On harmonic functions on and the polynomial Pell equation
There have been many studies on “” harmonic functions, differential forms, or spinors recently, for example [1][2][3][4]. It is then natural to ask what the most simple case is like: harmonic functions on with point singularities? This paper studies such harmonic functions and finds its unexpected relationship with the polynomial Pell equation.
Remark: Unlike the more ambitious ongoing program of studying harmonic functions such that on the singular locus (see, for example, [3][4]), this paper only requires to have locally finite norms near , which means, is allowed to be unbounded near . The readers should not confuse what we study here with the other ongoing program.
Summary of the Results
Let be , where is a collection of an even number of isolated points. We will also identify with and suppose . Then there is a unique real line bundle in that has monodromy around each point in . An -valued function/differential form is also called a function/differential form.
Locally, a function can be regarded as an ordinary function under a trivialization of . This function changes sign in a different trivialization, which is why it is called a “” function. In particular, the derivatives of make sense if is differentiable (which are also functions).
Definition 0.1.
A harmonic function is a function whose Laplacian is . That is to say, under any local trivialization,
In addition, in this paper, we assume that has a bounded norm on any bounded open subset of (including an open neighborhood of ).
Choose a large disk that contains as a subset. Then the line bundle is trivial on . For convenience, we fix a trivialization of in . Under this trivialization, a function is an ordinary real-valued function in .
Here are the main results of this paper (proofs will be deferred to later sections).
Theorem 0.2.
Suppose
Then there is a natural 1-1 correspondence between any two of the following three objects:
-
•
A harmonic function on .
-
•
An admissible entire function on , where admissible means for any ,
and Re means the real part.
-
•
An equivalence class of harmonic functions in , where is the origin, given by:
where is a constant.
To be more precise, their relationships are given by
Remark 0.3.
Strictly speaking, the integral may depend on a path chosen from to in and a sign chosen for . But for any paths chosen from to each and any sign chosen for the admissible conditions are equivalent.
Moreover, the admissible condition is equivalent to saying that the integral is well defined up to a sign (or, to be more precise, as a section of the line bundle ). Details can be found in section 1 remark 1.5.
There is an analog correspondence in the higher-dimensional case but even simpler: We assume that the ambient space is , where and is a compactly smoothly embedded sub-manifold of dimension in . Let be the unique real line bundle on with monodromy around . We can define harmonic functions on with monodromy around in the same way. Then
Theorem 0.4.
There is a bijection between:
-
•
harmonic functions on .
-
•
Harmonic functions on .
The bijection is given by
where is the coordinates of .
Remark 0.5.
Soon after the author wrote the first draft of this paper, professor Mazzeo told the author a shorter argument to prove the theorem 0.4, which indicates that this result might not be unexpected for experts. (Note that the case is different.) But the author chose not to change the original argument here for two reasons: 1. Mazzeo’s argument relies on more background knowledge on geometrical analysis, while the current one, though more tedious in details, is more elementary in a certain sense. 2. The author believes that the original argument could also potentially be illuminating in other studies (for example, the study of more complicated harmonic objects).
Nevertheless, the following corollary of theorem 0.4 is new and partially answers a question asked by Mazzeo which roughly says:
Suppose that there is a unit circle in the Euclidean space , is there a harmonic function on with monodromy around such that extends to a continuous function throughout ? Will it be useful in gluing arguments in gauge theory?
Corollary 0.6.
Suppose is a unit circle embedded in the Euclidean space . Let be the distance to the circle . Then for any positive integer , there is a harmonic function on with monodromy around such that as .
The proof is given in section 1.
After the digression on the higher-dimensional case, we come back to the two-dimensional case. Suppose and are given. Consider the following harmonic function in :
By theorem 0.2, there is a unique harmonic function (denoted , which depends on ) and a unique admissible entire function that corresponds to it. Then they satisfy the following properties:
Theorem 0.7.
