Orthogonal rational functions with real poles, root asymptotics, and GMP matrices

By Benjamin Eichinger, Milivoje Lukić, and Giorgio Young

Abstract

There is a vast theory of the asymptotic behavior of orthogonal polynomials with respect to a measure on and its applications to Jacobi matrices. That theory has an obvious affine invariance and a very special role for . We extend aspects of this theory in the setting of rational functions with poles on , obtaining a formulation which allows multiple poles and proving an invariance with respect to -preserving Möbius transformations. We obtain a characterization of Stahl–Totik regularity of a GMP matrix in terms of its matrix elements; as an application, we give a proof of a conjecture of Simon – a Cesàro–Nevai property of regular Jacobi matrices on finite gap sets.

1. Introduction

There is a vast theory of orthogonal polynomials with respect to measures on and their root asymptotics, exemplified by the Ullman–Stahl–Totik theory of regularity. Let be a compactly supported probability measure and the corresponding orthonormal polynomials, obtained by the Gram–Schmidt process from in . Then

for outside the convex hull of , where is the essential support of and denotes the potential theoretic Green function for the domain ; if that domain is not Greenian, one takes instead. For measures compactly supported in , this theory can be interpreted in terms of self-adjoint operators. In particular, for any bounded half-line Jacobi matrix

with , ,

where denotes logarithmic capacity. For both of these universal inequalities, the case of equality (and existence of limit) is called Stahl–Totik regularity Reference 27; the theory originated with the case , first studied by Ullman Reference 30.

We extend aspects of this theory to the setting of rational functions with poles in . One motivation for this is the search for a more conformally invariant theory. Statements such as Equation 1.1, Equation 1.2 rescale in obvious ways with respect to affine transformations (automorphisms of ) which preserve , so it is obvious that an affine pushforward of a Stahl–Totik regular measure is Stahl–Totik regular. However, the point has a very special role throughout the theory: for a Möbius transformation which does not preserve , are rational functions with a pole at , and as defined by the functional calculus is not a finite band matrix. Thus, it is a nontrivial question whether a Möbius pushforward of a Stahl–Totik regular measure is Stahl–Totik regular.

The set of Möbius transformations which preserve is the semidirect group product , whose normal subgroup corresponds to the orientation preserving case. Denote by the pushforward of , defined by for Borel sets . As an example of our techniques, we obtain the following:

Theorem 1.1.

Let . If is a Stahl–Totik regular measure on and , then the pushforward measure is also Stahl–Totik regular.

However, we will mostly work in the more general setting when multiple poles on are allowed, which arises naturally in the spectral theory of self-adjoint operators. Denote by the multiplication operator by in . The matrix representation for in the basis of orthogonal polynomials is a Jacobi matrix, and through this classical connection, the theory of orthogonal polynomials is inextricably linked to the spectral theory of Jacobi matrices. In this matrix representation, resolvents are not finite-diagonal matrices. However, in a basis of orthogonal rational functions with poles at , the multiplication operators , …, , all have precisely nontrivial diagonals. The corresponding matrix representations are called GMP matrices; they were introduced by Yuditskii Reference 32.

Self-adjoint operators and their matrix representations are an important part of this work, so we choose to present the theory in a more self-contained way, using self-adjoint operators from the ground up; this has similarities with Reference 22. Some proofs could be shortened by using orthogonal polynomials with respect to varying weights Reference 27, Chapter 3, but some facts rely on the precise structure obtained by the periodically repeating sequence of poles.

We should also compare this to the case of CMV matrices: for a measure supported on the unit circle, Stahl–Totik regularity is still defined in terms of orthogonal polynomials, but the CMV basis Reference 4Reference 22 is given in terms of positive and negative powers of , i.e., orthonormal rational functions with poles at and . The symmetries in that setting lead to explicit formulas for the CMV basis in terms of the orthogonal polynomials; it is then a matter of calculation to relate the exponential growth rate of the CMV basis to that of the orthogonal polynomials, and to interpret regularity in terms of the CMV basis. In our setting, there is no such symmetry and no formula for orthonormal rational functions in terms of orthonormal polynomials.

In order to state our results in a conformally invariant way, we will use the following notations and conventions throughout the paper. The measure will be a probability measure on . We denote by its support in , and we consider its essential support (the support with isolated points removed), denoted

We will always assume that is nontrivial; equivalently, .

Fix a finite sequence with no repetitions, with for all . Consider the sequence where and for , ,

Applying the Gram–Schmidt process to this sequence in gives the sequence of orthonormal rational functions whose behavior we will study. We note that the special case , , corresponds to the standard construction of orthonormal polynomials associated to the measure (note that, since we denote by the support in , the statement implies that is compactly supported in ), and our first results are an extension of the same techniques.

