This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On 2\mathbb{Z}_{2} harmonic functions on 2\mathbb{R}^{2} and the polynomial Pell equation

Weifeng Sun

There have been many studies on “2\mathbb{Z}_{2}” 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: 2\mathbb{Z}_{2} harmonic functions on 2\mathbb{R}^{2} with point singularities? This paper studies such 2\mathbb{Z}_{2} harmonic functions and finds its unexpected relationship with the polynomial Pell equation.

Remark: Unlike the more ambitious ongoing program of studying 2\mathbb{Z}_{2} harmonic functions ff such that |df|=0|df|=0 on the singular locus KK (see, for example, [3][4]), this paper only requires |df||df| to have locally finite L2L^{2} norms near KK, which means, |df||df| is allowed to be unbounded near KK. The readers should not confuse what we study here with the other ongoing program.

Summary of the Results

Let XX be 2\K\mathbb{R}^{2}\backslash K, where KK is a collection of an even number of isolated points. We will also identify 2\mathbb{R}^{2} with \mathbb{C} and suppose K={z1,z2,,z2k}K=\{z_{1},z_{2},\cdots,z_{2k}\}\subset\mathbb{C}. Then there is a unique real line bundle \mathcal{L} in XX that has monodromy 1-1 around each point in KK. An \mathcal{L}-valued function/differential form is also called a 2\mathbb{Z}_{2} function/differential form.

Locally, a 2\mathbb{Z}_{2} function ff can be regarded as an ordinary function under a trivialization of \mathcal{L}. This function changes sign in a different trivialization, which is why it is called a “2\mathbb{Z}_{2}” function. In particular, the derivatives of ff make sense if ff is differentiable (which are also 2\mathbb{Z}_{2} functions).

Definition 0.1.

A 2\mathbb{Z}_{2} harmonic function ff is a 2\mathbb{Z}_{2} function whose Laplacian is 0. That is to say, under any local trivialization,

Δf=(x2+y2)f=0.\Delta f=(\partial_{x}^{2}+\partial_{y}^{2})f=0.

In addition, in this paper, we assume that |f||\nabla f| has a bounded L2L^{2} norm on any bounded open subset of XX (including an open neighborhood of KK).

Choose a large disk BB\subset\mathbb{C} that contains KK as a subset. Then the line bundle \mathcal{L} is trivial on \B\mathbb{C}\backslash B. For convenience, we fix a trivialization ι\iota of \mathcal{L} in \B\mathbb{C}\backslash B. Under this trivialization, a 2\mathbb{Z}_{2} function ff is an ordinary real-valued function in \B\mathbb{C}\backslash B.

Here are the main results of this paper (proofs will be deferred to later sections).

Theorem 0.2.

Suppose

D(z)=i=12k(zzi).D(z)=\prod\limits_{i=1}^{2k}(z-z_{i}).

Then there is a natural 1-1 correspondence between any two of the following three objects:

  • A 2\mathbb{Z}_{2} harmonic function ff on XX.

  • An admissible entire function u(z)u(z) on \mathbb{C}, where admissible means for any j=2,3,,2kj=2,3,\cdots,2k,

    Re(z1zju(z)D(z)𝑑z)=0,\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz)=0,

    and Re means the real part.

  • An equivalence class [h][h] of harmonic functions in \{O}\mathbb{C}\backslash\{O\}, where OO is the origin, given by:

    h1h2if and only ifh1h2Cwhen|z|+,h_{1}\sim h_{2}~{}~{}~{}~{}\text{if and only if}~{}~{}~{}h_{1}-h_{2}\rightarrow C~{}~{}\text{when}~{}~{}|z|\rightarrow+\infty,

    where CC is a constant.

To be more precise, their relationships are given by

f(z)=Re(z1zu(z)D(z)𝑑z)andlim|z|+(f(z)h(z))=Cfor some constant C.f(z)=\text{Re}(\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{D(z)}}dz)~{}~{}\text{and}~{}~{}\lim\limits_{|z|\rightarrow+\infty}(f(z)-h(z))=C~{}~{}~{}\text{for some constant $C$.}
Remark 0.3.