The triple is “minimal” in the following sense:
-
•
Any harmonic function such that as is a scalar multiplication of by a real-valued constant.
-
•
The zero locus of : is compact. Moreover, any harmonic function that satisfies this property is a multiplication of by a real-valued constant.
-
•
Any admissible entire function in such that as is a scalar multiplication of by a real-valued constant.
Moreover, we have the following mysterious relationship between and the polynomial Pell equation:
Theorem 0.8.
If is a pair of polynomials that satisfies the following polynomial Pell equation:
then is a multiplication of up to a multiplication by a real valued constant.
Moreover, a pair satisfies the polynomial Pell equation if any only if the following algebraic expression of the Abel integration holds:
Remark 0.9.
-
•
Not all whose corresponding polynomial Pell equation has a solution. If it has a solution, then is called Pellian.
-
•
We may easily extend the definition of to the case where has roots of multiplicity greater than . See definition 2.3.
- •
-
•
Our definition of (through the correspondence given by theorem 0.2), which still makes sense even when is non-Pellian, to the author’s knowledge, is also new.
With the help of our generalized definition of (which makes sense even if is non-Pellian), there is a natural practical method to check whether any given is Pellian or not. (Roughly speaking, for any given , Pellian or not, one may find first and then check whether the Abel integration is well-defined or not.)
Details of this method are described in section 2. The author does not know whether this is the only known method or not. (The author suspects that it is not, since polynomial Pell equations have been a well-studied subject for a long time, and the author is not an expert in that.) But the approach through harmonic functions is new.
There are also some miscellaneous studies on the zero locus of harmonic functions (and their relationship with ) which are stated in section 3 but omitted in the introduction.
Acknowledgment: I want to thank my Ph.D. advisor, Prof. Taubes, for his endless support, inspiration, and encouragement. Several arguments used in the paper are due to Taubes (which will be mentioned precisely), and more ideas were inspired by him. This paper cannot be finished without his help. I also want to thank Zhenkun Li, who helped me calculate the first homology group of the double branched covering over and Prof. Mazzeo, who encourages and supports me all the time. And I want to thank Boyu Zhang, Kai Xu, Ziquan Yang, Yongquan Zhang, Siqi He for helpful discussions with them.
1 Properties of harmonic functions
Lemma 1.1.
If is a harmonic function on , then for each , there exists an open neighborhood of , and a holomorphic function on , such that
where Re is the real part and
We call the half integer the degree of at , or .
Proof.
Suppose , where is the polar coordinates. Take the Fourier series of with respect to in a small neighborhood of . Locally can be written as
where are complex-valued functions in .
(The Fourier series exists because is an ordinary smooth function on the double cover of a neighborhood of with the exclusion of itself, and one can do the Fourier series there. The integer terms disappear because has monodromy .)
Based on the Laplacian equation,
So , where and are complex-valued constants. Since has locally finite norm, must be . So
Let and let be the smallest integer such that Then the proof is finished. ∎
Here are two immediate corollaries.
Corollary 1.2.
If is a harmonic function, then can be extended to continuously by setting on .
Corollary 1.3.
The expression represents a holomorphic function in up to a sign. Suppose
then represents a holomorphic function in up to a sign. When , has a zero point of degree at .
Note that in general, is not a global holomorphic function.
Theorem 1.4.
There exists an entire function on such that
Moreover, if , then has a zero point at of degree .
Proof.
Both and are 1-forms on . And
Note that
So is a holomorphic function on with monodromy at .
Let . Then is an ordinary holomorphic function in . (Strictly speaking, is determined only up to a sign here. But if we fix a sign convention for as a section of , then is unique. We will always assume that this is the case.)
Let be the function defined in lemma 1.1 and corollary 1.3, then is holomorphic in a neighborhood of and
Thus is also holomorphic in a neighborhood of . And is an entire function in . ∎
Remark 1.5.