The first result is a universal lower bound on the growth of in terms of a potential theoretic quantity. If is not a polar set, we use the (potential theoretic) Green function for the domain , denoted , and we define

Theorem 1.2.

For all ,

This is a good place to point out that our current setup is not related to the recent paper Reference 13, in which the behavior was compared to a Martin function at a boundary point of the domain. Here, the behavior is compared to a combination of Green functions Equation 1.4, all the poles are in the interior of the domain , and the difficulty comes instead from the multiple poles.

Another universal inequality for orthonormal polynomials comes from comparing their leading coefficients to the capacity of . In our setting, the analog of the leading coefficient must be considered in a pole-dependent way. Denote

By the nature of the Gram–Schmidt process, there is a such that

The Gram–Schmidt process can be reformulated as the -extremal problem

By strict convexity of the -norm, these -extremal problems have unique extremizers given by , and is explicitly characterized as a kind of leading coefficient for with respect to the pole at where , . Below, we will also relate the constants to off-diagonal coefficients of certain matrix representations.

The growth of the leading coefficients will be studied along sequences for a fixed , and bounded by quantities related to the pole . If is not a polar set, it is a basic property of the Green function that the limits

exist. Note that if , is precisely the Robin constant for the set . We further define constants by

Theorem 1.3.

For all , for the subsequence ,

Theorem 1.4.

The following are equivalent:

(i)

For some , for the subsequence ,

(ii)

For all , for the subsequence ,

(iii)

(iv)

For q.e. , we have ;

(v)

For some , ;

(vi)

For all , ;

(vii)

Uniformly on compact subsets of , .

Definition 1.5.

The measure is -regular if it obeys one (and therefore all) of the assumptions of Theorem 1.4.

In this terminology, Stahl–Totik regularity is precisely -regularity, i.e., -regularity for the special case , , . Theorems 1.2, 1.3, 1.4 are closely motivated by foundational results for Stahl–Totik regularity. A new phenomenon appears through the periodicity with which poles are taken in Equation 1.3 and the resulting subsequences : since is a normalization constant for , it is notable that control of along a single subsequence in Theorem 1.4i provides control over the entire sequence. This phenomenon doesn’t have an exact analog for orthogonal polynomials, where . We will also see below that this is essential in order to characterize the regularity of a GMP matrix using only the entries of the matrix itself and not its resolvents.

Moreover, we show that the regular behavior described by Theorem 1.4 is independent of the set of poles :

Theorem 1.6.

Let be two finite sequences of elements from , not necessarily of the same length. Then is -regular if and only if it is -regular.

Corollary 1.7.

Let . Let be a finite sequence of elements from . Then is -regular if and only if it is Stahl–Totik regular.

Thus, Theorem 1.4 should not be seen as describing equivalent conditions for a new class of measures, but rather a new set of regular behaviors for the familiar class of Stahl–Totik regular measures.

We consistently work with poles on since our main interest is tied to self-adjoint problems. Some of our results are in a sense complementary to the setting of Reference 27, Section 6.1, where poles are allowed in the complement of the convex hull of , and the behavior of orthogonal rational functions is considered with respect to a Stahl–Totik regular measure. Due to this, it is natural to expect that these results hold more generally, for measures on and general collections of poles and Möbius transformations. Moreover, in our setup the poles are repeated exactly periodically, but we expect this can be generalized to a sequence of poles which has a limiting average distribution. Related questions for orthogonal rational functions were also studied by Reference 3Reference 10.

As noted in Reference 27, Section 6.1, poles in the gaps of can cause interpolation defects in the problem of interpolation by rational functions. In our work, these interpolation defects show up as possible reductions in the order of the poles. For example, consider . Then, by construction, is allowed a pole at of order at most . However, if is symmetric with respect to , the functions will have an even/odd symmetry. Since contains a nontrivial multiple of , it follows that . By this symmetry, the actual order of the pole at is for some even , so it cannot be equal to (it will follow from our results that in this case, the order of the pole is ). The same effect can be seen for the pole at for . In the polynomial case, this does not occur: always has a pole at of order exactly .

We will consider at once the distribution of zeros of and the possible reductions in the order of the poles. We will prove that all zeros of are real and simple, and that . We define the normalized zero counting measure

Although we normalize by , may not be a probability measure: however . Therefore, normalizing by instead of by would not affect the limits as .

We will now describe the weak limit behavior of the measures as . To avoid pathological cases, we assume that is not polar; in that case, denoting by the harmonic measure for the domain at the point , we define the probability measure on ,

The results below describe weak limits of measures in the topology dual to .

Theorem 1.8.