Strictly speaking, the integral z1zju(z)D(z)𝑑z\displaystyle\int_{z_{1}}^{z_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz may depend on a path chosen from z1z_{1} to zjz_{j} in KK and a sign chosen for D(z)\sqrt{D(z)}. But for any paths chosen from z1z_{1} to each zjz_{j} and any sign chosen for D(z)\sqrt{D(z)} the admissible conditions are equivalent. Moreover, the admissible condition is equivalent to saying that the integral Re(z1zu(z)D(z)𝑑z)\displaystyle\text{Re}(\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{D(z)}}dz) is well defined up to a sign (or, to be more precise, as a section of the line bundle \mathcal{L}). 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 X=n\KX=\mathbb{R}^{n}\backslash K, where n3n\geq 3 and KK is a compactly smoothly embedded sub-manifold of dimension n2n-2 in n\mathbb{R}^{n}. Let \mathcal{L} be the unique real line bundle on XX with monodromy 1-1 around KK. We can define 2\mathbb{Z}_{2} harmonic functions ff on XX with monodromy 1-1 around KK in the same way. Then

Theorem 0.4.

There is a bijection between:

  • 2\mathbb{Z}_{2} harmonic functions ff on n\K\mathbb{R}^{n}\backslash K.

  • Harmonic functions hh on n\mathbb{R}^{n}.

The bijection is given by

fh0when|𝐱|+,f-h\rightarrow 0~{}~{}~{}~{}~{}~{}\text{when}~{}~{}~{}~{}|\mathbf{x}|\rightarrow+\infty,

where 𝐗\mathbf{X} is the coordinates of n\mathbb{R}^{n}.

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 n=2n=2 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 2\mathbb{Z}_{2} 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 CC in the Euclidean space 3\mathbb{R}^{3}, is there a 2\mathbb{Z}_{2} harmonic function ff on 3\C\mathbb{R}^{3}\backslash C with monodromy 1-1 around CC such that |f||f| extends to a continuous function throughout 3\mathbb{R}^{3}? Will it be useful in gluing arguments in gauge theory?

Corollary 0.6.

Suppose CC is a unit circle embedded in the Euclidean space 3\mathbb{R}^{3}. Let ρ\rho be the distance to the circle CC. Then for any positive integer kk, there is a 2\mathbb{Z}_{2} harmonic function ff on 3\C\mathbb{R}^{3}\backslash C with monodromy 1-1 around CC such that |f|=O(ρk)|f|=O(\rho^{k}) as ρ0\rho\rightarrow 0.

The proof is given in section 1.

After the digression on the higher-dimensional case, we come back to the two-dimensional case. Suppose KK and D(z)D(z) are given. Consider the following harmonic function in \{O}\mathbb{C}\backslash\{O\}:

h(z)=ln|z|.h(z)=\ln|z|.

By theorem 0.2, there is a unique 2\mathbb{Z}_{2} harmonic function f(z)f(z) (denoted GD(z)G_{D}(z), which depends on D(z)D(z)) and a unique admissible entire function UD(z)U_{D}(z) that corresponds to it. Then they satisfy the following properties:

Theorem 0.7.

The triple (GD(z),UD(z),ln|z|)(G_{D}(z),U_{D}(z),\ln|z|) is “minimal” in the following sense:

  • Any 2\mathbb{Z}_{2} harmonic function ff such that f=o(|z|)f=o(|z|) as |z|+|z|\rightarrow+\infty is a scalar multiplication of GD(z)G_{D}(z) by a real-valued constant.

  • The zero locus of GD(z)G_{D}(z): Z(GD(z)):={z|GD(z)=0}Z(G_{D}(z)):=\{z\in\mathbb{C}~{}|~{}G_{D}(z)=0\} is compact. Moreover, any 2\mathbb{Z}_{2} harmonic function that satisfies this property is a multiplication of GD(z)G_{D}(z) by a real-valued constant.

  • Any admissible entire function u(z)u(z) in \mathbb{C} such that u(z)=o(|z|k)u(z)=o(|z|^{k}) as |z|+|z|\rightarrow+\infty is a scalar multiplication of UD(z)U_{D}(z) by a real-valued constant.

Moreover, we have the following mysterious relationship between UD(z)U_{D}(z) and the polynomial Pell equation:

Theorem 0.8.