If is a harmonic function, then . So we can always write as
for some entire function in . However, clearly only admissible entire functions in make the above expression well-defined harmonic functions.
In general, an integral such as
depends on the path chosen in from a lifting of to a lifting of , where is the surface described by , regarded as a double branched cover of .
The first homology group of is generated by the loops as follows:
Choose any path from to in that does not meet any other point in . Then is defined to be going along from to in any branch in , and go back along from to but in another branch.
In fact,
one can visualize the double-branched cover from to as follows:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/fede4d23-78b8-43b9-93e3-ad15b7ee15ce/1.jpg)
In this graph, let be the rotation of along the axis by . Then the branched double cover from to is given by the quotient . The intersection of with the axis is exactly the branching points . The two punctured points and are the two points above infinity.
Proposition 1.6.
The definition of admissible in thoerem 0.2 is well-defined.
Proof.
Clearly if one chooses a different sign of ,
only changes sign.
Now suppose for some chosen paths from to , and for each ,
Let be the loop on from to along on the one branch, and then coming back from to along but on the other branch. Then
But since are generators of , we know the integral
is purely imaginary along any closed loop in . So, the integral
has a well-defined real part which does not depend on the path chosen from to . In particular, the condition that “their real parts are all zero” does not depend on the specific paths that are used to define them and is well defined.
∎
Theorem 1.7.
There is a bijection between
-
•
Admissible entire functions on ;
-
•
harmonic functions on .
The bijection is given by
Proof.
-
•
: As shown in the Appendix A, if is admissible, then the real part of the integration
doesn’t depend on the path chosen from to except its sign (which is also fixed is a sign convention for is chosen). So its real part is a well-defined harmonic function. Call this function . We have for any . In order for to be a function, we need to check that the monodromy of around each point is .
Near each , the Taylor series looks like
So
which indeed has the correct monodromy around .
∎
Since is uniquely determined by , sometimes it is convenient to use the notation to emphasize its dependence on . However, if there is no ambiguity, the subscript can be omitted.
Suppose are the polar coordinates centered at the origin. Suppose is a ball with a large radius centered at the origin such that . On . Using Fourier series,
(1.8) |
where are complex numbers except that must be real. Capital letters and , representing the first and second parentheses, respectively, are the “unbounded” and “bounded” parts at infinity. We may also call them and respectively to emphasize their dependency on . Note that is actually a harmonic function on the whole , where is the origin. (But is not.)
Theorem 1.9.
Suppose that there are two harmonic functions in such that , then . In particular, if , then .
Proof.
Since is a harmonic function with . So it suffices to prove .
Assume when ,
We may assume , since otherwise we can change the trivialization of and flip the sign of . Then is a super-harmonic function on the entire which approaches at infinity. By maximum principle, it cannot be negative. So in particular, when , .
Suppose . Then let is the smallest positive integer such that . Then
which is impossible when is large enough.
∎
Corollary 1.10.
If the zero locus of is bounded, then for some constant . In addition, such (if exists) is unique up to a scalar multiplication by a constant.
Proof.
Suppose the zero locus of is bounded and in general,
We may assume when for some large enough .
Then is a harmonic function on the entire which is bounded below by when . It is well known that such a harmonic function with a lower bound must be a constant. But this constant can only be based on the expression of . So, the only possibility is . The uniqueness of up to a scalar multiplication follows directly from theorem 1.9. ∎
One may wonder for which choice of the harmonic function exists. In fact, we have the following theorem:
Theorem 1.11.
There is a bijection between:
-
•
The equivalence classes of harmonic functions on , where is the origin, given by:
where is a constant.
-
•
harmonic functions on .
The bijection is given by sending to .
We have already shown that the map is injective. So, the task is to prove that it is also surjective.
We need a lemma first:
Lemma 1.12.
Fix a path from to each on the double branched cover of so that the following expression makes sense:
And let . Then the following vectors
are linearly independent over .
