston Julia, Mémoire sur la permutabilité des fractions rationnelles, Ann. Sci. École Norm. Sup. (3) 39 (1922), 131–215. MR 1509242
- [Kro57] L. Kronecker, Zwei Sätze über Gleichungen mit ganzzahligen Coefficienten, J. Reine Angew. Math. 53 (1857), 173–175. MR 1578994
- [Law83] Wayne M. Lawton, A problem of Boyd concerning geometric means of polynomials, J. Number Theory 16 (1983), no. 3, 356–362. MR 707608
- [Leh33] D. H. Lehmer, Factorization of certain cyclotomic functions, Ann. of Math. (2) 34 (1933), no. 3, 461–479. MR 1503118
- [Lyu83] M. Ju. Lyubich, Entropy properties of rational endomorphisms of the Riemann sphere, Ergodic Theory Dynam. Systems 3 (1983), no. 3, 351–385. MR 741393
- [Mah62] K. Mahler, On some inequalities for polynomials in several variables, J. London Math. Soc. 37 (1962), 341–344. MR 0138593
- [PST05] Jorge Pineiro, Lucien Szpiro, and Thomas J. Tucker, Mahler measure for dynamical systems on and intersection theory on a singular arithmetic surface, Geometric methods in algebra and number theory, Progr. Math., vol. 235, Birkhäuser Boston, Boston, MA, 2005, pp. 219–250. MR 2166086
- [Ran95] Thomas Ransford, Potential theory in the complex plane, London Mathematical Society Student Texts, vol. 28, Cambridge University Press, Cambridge, 1995. MR 1334766
- [Rit23] J. F. Ritt, Permutable rational functions, Trans. Amer. Math. Soc. 25 (1923), no. 3, 399–448. MR 1501252
- [RZ14] Mathew Rogers and Wadim Zudilin, On the Mahler measure of , Int. Math. Res. Not. IMRN (2014), no. 9, 2305–2326. MR 3207368
- [Sil07] Joseph H. Silverman, The arithmetic of dynamical systems, Graduate Texts in Mathematics, vol. 241, Springer, New York, 2007. MR 2316407
- [Smy71] C. J. Smyth, On the product of the conjugates outside the unit circle of an algebraic integer, Bull. London Math. Soc. 3 (1971), 169–175. MR 289451
- [Smy81] by same author, On measures of polynomials in several variables, Bull. Austral. Math. Soc. 23 (1981), no. 1, 49–63. MR 615132
- [ST12] L. Szpiro and T. J. Tucker, Equidistribution and generalized Mahler measures, Number theory, analysis and geometry, Springer, New York, 2012, pp. 609–638. MR 2867934
- [Zha95] Shouwu Zhang, Small points and adelic metrics, J. Algebraic Geom. 4 (1995), no. 2, 281–300. MR 1311351