Let be a probability measure on . Assume that is not a polar set.

(a)

If is regular, then .

(b)

If , then is regular or there exists a polar set such that .

We now turn to matrix representations of self-adjoint operators. Fix a sequence such that for some . A half-line GMP matrix Reference 32 is the matrix representation for multiplication by in the basis for this sequence ; its matrix elements are

The condition that for some guarantees that for , so these matrix elements generate a bounded operator on such that , where denotes the standard basis of . We say that .

GMP matrices have the property that some of their resolvents are also GMP matrices; namely, for any , where is the Möbius transform and .

Note that the special case , gives precisely a Jacobi matrix. A Jacobi matrix is said to be regular if it is obtained by this construction from a regular measure; analogously, we will call a GMP matrix regular if it is obtained from a regular measure. Just as regularity of a Jacobi matrix can be characterized in terms of its off-diagonal entries, we will show that regularity of a GMP matrix can be characterized in terms of its entries in the outermost nontrivial diagonal. We will also obtain a GMP matrix analog of the inequality Equation 1.2.

The GMP matrix has an additional block matrix structure; in particular, for a GMP matrix with , on the outermost nonzero diagonal , the only nonzero terms appear for , and those are strictly positive. Thus, we denote

Theorem 1.9.

Fix a probability measure with and a sequence with . Then

Moreover, the measure is Stahl–Totik regular if and only if

The proof will use a relation between the sequence and the constants . In particular, the characterization of regularity in Theorem 1.9 is made possible by the characterization of regularity in terms of the subsequence for any single . Theorem 1.9 also corroborates the perspective that regularity of the measure is the fundamental notion which manifests itself equally well in many different matrix representations.

Since the resolvents are also GMP matrices and their measures are pushforwards of the original measure, they are also regular GMP matrices; in this sense, Theorem 1.9 provides criteria for regularity, one corresponding to each subsequence , .

As an application of this theory, we show that it provides a proof of a theorem for Jacobi matrices originally conjectured by Simon Reference 23. Let be a compact finite gap set,

and denote by the set of almost periodic half-line Jacobi matrices with Reference 5Reference 14. Through algebro-geometric techniques and the reflectionless property, this class of Jacobi matrices has been widely studied for their spectral properties and quasiperiodicity (see also Reference 26Reference 31 for more general spectral sets). They also provide natural reference points for perturbations, which is our current interest. On bounded half-line Jacobi matrices , we consider the metric

On norm-bounded sets of Jacobi matrices, convergence in this metric corresponds to strong operator convergence. However, instead of distance to a fixed Jacobi matrix , we will consider the distance to ,

Denote by the right shift operator on , . The condition as is called the Nevai condition. For , this corresponds simply to the commonly considered condition , as Reference 18. In general, as a consequence of Reference 21, the Nevai condition implies regularity. The converse is false; however:

Theorem 1.10.

If is a compact finite gap set and is a regular Jacobi matrix with , then

The condition Equation 1.13 is described as the Cesàro–Nevai condition; it was first studied by Golinskii–Khrushchev Reference 15 in the OPUC setting with essential spectrum equal to . Theorem 1.10 was conjectured by Simon Reference 23 and proved in the special case when is the spectrum of a periodic Jacobi matrix with all gaps open by using the periodic discriminant and techniques from Damanik–Killip–Simon Reference 7 to reduce to a block Jacobi setting. It was then proved by Krüger Reference 17 by very different methods under the additional assumption . While this is a common assumption in the ergodic literature, regular Jacobi matrices do not always satisfy it: Reference 22, Example 1.4 can easily be modified to give a regular Jacobi matrix with spectrum and . We prove Theorem 1.10 in full generality by applying Simon’s strategy and, instead of the periodic discriminant and techniques from Reference 7, using the Ahlfors function, GMP matrices, and techniques of Yuditskii Reference 32.

For the compact finite gap set , among all analytic functions which vanish at , the Ahlfors function takes the largest value of . The Ahlfors function has precisely one zero in each gap, denoted for , a zero at , and no other zeros; see also Reference 25, Chapter 8. In particular, for the finite gap set , this generates a particularly natural sequence of poles .

The Ahlfors function was used by Yuditskii Reference 32 to define a discriminant for finite gap sets,

This function is not equal to the periodic discriminant, but it has some similar properties and it is available more generally (even when is not a periodic spectrum). Namely, extends to a meromorphic function on and . It was introduced by Yuditskii to solve the Killip–Simon problem for finite gap essential spectra. In fact, the discriminant is a rational function of the form

for some ; in particular, we will explain that the constants in Equation 1.15 match the general definition Equation 1.6.