Proof.
Suppose
Let
And let .
Then
Note that
So
where is a constant. But in order for to be well defined, must be . So . So . Hence , which implies . ∎
Now we prove that is surjective.
Choose any
Consider the following formal Laurent series
Let
Then is an entire function in such that
By the previous lemma, there exist real numbers and a corresponding polynomial
such that
for each .
Thus
defines a harmonic function.
Since
so
where is a constant. This implies that the unbounded part of is indeed the proscribed and that the proof of theorem 1.11 is finished.
Note that the combination of theorem 1.7 and theorem 1.11 is just theorem 0.2 stated in the introduction.
The remainder of this section concerns the case of the higher dimensions.
Recall that now the ambient space is , where and is a compact smooth manifold of dimension embedded smoothly into . Let be the unique real bundle in with monodromy around . We can similarly define the harmonic functions in with monodromy around .
Suppose is a harmonic function in with monodromy around , where is a smooth compactly embedded sub-manifold of codimension in , and is a bounded open subset of that contains . Suppose that the norm of is finite on .
In addition, choose a large enough number and an open ball of radius centered at the origin such that . Fix a trivialization of in . Then is viewed as an ordinary function on .
Lemma 1.13.
The function is a subharmonic function on . That is to say, for any ball centered at in , where ,
Note that we allow to have intersection with .
Proof.
If , then is an ordinary harmonic function in up to the sign. So, the inequality is true by the mean value property.
In general, suppose is the ball of radius centered at . Since , the inequality is true when is small such that .
Let
Here
where is a constant that doesn’t depend on .
On the other hand, fix any small enough , there exists a constant which doesn’t depend on , such that
Note that has locally finite norm. The above inequality implies that
So
So . And for any such that . ∎
Corollary 1.14.
If we extend by setting on , then is a sub-harmonic function on the entire .
Note that we haven’t proved that the extension is continuous yet.
Lemma 1.15.
Suppose is a small tubular neighborhood of which is a bundle over , where is a disk of radius whose local polar coordinates are . (Note that may not be well-defined globally along ). For a small contractible relatively open subset of , let be the tubular neighborhood of (which is roughly with minor modifications on the metric). Then there exists a constant which doesn’t depend on , but may depend on , such that
Here volume of .
In particular
Proof.
Because has modronomy around , we know that for any ,
where is a constant. So there is a constant such that
And there is another constant such that
So
The final assertion in the lemma is because
∎
Lemma 1.16.
Let be a subset of such that for each , there exists a ball centered at such that
Then .
Proof.
Clearly is a closed subset of . It suffices to prove that is dense in .
For any small contractible relatively open subset of , we know from the previous lemma that
But
So the “average value” of on approaches as goes to . This implies that there exists at least one point such that is also in . Since this is true for any small contractible relative open subset of , we know that is dense. ∎
Theorem 1.17.
The function can be extended continuously in all setting in . This extended version of is a subharmonic function in .
Proof.
We have already proved that the extended version of is subharmonic. So, it suffices to prove that it is continuous at any point in .
From the previous lemma, for any small enough , . So there exists a ball centered at such that
Let be the ball centered at whose radius is half of the radius of . For any , let be the ball centered at whose radius is also half of the radius of . Then . And . So
Since is subharmonic, for any ,
Letting implies that
Thus is continuous at .
∎
Corollary 1.18.
There is a unique harmonic function on such that
Taubes’ proof of corollary 1.18
Consider all smooth function that equals when is large enough. Let the Dirichlet energy be
Consider the following space:
Suppose is a smooth function such that far away from the origin and that extends continuously to by setting on . Then we define the “norm” of as its Dirichlet energy:
The space of completion of the above type of smooth functions under the norm is still denoted as .
Note that if , then by Sobolev embedding, the norm of is bounded up to a constant by . As a corollary, also has finite norm for any ball whose closure is in .