If (p(z),q(z))(p(z),q(z)) is a pair of polynomials that satisfies the following polynomial Pell equation:

p(z)2D(z)q(z)2=1,p(z)^{2}-D(z)q(z)^{2}=1,

then p(z)q(z)\dfrac{p^{\prime}(z)}{q(z)} is a multiplication of UD(z)U_{D}(z) up to a multiplication by a real valued constant.

Moreover, a pair (p(z),q(z))(p(z),q(z)) satisfies the polynomial Pell equation if any only if the following algebraic expression of the Abel integration holds:

UD(z)D(z)𝑑z=Cln(p(z)+q(z)D(z)),where C is a real-valued constant.\int\dfrac{U_{D}(z)}{\sqrt{D(z)}}dz=C\ln(p(z)+q(z)\sqrt{D(z)}),~{}~{}~{}\text{where $C$ is a real-valued constant}.
Remark 0.9.
  • Not all D(z)D(z) whose corresponding polynomial Pell equation has a solution. If it has a solution, then D(z)D(z) is called Pellian.

  • We may easily extend the definition of UD(z)U_{D}(z) to the case where D(z)D(z) has roots of multiplicity greater than 11. See definition 2.3.

  • When D(z)D(z) is Pellian, the existence of such a polynomial UD(z)U_{D}(z) that satisfies the properties in theorem 0.8 is known. (See [5] and [6] for example.) However, to the author’s knowledge, its natural relationship with 2\mathbb{Z}_{2} harmonic functions is a new result.

  • Our definition of UD(z)U_{D}(z) (through the correspondence given by theorem 0.2), which still makes sense even when D(z)D(z) is non-Pellian, to the author’s knowledge, is also new.

With the help of our generalized definition of UD(z)U_{D}(z) (which makes sense even if P(z)P(z) is non-Pellian), there is a natural practical method to check whether any given D(z)D(z) is Pellian or not. (Roughly speaking, for any given D(z)D(z), Pellian or not, one may find UD(z)U_{D}(z) 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 2\mathbb{Z}_{2} harmonic functions is new.

There are also some miscellaneous studies on the zero locus of 2\mathbb{Z}_{2} harmonic functions (and their relationship with u(z)u(z)) 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 \mathbb{C} 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 2\mathbb{Z}_{2} harmonic functions

This section mainly proves theorem 0.2, theorem 0.4, and corollary 0.6 in the introduction.

Lemma 1.1.

If ff is a 2\mathbb{Z}_{2} harmonic function on XX, then for each ziKz_{i}\in K, there exists an open neighborhood UiU_{i} of ziz_{i}, and a 2\mathbb{Z}_{2} holomorphic function φi(z)\varphi_{i}(z) on Ui\{zi}U_{i}\backslash\{z_{i}\}, such that

f=Re(φi(z)),forzUi\{zi}f=\text{Re}(\varphi_{i}(z)),~{}~{}~{}~{}\text{for}~{}~{}z\in U_{i}\backslash\{z_{i}\}

where Re is the real part and

φi(z)=ai(zzi)ki+12+O(|zzi|ki+32).\varphi_{i}(z)=a_{i}(z-z_{i})^{k_{i}+\frac{1}{2}}+O(|z-z_{i}|^{k_{i}+\frac{3}{2}}).

We call the half integer ki+12k_{i}+\dfrac{1}{2} the degree of ff at z=ziz=z_{i}, or degzif(z)\mathrm{deg}_{z_{i}}f(z).

Proof.

Suppose zzi=ρeiθz-z_{i}=\rho e^{i\theta}, where (ρ,θ)(\rho,\theta) is the polar coordinates. Take the Fourier series of ff with respect to θ\theta in a small neighborhood of ziz_{i}. Locally ff can be written as

f=Re(m=0+am(ρ)e(2m+1)iθ2),f=\text{Re}(\sum\limits_{m=0}^{+\infty}a_{m}(\rho)e^{\frac{(2m+1)i\theta}{2}}),

where am(ρ)a_{m}(\rho) are complex-valued functions in ρ\rho.