As a first glimpse of our proof of Theorem 1.10, we note that it uses the following chain of implications. Starting with a regular Jacobi matrix with essential spectrum , by a change of one Jacobi coefficient, which does not affect regularity, we can assume that (Lemma 7.1). Under this assumption, regularity of the Jacobi matrix implies regularity of the corresponding GMP matrix and the resolvents , , which can be characterized in terms of their coefficients by Theorem 1.9. By properties of the Yuditskii discriminant, this further implies regularity of the block Jacobi matrix . Let us briefly recall that a block Jacobi matrix is of the form

where and are matrices, , and for each . Type 3 block Jacobi matrices have each lower triangular and positive on the diagonal. An extension of regularity to block Jacobi matrices was developed by Damanik–Pushnitski–Simon Reference 8; in particular, is regular for the set if and

This chain of arguments will result in Lemma 1.11:

Lemma 1.11.

Let be a regular Jacobi matrix, a finite gap set, and the corresponding sequence of zeros of the Ahlfors function. Assuming for , denote by the GMP matrix corresponding to with respect to the sequence . Then is a regular type 3 block Jacobi matrix with essential spectrum .

With Lemma 1.11, it will follow that obeys a Cesàro–Nevai condition. That Cesàro–Nevai condition will imply Equation 1.13 by a modification of arguments of Reference 32. The strategy is clear: just as Reference 32 uses a certain square-summability in terms of to prove finiteness of -norm of , we will use Cesàro decay in terms of to conclude the Cesàro decay Equation 1.13. This can be expected due to a certain locality in the dependence between the terms of the series considered; this idea first appeared in Reference 23 in the setting of periodic spectra with all gaps open. However, some care is needed, since the locality is only approximate in some steps; this is already visible in Equation 1.12. Also, substantial modifications are needed throughout the proof due to the possibility of (this cannot happen in the Killip–Simon class), which locally breaks some of the estimates. The fix is that this can only happen along a sparse subsequence, but the combination of a bad sparse subsequence and approximate locality means that we cannot simply ignore a bad subsequence once from the start; we must maintain it throughout the proof. A related issue arises with the Cesàro version of a Killip–Simon type functional. We will describe the necessary modifications to the detailed analysis in Reference 32.

The rest of the paper will not exactly follow the order given in this section. In Section 2, we describe the behavior of our problem with respect to Möbius transformations, and we describe the distribution of zeros of the rational function . In Section 3, we recall the structure of GMP matrices and relate their matrix coefficients to the quantities , and use this to provide a first statement about exponential growth of orthonormal rational functions on . In Section 4, we combine this with potential theoretic techniques to characterize limits of as and prove the universal lower bounds. In Section 5, we prove the results for -regularity and Stahl–Totik regularity. In Section 6 we describe a proof of Theorem 1.10.

2. Orthonormal rational functions and Möbius transformations

In Section 1, starting from the measure and sequence of poles , we defined a sequence and the orthonormal rational functions . In the next statement, we will denote these by and , in order to state precisely the invariance of the setup with respect to Möbius transformations.

Lemma 2.1.

If is a Möbius transformation which preserves , then

where and

Proof.

Note that the sequence does not have this property: is not equal to . However, if we denote

then it suffices to have

for some constants . If Equation 2.2 holds, then applying the Gram–Schmidt process to the sequences and will give the same sequence of orthonormal functions, up to the sign change , which is precisely Equation 2.1.

Note that if Equation 2.1 holds for , it holds for their composition, so it suffices to verify Equation 2.2 for a set of generators of . In particular, Equation 2.2 is checked by straightforward calculations for affine transformations and for the inversion , which implies the general statement since affine maps and inversion generate .

Let us emphasize what Lemma 2.1 does and what it doesn’t do. Since the Möbius transformation acts on both the measure and the sequence of poles, Lemma 2.1 does not by itself prove Theorem 1.1. Lemma 2.1 can only say that if is Stahl–Totik regular, then is -regular, which is not sufficient unless is affine. The proof of Theorem 1.1 will be more involved.

However, Lemma 2.1 provides a very useful conformal invariance for many of our proofs. This can be compared to choosing a convenient reference frame. Since potential theoretic notions such as Green functions are conformally invariant, our results will be invariant with respect to Möbius transformations. We will often use this invariance in the proofs to fix a convenient point at .

Note that this will be possible even though some objects are not conformally invariant. Some of our results compare the sequences with the , and although those objects are not preserved under conformal transformations, both sequences are affected in a compatible way so that the inequalities and equalities are preserved. Explicitly, fix and and a Möbius transformation (a reflection can be considered separately). Let us denote a local dilation factor . Then, we use Lemma 2.1 to compute

where is the leading coefficient . If is nonpolar, the Green function is conformally invariant so we find by another computation

where we have used that on . Thus, . If is polar, then is as well. From these calculations, it becomes elementary to verify that statements such as those in Theorems 1.3, 1.4 are conformally invariant.