Choose a sequence of smooth functions whose Dirichlet energy approaches the infimum in . One may assume, by possibly choosing a subsequence, that
-
•
weakly converges to in .
-
•
weakly converges to in .
The second bullet implies . The first bullets implies
Since , by the way that the sequence was constructed, achieves the minimal Dirichlet energy in .
By the method of variation, under any local trivialization of , one sees that is a weak solution of the Laplacian equation on any small ball whose closure is in . Since is also in , by a standard elliptic regularity argument, is harmonic on . Thus is a harmonic function on the entire .
Now we return to theorem 0.4. We first need a lemma.
Lemma 1.19.
Let be a connected, bounded open subset of such that . Suppose that the boundary is smooth. Suppose that a trivialization of on is given. Suppose is a smooth function on . Then there exists a unique harmonic function on (that is to say, harmonic on and continuous on ) with monodromy around (we just call it harmonic function when there is no ambiguity) such that
Proof.
Consider all smooth functions on that equals on the boundary.
Consider the energy function
The energy function is bounded above by the norm. By Poincare inequality, the norm is also bounded above by a constant times plus a constant which only depends on and .
Thus we can choose a sequence of smooth functions on that converges to the infimum of the energy among all such functions and weakly converges to a weak solution of the Laplacian equation in , which by choosing a further sub-sequence strongly converges in . By a standard regularity theorem, this converging function as a weak solution of the Laplacian equation is a harmonic function.
Suppose there are two harmonic functions on which has the same boundary values on . Then their difference is a harmonic function which equals on . But if , then let be any non-empty open connected component of , where is the zero locus of . Then such exists and is an ordinary non-zero harmonic function on whose boundary equals zero. This violates the maximal principle. Hence, the harmonic function is unique. ∎
Proof of theorem 0.4.
-
•
Let be the harmonic function on defined in the previous lemma. We may extend continuously to the entire by setting outside of , still denoted as . Suppose . Let be the short for which is the energy defined in the previous lemma.
Since outside of , we have
Let
Note that by the way was chosen, has the minimal energy among all functions in equal to in . And is an example of such function. Thus,
So . That is, does not increase in . Since it is bounded below by , it must converge. We may assume .
Suppose is large enough such that . Then for any , . In particular, . On the other hand, has the minimal energy among all harmonic functions on that is equal to on . And is one of such kind of functions. So
And
Note that has bounded support in . By Sobolev embedding, there exists a constant which does not depend on such that
This implies that is a Cauchy sequence (as goes to infinity) in . So, it converges to some function as . This function is locally a weak solution of the Laplacian equation and locally in , so it is a harmonic function.
On the other hand,
So
So for any ball of radius in ,
where is a large constant.
But is a harmonic function in . So, by a regularity argument, assuming that is the ball of radius with the same center as , we have another constant such that
Note that does not depend on the particular ball chosen. So, in fact, for some , we have
Letting implies that converges to uniformly as .
-
•
We may use the same argument as in the case. However, we introduce a different argument here.
In the spherical coordinates, the Laplacian equation is written as
And when is large enough,
where and are spherical harmonics with eigenvalue , and are constants.
Let
Then is the harmonic function on that we want.
Remark 1.20.
Like in the case , one sees from the proof that the function depends on and continuously in a suitable sense.
Now we prove the corollary 0.6. We may assume that our embeds in as follows:
Clearly there is an action of that restricts to a rotation on the embedded .
We may also choose local coordinates in a neighborhood of to represent the point: .
Lemma 1.21.
There is a sequence of linearly independent harmonic functions in that are invariant under the action.
Proof.
In fact, these harmonic functions can be chosen to have polynomial growth. Otherwise, there are even uncountable such harmonic functions.
Examples are: , where is the distance to the origin, and are Lendgre polynomials. It is well known that are spherical harmonics, and hence are the harmonic functions as desired.