(The Fourier series exists because ff is an ordinary smooth function on the double cover of a neighborhood of PP with the exclusion of PP itself, and one can do the Fourier series there. The integer terms disappear because ff has monodromy 1-1.)

Based on the Laplacian equation,

d2am(ρ)dρ2+1ρdam(ρ)dρ(2m+1)24ρ2am(ρ)=0.\dfrac{d^{2}a_{m}(\rho)}{d\rho^{2}}+\dfrac{1}{\rho}\dfrac{da_{m}(\rho)}{d\rho}-\dfrac{(2m+1)^{2}}{4\rho^{2}}a_{m}(\rho)=0.

So am(ρ)=Amρ2m+12+A~mρ2m+12a_{m}(\rho)=A_{m}\rho^{\frac{2m+1}{2}}+\tilde{A}_{m}\rho^{-\frac{2m+1}{2}}, where AmA_{m} and A~m\tilde{A}_{m} are complex-valued constants. Since |f||\nabla f| has locally finite L2L^{2} norm, A~m\tilde{A}_{m} must be 0. So

f=Re(m=0+Amρ2m+12e(2m+1)iθ2)=Re(n=0+Am(zzi)2m+12).f=\text{Re}(\sum\limits_{m=0}^{+\infty}A_{m}\rho^{\frac{2m+1}{2}}e^{\frac{(2m+1)i\theta}{2}})=\text{Re}(\sum\limits_{n=0}^{+\infty}A_{m}(z-z_{i})^{\frac{2m+1}{2}}).

Let φi(z)=n=0+Am(zzi)2m+12\varphi_{i}(z)=\sum\limits_{n=0}^{+\infty}A_{m}(z-z_{i})^{\frac{2m+1}{2}} and let m=kim=k_{i} be the smallest integer such that Am0.A_{m}\neq 0. Then the proof is finished. ∎

Here are two immediate corollaries.

Corollary 1.2.

If ff is a 2\mathbb{Z}_{2} harmonic function, then |f||f| can be extended to 2\mathbb{R}^{2} continuously by setting |f|=0|f|=0 on KK.

Corollary 1.3.

The expression φi(z)(zzi)ki+12\dfrac{\varphi_{i}(z)}{(z-z_{i})^{k_{i}+\frac{1}{2}}} represents a holomorphic function in UiU_{i} up to a sign. Suppose

D(z)=i=12k(zzi),D(z)=\prod\limits_{i=1}^{2k}(z-z_{i}),

then φi(z)D(z)\dfrac{\varphi_{i}(z)}{\sqrt{D(z)}} represents a holomorphic function in UiU_{i} up to a sign. When ki1k_{i}\geq 1, φi(z)D(z)\dfrac{\varphi_{i}(z)}{\sqrt{D(z)}} has a zero point of degree kik_{i} at z=ziz=z_{i}.

Note that in general, φi(z)D(z)\dfrac{\varphi_{i}(z)}{\sqrt{D(z)}} is not a global holomorphic function.

Theorem 1.4.

There exists an entire function u(z)u(z) on \mathbb{C} such that

df=Re(u(z)dzD(z)).df=\text{Re}(\dfrac{u(z)dz}{\sqrt{D(z)}}).

Moreover, if ki1k_{i}\geq 1, then u(z)u(z) has a zero point at z=ziz=z_{i} of degree kik_{i}.

Proof.

Both dfdf and df*df are 2\mathbb{Z}_{2} 1-forms on 2\K\mathbb{R}^{2}\backslash K. And

df+idf=(fxdx+fydy+ifxdyifydx)df+i*df=(f_{x}dx+f_{y}dy+if_{x}dy-if_{y}dx)
=(fxify)(dx+idy)=(fxify)dz.=(f_{x}-if_{y})(dx+idy)=(f_{x}-if_{y})dz.

Note that

¯(fxify)=12(x+iy)(fxify)=12(fxx+fyy)=0.\bar{\partial}(f_{x}-if_{y})=\dfrac{1}{2}(\partial_{x}+i\partial_{y})(f_{x}-if_{y})=\dfrac{1}{2}(f_{xx}+f_{yy})=0.

So fxifyf_{x}-if_{y} is a 2\mathbb{Z}_{2} holomorphic function on 2\K\mathbb{R}^{2}\backslash K with monodromy 1-1 at KK.