Note that technical ingredients of the proof, such as polynomial factorizations, give a preferred role to so they break symmetry. For instance, we will often use the observation that the subspace can be represented as

for some suitable polynomial with factors which account for finite poles . We will use the representation Equation 2.3 after placing a convenient point at . This idea is already seen in the next proof.

Lemma 2.2.

All zeros of the rational function are simple and lie in . Moreover, .

Let , , and denote by the connected component of in . Then has no zeros in and at most one zero in any other connected component of .

Proof.

Fix and without loss of generality, assume . Then, in the representations Equation 2.3, we can notice that . In particular, then implies the representation for some polynomial of degree .

Recall that , is the unique minimizer for the extremal problem Equation 1.5. By complex conjugation symmetry, the minimizer is real. To proceed further, we study zeros of by using Markov correction terms.

We say that a rational function is an admissible Markov correction term if a.e. on and . In this case, using , we see that the function obeys

Thus, for small enough , the function

obeys . Since is of the form for some and in particular has the same leading coefficient as , the function contradicts extremality of . In other words, for the extremizer , there cannot be any admissible Markov correction terms.

Assume that has a nonreal zero . Then, since is real, , so the Markov correction term would be admissible, leading to contradiction.

Assume that has two zeros in the same connected component of ; then, the Markov correction term

would be admissible, leading to contradiction.

There are no zeros of in . Otherwise, if was a zero, the Markov term

would be admissible.

Finally, all zeros of are simple: otherwise, if was a double zero, the Markov term

would be admissible.

The properties of zeros of follow from those of . There may be cancellations in the representation , but since has at most a simple zero at , the only possible cancellations are simple factors , . Thus, .

The use of Markov correction factors is standard in the Chebyshev polynomial literature and is applied here with a modification for the -extremal problem (in the -setting, singularities in are treated with a separate argument near the singularity, which would not work here).

Corollary 2.3.

The measures are a precompact family with respect to weak convergence on . Any accumulation point is a probability measure and .

Proof.

By Lemma 2.2, , so precompactness follows by the Banach-Alaoglu theorem. If , then since , .

Let be a connected component of . Let us prove that . By Möbius invariance, it suffices to assume that is a bounded subset of .

Fix . As is a discrete set, we have

So, by Lemma 2.2, and by the Portmanteau theorem and sending , and .

3. GMP matrices and exponential growth of orthonormal rational functions

In this section, we consider orthonormal rational functions through the framework of GMP matrices. We begin by recalling the structure of GMP matrices Reference 32. The GMP matrix has a tridiagonal block matrix structure, with the beginnings of new blocks corresponding to occurrences of . Explicitly,

where is a matrix, is a matrix. For , are matrices; while for these appear in unmodified in the above, and are projections of and respectively. More precisely, let denote the upper triangular part of a matrix (excluding the diagonal) and the lower triangular part (including the diagonal). Then, indexing the entries of , , from to , we see they are of the form

where , with and (with the obvious modification if or ) and denotes the standard first basis vector of . and are projections of and ,

with the block matrix . We will refer to as the GMP coefficients of . While the precise structure will not be essential throughout the paper, we point out two things. First on the outermost diagonal of in each block there is only one nonvanishing entry, given by , which is positive and which is at a different position depending on the position of in the sequence . And secondly, in general as a self-adjoint matrix could depend on parameters, but we see that in fact they only depend on . This is not that surprising due to their close relation to three-diagonal Jacobi matrices. A similar phenomenon also appears for their unitary analogs Reference 6.

Remark 1.

For later reference, we provide an alternative point of view on the block structure of . The structure provided above is chosen so that is at the first diagonal position of the -blocks. Recall also that to these blocks we attached a column (with positive first entry ) to the right and a row at the bottom. If, instead of viewing this as a block matrix structure with blocks of size , we view this structure as overlapping blocks of size which overlap at the positions of , then those would contain all nonvanishing entries of the GMP matrix (i.e., it would also include the vector ). Moreover, the positive entries are exactly at the upper right and the lower left corner of the bigger block. Now placing the window of size on the top of the bigger block corresponds to the structure presented above. We will encounter in Section 6 that in other settings it may be more natural to place the block at the lower corner, and in this case the blocks will have structure similar to Equation 3.1.