∎
By theorem 0.4, we immediately have the following corollary:
Corollary 1.22.
There is a sequence of linearly independent harmonic functions in with monodromy around that are invariant under the action.
Here is a sketch of the proof of corollary 0.6:
Proof.
In the local coordinates,
Suppose is invariant under action, that is to say, independent with . Then and
We can construct the Fourier series of as in the proof of lemma 1.1.
Suppose
where are complex-valued functions in .
From corollary 1.22, there are infinitely many choices of linearly independent harmonic functions that clearly satisfy . We use induction on the power.
By induction, suppose that there are infinitely many choices of linearly independent such that as . All such form an infinite-dimensional vector space . For each , we have for all . From the standard regular singular point theory in ODE, we know that for all and that . The map from to sending to has an infinite-dimensional kernel. Clearly, any harmonic function in the kernel behaves like as . ∎
Remark 1.23.
Strictly speaking, in the above argument we need to check that all the higher-order remainder terms have all derivatives of the correct order as . Moreover, this boundary behavior is uniform for different Fourier components . However, these are all easy to check and are omitted here. In fact, there are many much stronger conclusions about what a harmonic function is like near its singular points (as a polyhomogeneous expansion) in [4],[3], [7], etc. We refer interested readers to these references.
2 The polynomial Pell equation
In this section we prove theorem 0.7, theorem 0.8, and describe the details of the practical method to check whether a given is Pellian or not. We work in the two-dimensional case. We assume that and are given.
By theorem 0.2, there exists a unique harmonic function, denoted as (or for short), such that the unbounded part at infinity has only a term . By corollary 1.10, it is also the unique harmonic function with bounded zero locus up to a real-valued scalar multiplication. This is the second bullet of the theorem 0.7.
Remark 2.1.
The reason we call it is due to the following observation of Taubes: if one uses a conformal map to send to , where is the south pole (corresponding to the point of infinity), then a harmonic function on becomes a harmonic function on . And becomes the version of the Green function on evaluated at . More details of this Green function can be found implicitly in [1] [2] or [3].
Theorem 2.2.
When , the entire function corresponding to in theorem 1.4 is a polynomial of degree , and vice versa (up to a multiplication of a constant of real value). We call this polynomial . Then is uniquely determined by (up to a multiplication by a constant of real value as always).
Proof.
Definition 2.3.
For each even degree polynomial (may not have distinct roots), we define a polynomial (up to a multiplication by a real number) in the following way:
-
•
If all roots of are distinct, then is the unique polynomial from the previous theorem.
-
•
Suppose , such that all roots of are distinct. Then .
If there is no ambiguity, we may omit the subscript and simply write it as .
Theorem 2.4.
Suppose is a harmonic function. Suppose the unbounded part of at infinity is
Then is a polynomial of degree if and only if is the largest integer such that is not zero.
Proof.
So
Therefore, has a pole of order at infinity if and only if is the largest integer such that is non-zero. But is an entire function. So, it has a pole of order at infinity if and only if it is a polynomial of degree . ∎
Note that there is a corollary: the leading coefficient of is real.
Remark: The harmonic function depends on and continuously in a suitable sense. In fact, in a suitable sense (for example, the topology, where is a large enough disk that contains all the roots of ), the following expressions for :
and the following expressions for
are all continuous, where is defined in the proof of theorem 0.4.
Recall that suppose is a polynomial with complex coefficients. The following equation is called the polynomial Pell equation, or the Pell equation for short:
where and are unknown polynomials with complex coefficients.
If the Pell equation for has a nontrivial solution (that is, ), then is called Pellian. Ohterwise, it is called non-Pellian.
One obvious observation is that, is non-Pellian when the degree is odd. And when the degree of is , it is Pellian if and only if it is square free. It is believed that when the degree of is at least , then Pellian polynomials are extremely rare. There are many studies on this topic; see, for example, [5] [8] [6].
For the remainder of this section, we assume that the degree of is even.
The following proposition and its corollaries are well known and are very easy to prove, and we omit the proofs here:
Proposition 2.5.
If is a solution to the Pell equation for with minimal degree. Let
Then are all the solutions to the Pell equation for .
Corollary 2.6.
Suppose is a solution to the Pell equation for . Let . Then different solutions will give the same up to a multiplication by a real-valued constant.
Corollary 2.7.
If is Pellian and if is a solution to the Pell equation. Let . Then is a polynomial and the following Abel integration has an algebraic expression:
Thus in order to prove theorem 0.8, the only task is to verify that given above is the same as our up to a scalar multiplicity by a real-valued constant. Here is the proof of theorem 0.8:
Proof.
We only need to check that defines a harmonic function. Then by uniqueness given by theorem 0.2, must be the same as .
First of all,
So
Thus has a well-defined real part on the entire except a sign ambiguity for . But a different choice of the sign of also changes only the sign of . So, this sign ambiguity corresponds to the correct monodromy for a function. Thus, is a well-defined harmonic function on .
∎
Note that this above theorem partially explains why we choose to have the second bullet of the definition 2.3. The author wonders whether can be computed from in a purely algebraic way (that is, not doing the integrals).
When the coefficients of are real numbers, then the coefficients of are also real numbers. In fact, it is not hard to check that , where means the polynomial whose coefficients are conjugates of the coefficients of .
Finally, inspired by theorem 0.2, we have the following general method to check whether a polynomial is Pellian or not.
Theorem 2.8.
Suppose is a monic polynomial of even degree with distinct roots. Then is Pellian if and only if all the following integrals
are integer multiplications of .
Proof.
If is Pellian, suppose
Then and
But from Pell equation, for any , so
is an integer multiplication of .
The other direction: if all
are integer multiplications of , then
defines a holomorphic function on for the following reason:
We know that the topology of is a surface of genus minus two points. The first homology of is generated by the following cycles:
For each , one goes from to in one branch and go back from to using the same path but in another branch. This loop on is denoted as .
We have assumed that the path integration of is a multiplication of along any other loop . So the exp of this integral is well-defined holomorphic function on without ambiguity.
Finally, by GAGA (see for example, [9]), a holomorphic function on with polynomial growth must be a polynomial in and . Since , we may write this polynomial as
And the unique harmonic function with bounded zero locus is
Since has monodromy around any point , a change of the sign of leads to a change of sign in as well. We have
So We may assume , where is a complex-valued constant with norm . Then is a nontrivial solution to the Pell equation for . ∎
Remark: Although we assumed that does not have distinct roots here, it is standard to reduce the case more general to the distinct roots case because of the following well-known and easy-to-proof theorem (whose proof is thus omitted):
Theorem 2.9.
If has a common factor with . Let be the order of in and be the order of in . Then . In addition, is Pellian if and only if .
To finish this section, we write down a concrete example of a harmonic function. This example is due to Taubes, who told me in a private conversation. Exactly this example inspired me to think about the relationship between the polynomial Pell equation and harmonic functions.
Example 2.10.
Let , then
3 The zero locus
This short section studies the zero locus of the harmonic functions in and its relationship to in the correspondence given by theorem 0.2.
For an ordinary harmonic function in , there is a famous theorem that states that the zero locus is a union of smooth curves such that, at any intersection points of the curves, there are only finitely many curves that intersect equiangularly. See, for example, the textbook [10] chapter X section 9. We deduce an analog of this theorem for harmonic functions on .
Theorem 3.1.
Suppose is a harmonic function on . Then the zero locus of is the union of locally finite smooth curves whose boundaries are empty or belong to . At any intersection point of those smooth curves (including intersections at boundaries), they form an equiangular system. In addition to that, any subset of the zero locus of does not form a loop.
Proof.