Now the various notations for the off-diagonal blocks , the vectors which determine them, and the coefficients defined in Equation 1.8 are related as

The coefficients are a special case of the coefficients defined for , as

Namely , and the coefficients for instead occur as outermost diagonal coefficients for the GMP matrix . In our later applications to the discriminant of , both the coefficients of and of its resolvents will appear, so we will work with throughout.

Next, we connect the coefficients Equation 3.2 to the solutions of the -extremal problem Equation 1.5.

Lemma 3.1.

For all ,

Proof.

Let . By self-adjointness,

for some . By orthogonality, , so implies that

We now adapt to GMP matrices ideas from the theory of regularity for Jacobi matrices Reference 22.

Lemma 3.2.

Let . For all , .

Proof.

Fix and denote . For any ,

Since the vectors are orthonormal, by the Bessel inequality,

since .

Lemma 3.3.

For ,

Proof.

We adapt the proof of Reference 22, Proposition 2.2. It suffices to prove Equation 3.4 along the subsequences , , for . Moreover, due to -preserving conformal invariance, it suffices to fix and prove

under the assumption that . This allows us to use the associated GMP matrix .

Note that for any , since is an orthonormal basis of ,

This equality holds in , but since all functions are rational, it also holds pointwise. Thus, if we fix , the sequence is a formal eigensolution for at energy , i.e. componentwise. Since is represented as a block tridiagonal matrix, let us also write in a matching block form, as where

We also consider the projection of onto the first blocks,

and compute . By the block tridiagonal structure of , for we have . For , we have

so that . Moreover,

For , we again have . In conclusion, has only two nontrivial blocks,

In particular, we can compute

Since is self-adjoint and , by a standard consequence of the spectral theorem Reference 29, Lemma 2.7.,

Using Equation 3.6 and the Cauchy–Schwarz inequality gives

By Lemma 3.2, with ,

Applying the AM-GM inequality to the right-hand side of Equation 3.7 gives

which together with Equation 3.7 implies

Since , this implies that

Since , this implies by induction that

Combining this with Equation 3.8 gives a lower bound on which implies Equation 3.5.

The estimates in the previous proof also lead to the following:

Corollary 3.4.

For any , the quantities

are independent of .

Proof.

Assume . For , the estimate Equation 3.8 gives

which implies

and

Clearly, the right-hand sides don’t change if is shifted by , so Equation 3.9, Equation 3.10 hold for all with . By symmetry, since the roles of can be switched, we conclude that equality holds in Equation 3.9, Equation 3.10.

4. Growth rates of orthonormal rational functions

In this section, we will combine the positivity Equation 3.4 with potential theory techniques in order to study exponential growth rates of orthonormal rational functions. Our main conclusions will be conformally invariant, but our proofs will use potential theory arguments and objects such as the logarithmic potential of a finite measure ,

which is well defined when does not contain .

Theorem 4.1.

Fix and denote by the connected component of containing . Suppose there is a subsequence such that and as . Then uniformly on compact subsets of , we have

The function is determined by and ; in particular, if ,

Moreover,

(a)

is impossible;

(b)

If , the limit is ;

(c)

If , the limit extends to a positive harmonic function on such that

Proof.

By using -preserving conformal invariance, we can assume without loss of generality that . We will use the representation Equation 2.3 of the subspace . For , counting degrees of the poles leads to

with . This may not be the minimal representation of , but by the proof of Lemma 2.2, the only possible cancellations are simple factors for each , so we get the minimal representation with

where for each . All that matters is that as . It will be useful to turn this rational function representation into a kind of Riesz representation,

Since , note that is a compact subset of . Denote . For any , the map is continuous on , so as . In fact, convergence is uniform on compact subsets of : since and for all , the estimate

implies uniform equicontinuity of the potentials on compact subsets of , and the Arzelà–Ascoli theorem implies uniform convergence on compacts.

Note that b follows from Equation 4.1. By Corollary 2.3, and is harmonic on , so the right hand side extends to a harmonic function on and we denote this extension also by . By Lemma 3.3, is positive on , so ; moreover, by the mean value property, is positive on .

The remaining asymptotic properties follow from Equation 4.1. Under the assumption , is a compact subset of , and , . It then follows that as . Of course, near each .

Theorem 4.1 motivates interest in positive harmonic functions on . If is polar, by Myrberg’s theorem Reference 2, Theorem 5.3.8, any such function is constant. If is not polar, knowing the asymptotic behavior of at the poles, positivity of improves to the following lower bound on . Lemma 4.2 reflects a standard minimality property of the Green function Reference 11, Section VII.10.

Lemma 4.2.

Assume that is a nonpolar closed subset of . Let be a positive superharmonic function on . Suppose has an existent limit at for each finite , and has an existent limit at if one of the . Then

for . For , define

Then

Proof.