Here is the sketch. Suppose is a zero point of a harmonic function .
If does not belong to , then it is the same as the usual situation and a proof can be found in the textbook [10]. Otherwise, belongs to . Near ,
Suppose is the smallest number such that . We may assume is real, then
So
The above two expressions, together with the implicit function theorem, are enough to show that in a neighborhood of , the zero locus of are smooth curves that intersect at their common boundary in an equiangular way.
That the zero locus does not form a closed loop is just a consequence of the maximal principle. And here is the argument:
If on the contrary, the zero locus of has a closed loop as a subset. Then at least one connected component of is bounded, denoted as . Note that setting defines a trivialization of on . And is an ordinary harmonic function on without zero points in it. In addition, can be extended as a continuous function to the closure of by setting on the boundary. This implies that has a positive local maximum in , which is impossible. ∎
Definition 3.2.
Suppose is a zero point of a harmonic function . Suppose that there are a non-zero number and a half integer such that in a neighborhood of , then is called the degree of at . From the proof of the previous theorem, we know that the zero locus of near is an intersection of curves that share a common boundary point at and intersect at equiangularly.
Theorem 3.3.
Suppose is a zero point of a harmonic function . Suppose is the entire function defined in section 1. If the degree of at is larger than , then is a zero point of with degree (the largest integer that is no greater than ).
Proof.
If is a zero point of whose degree , then is also a zero point of whose degree is . Since , is a zero point of with degree . ∎
Here are some easy examples.
Example 3.4.
Suppose is Pellian, and suppose
Then the zero locus of is given by the set
This implies at a point in the zero locus,
In particular,
-
•
If has only two roots and if they are distinct, then
Then we can choose , and the zero locus is the straight line segment connecting and . In this situation, .
-
•
If , and , then the zero locus is
An typical example of this kind is: .
Example 3.5.
If is a polynomial of degree with distinct roots and suppose , then the degree of is one. We may assume . (Note that may not be Pellian in general.) Then there are two possibilities:
-
•
The zero locus is a combination of three curve segments and we may assume that they share a common end at and their other ends are , , respectively. They intersect equiangularly at and do not intersect at any other point. In this situation, . An typical example of this kind is: .
-
•
The zero locus is a combination of two curve segments and we may assume that one of them connects and , the other one connects and . If they intersect at a point in an orthogonal way, then this point must be . Otherwise, they do not intersect.
The author is very curious how the four points and determine in general.
References
- [1] C. H. Taubes and Y. Wu, “Examples of singularity models for Z/2 harmonic 1-forms and spinors in dimension three,” J. G¨okova Geom. Topol., 2021.
- [2] H. T. Clifford and W. Yingying, “Topological aspects of Z/2Z eigenfunctions for the Laplacian on S2.”
- [3] S. Donaldson, “Deformations of Multivalued Harmonic Functions,” Quarterly Journal of Mathematics, vol. 72, no. 1-2, 2021.
- [4] Siqi He, “Existence of nondegenerate Z2 harmonic 1-forms via Z3 symmetry,” 2022.
- [5] A. Dubickas and J. Steuding, “The polynomial Pell equation,” Elemente der Mathematik, vol. 59, no. 4, 2004.
- [6] M. B. Nathanson, “Polynomial Pell’s equations,” Proceedings of the American Mathematical Society, vol. 56, no. 1, 1976.
- [7] G. J. Parker, “Deformations of $\mathbb Z_2$-Harmonic Spinors on 3-Manifolds,” 1 2023.
- [8] N. D. Kalaydzhieva, On problems related to multiple solutions of Pell’s equation and continued fractions over function fields. PhD thesis, 2020.
- [9] A. Neeman, Algebraic and Analytic Geometry. 2007.
- [10] C. N. Moore and O. D. Kellogg, “Foundations of Potential Theory.,” The American Mathematical Monthly, vol. 37, no. 10, 1930.