We will use a stronger, q.e. version of the maximum principle Reference 20, Thm 3.6.9. Define

which is bounded at for and so extends to a subharmonic function on . Since vanishes q.e. on , we have for q.e. ,

Now we show is bounded above on . Let be a union of small neighborhoods containing the points in . By the definition of the Green function, defines a harmonic and bounded function on . That is, there exists such that for all we have

Since , it follows on that

On the other hand, by properties of the Green functions we have

Then, by assumption, for , as and, in particular, the difference is bounded in a small neighborhood of . Thus, is bounded above on .

So, by the maximum principle on . Since , we have Equation 4.6.

Lemma 4.3.

Under the same assumptions as Lemma 4.2, the following are equivalent:

(i)

Equality in Equation 4.6 for all with

(ii)

Equality in Equation 4.6 for a single with

(iii)

Equality holds in Equation 4.5

Proof.

i ii is trivial. Suppose then ii; with the notation of Lemma 4.2, by assumption, and achieves a global maximum. By the maximum principle for subharmonic functions Reference 20, Theorem 2.3.1, on . Finally, if iii holds, then evaluating for each yields i.

We will now prove Theorems 1.2 and 1.3.

Proof of Theorem 1.2.

Using conformal invariance, we take . Fix and select a sequence such that

By precompactness of the , we may pass to a further subsequence, which we denote again by , so that and for some and . Then for as in Theorem 4.1,

on . If , then there is nothing to show. Suppose . If is not polar we apply a of Theorem 4.1 to find , and we may use c of the same theorem and Lemma 4.2 to conclude.

If instead is polar, by Myrberg’s theorem, is constant on . Computing the limit at we see . In particular, for .

Proof of Theorem 1.3.

Fix and assume again by conformal invariance that . Using precompactness of the measures , we find a subsequence with

and . If , we are done. Suppose then , then we have by Theorem 4.1a, . Furthermore, if is nonpolar, by c and Lemma 4.2, on . In particular, by the representation Equation 4.1 we see that , and so Equation 4.6 yields the desired inequality.

If instead is polar, by Theorem 1.2, for each ,

and so by Theorem 4.1b, .

5. Regularity

We will begin by proving a version of Theorem 1.4 for a fixed .

Lemma 5.1.

Fix . Along the subsequence , the following are equivalent:

(i)

;

(ii)

For q.e. , we have ;

(iii)

For some , ;

(iv)

For all , ;

(v)

Uniformly on compact subsets of , .

Proof.

Using conformal invariance, we will assume throughout the proof that . First, suppose that is polar. In this case ii is vacuous, and since , iii and iv are trivially true. Since , i follows from Theorem 1.3. As in the proof of Theorem 4.1, weak convergence of measures implies uniform on compacts convergence of their potentials. Thus, since are a precompact family, so are . Thus, the convergence implies that uniformly on compact subsets of , so v holds.

For the remainder of the proof, we will assume is not polar. Moreover, we will repeatedly use the fact that if any subsequence of a sequence in a topological space has a further subsequence which converges to a limit, then the sequence itself converges to this limit. In particular, when concluding v, we apply this fact in the Fréchet space of harmonic functions on with the topology of uniform convergence on compact sets.

iiiv: Given a subsequence of , using precompactness of the measures , we pass to a further subsequence with and , with real or infinite. By Theorem 4.1, uniformly on compact subsets of ,

with given by Equation 4.1. Using the assumption, for some , we have

So, by Theorem 4.1, and has a harmonic extension to . Furthermore, by Lemma 4.2, . By assumption, we have the opposite inequality at , and so, by the maximum principle for harmonic functions, on , and in particular on . Thus, we have v.

viv: For , and there is nothing to show. Fix and let be a subsequence with . By passing to a further subsequence, we may assume , and where is real or infinite. By the assumption, we have on . So, by a and b, and extends to a harmonic function on . By the representation Equation 4.1, we may extend subharmonically to . On this set, is also subharmonic, so, by the weak identity principle Reference 20, Theorem 2.7.5, on . Thus, by the principle of descent Reference 27, A.III, we have

and iv follows.

v i: Given a subsequence of , we use precompactness of the to pass to a further subsequence with and . Then in the notation of Theorem 4.1 and by assumption, for a

So by Lemma 4.3, . Thus, is the only accumulation point of in and we have i.

iv: As before, we fix a subsequence of and use precompactness to pass to a further subsequence with . Then, by Theorem 4.1 and in the notation introduced there, uniformly on compact subsets of ,

where is given by Equation 4.1 with . Thus, by Lemma 4.3ii, on . Since the initial subsequence was arbitrary, we have v.

iv ii: Recalling that the Green function vanishes q.e. on , the claim follows.

ii v: Fixing a subsequence of , we again use precompactness to select a further subsequence such that and , . By the upper envelope theorem, there is a polar set such that on , . Now, we let , which is polar by assumption, and , which is polar by Reference 20, Theorem 3.5.1. Then, for a , we have

So by Theorem 4.1a. Thus, by c of the same theorem, uniformly on compact subsets of

and extends to a positive harmonic function on with logarithmic poles at each of the . So, extends to a harmonic function on , and there by Lemma 4.2. We now show that in fact using the stronger, q.e. maximum principle.

We use the equality in Equation 4.1 to extend to a subharmonic function on . By the upper envelope theorem and the assumption again, for

Then, for these , since is positive, we have

by upper semicontinuity. So, for q.e. .

Since is upper semicontinuous on the compact set , there is an so that . As in the above, now for any , we have

So, there is a neighborhood of with . Since the difference is harmonic on , we conclude that . Thus, by the maximum principle and the reverse inequality, on . Since the first sequence was arbitrary, we have v.

Since the implication iviii is clear, we may conclude.

We now put the subsequences together and use Corollary 3.4 to show that regular behavior occurs for one if and only if it happens for all.

Proof of Theorem 1.4.

Applying Lemma 5.1 for all implies equivalence of conditions (ii), (iv), (v), (vi), (vii) from Theorem 1.4. By Corollary 3.4, for some , the condition

holds for one value of if and only if it holds for all. Due to Lemma 5.1, this immediately implies equivalence of conditions (i) and (iii) from Theorem 1.4. It remains to prove equivalence of (ii), (iii).

ii iii: For and , denote by the integer such that and is divisible by . Then as so ii implies . Taking the product over gives iii.

iii ii: Similarly to the above, Theorem 1.3 shows that for all ,

Thus, if ii was false, this would mean that for some , . Taking products over , we would have

(the last step again uses Equation 5.2 for all ). This would contradict iii, so the proof is complete.

We now prove a seemingly special case of Corollary 1.7.

Proposition 5.2.

Assume that the sequence contains . Then is Stahl–Totik regular if and only if it is -regular.

Proof.

Let us assume that is -regular and let denote the orthonormal polynomial with respect to . Fix . Since is in , , so the orthonormal polynomials can be expressed on the basis of orthonormal rational functions as

Thus, in particular, and we get

By Theorem 1.4, for q.e. ,

Thus, for q.e. , Equation 5.3 implies

Thus, is Stahl–Totik regular.

Conversely, assume that is Stahl–Totik regular. For , the polynomial is a divisor of , so we can write where . For any there exists a polynomial such that on . Thus,

and since is normalized. Since is a polynomial of degree at most , similarly to the above, representing it in the basis of polynomials shows

Since as , the supremum in Equation 5.7 grows subexponentially whenever Equation 5.5 holds. By Equation 5.6, this implies

Since is arbitrary, we conclude that Equation 5.5 implies Equation 5.4, so Equation 5.4 holds q.e. on .

From this seemingly special case, Theorem 1.6 and Corollary 1.7 follow easily:

Proof of Theorem 1.6.

By applying a conformal transformation, the special case shows that is -regular if and only if it is -regular for any single in . By applying this twice, we conclude that if , have a common element, then is -regular if and only if it is -regular.

By applying that conclusion twice, we will finish the proof. Namely, for arbitrary , , choose a sequence which has common elements with both and . Then is -regular if and only if it is -regular if and only if it is -regular.

Proof of Corollary 1.7.

The result follows by taking in Theorem 1.6.

Proof of Theorem 1.1.

By Lemma 2.1, is Stahl–Totik regular if and only if is -regular, and by Corollary 1.7, this is equivalent to Stahl–Totik regularity of .

Proof of Theorem 1.8.

(a) We note that by Corollary 1.7 we may use Theorem 1.4. Fix , and use conformal invariance to assume . Given a subsequence of , we use precompactness to pass to a further subsequence with . We write

which we will use to show . By ii, we may apply Theorem 4.1 with . Then, vii yields off the real line, and thus the equality between the representations Equation 4.1 and Equation 5.8 gives on . By the weak identity principle, this equality extends to . Applying the distributional Laplacian to both sides gives . Thus, .

(b) The main ingredient is a variant of Schnol’s theorem; for any , , so

By Tonelli’s theorem, it follows that -a.e., so there exists a Borel set with such that

Suppose is not regular. Then, by Theorem 1.4ii, there is a with

Using conformal invariance, we may assume , and we can pass to a subsequence such that , where by Theorem 4.1a. Since , by comparing Equation 4.1 and Equation 5.8, we have for ,

where