Let u(z)=D(z)(fxify)u(z)=\sqrt{D(z)}(f_{x}-if_{y}). Then u(z)u(z) is an ordinary holomorphic function in 2\K\mathbb{R}^{2}\backslash K. (Strictly speaking, u(z)u(z) is determined only up to a sign here. But if we fix a sign convention for D(z)\sqrt{D(z)} as a section of \mathcal{L}, then u(z)u(z) is unique. We will always assume that this is the case.)

Let φi(z)\varphi_{i}(z) be the function defined in lemma 1.1 and corollary 1.3, then φi(z)D(z)\dfrac{\varphi_{i}(z)}{\sqrt{D(z)}} is holomorphic in a neighborhood of ziz_{i} and

fxify=φi(z)in a neighborhood of zi.f_{x}-if_{y}=\varphi_{i}^{\prime}(z)~{}~{}~{}~{}\text{in a neighborhood of $z_{i}$}.

Thus u(z)=D(z)φi(z)=(φi(z)D(z))D(z)+12(φi(z)D(z))D(z)u(z)=\sqrt{D(z)}\varphi_{i}^{\prime}(z)=(\dfrac{\varphi_{i}(z)}{\sqrt{D(z)}})^{\prime}D(z)+\dfrac{1}{2}(\dfrac{\varphi_{i}(z)}{\sqrt{D(z)}})D^{\prime}(z) is also holomorphic in a neighborhood of ziz_{i}. And u(z)u(z) is an entire function in \mathbb{C}. ∎

Remark 1.5.

If ff is a 2\mathbb{Z}_{2} harmonic function, then f(z1)=0f(z_{1})=0. So we can always write ff as

f(z)=Re(z1zu(z)D(z)𝑑z)f(z)=\text{Re}(\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{D(z)}}dz)

for some entire function u(z)u(z) in \mathbb{C}. However, clearly only admissible entire functions u(z)u(z) in \mathbb{C} make the above expression well-defined 2\mathbb{Z}_{2} harmonic functions.

In general, an integral such as z1zu(z)D(z)𝑑z\displaystyle\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{D(z)}}dz depends on the path chosen in Σ\Sigma from a lifting of z1z_{1} to a lifting of zz, where Σ\Sigma is the surface described by {(w,z)2|w2=D(z)}\{(w,z)\in\mathbb{C}^{2}~{}|~{}w^{2}=D(z)\}, regarded as a double branched cover of \mathbb{C}.

The first homology group of Σ\Sigma is generated by the loops C2,,C2nC_{2},\cdots,C_{2n} as follows:

Choose any path γj\gamma_{j} from z1z_{1} to zjz_{j} in \mathbb{C} that does not meet any other point in KK. Then CjC_{j} is defined to be going along γj\gamma_{j} from z1z_{1} to zjz_{j} in any branch in Σ\Sigma, and go back along γj\gamma_{j} from zjz_{j} to z1z_{1} but in another branch.

In fact, one can visualize the double-branched cover from Σ\Sigma to \mathbb{C} as follows:

[Uncaptioned image]

In this graph, let ι\iota be the rotation of Σ\Sigma along the axis by 180o180^{o}. Then the branched double cover from Σ\Sigma to \mathbb{C} is given by the Σ\Sigma quotient ι\iota. The intersection of Σ\Sigma with the axis is exactly the branching points z1,z2,,z2nz_{1},z_{2},\cdots,z_{2n}. The two punctured points ++\infty and -\infty 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 D(z)\sqrt{D(z)},

Re(z1zju(z)D(z))dz\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{u(z)}{\sqrt{D(z)}})dz

only changes sign.

Now suppose for some chosen paths γj\gamma_{j} from z1z_{1} to zjz_{j}, and for each jj,

Re(z1zju(z)D(z)𝑑z)=Re(γju(z)D(z)𝑑z)=0.\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz)=\text{Re}(\int_{\gamma_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz)=0.

Let CjC_{j} be the loop on Σ\Sigma from z1z_{1} to zjz_{j} along γj\gamma_{j} on the one branch, and then coming back from zjz_{j} to z1z_{1} along γj\gamma_{j} but on the other branch. Then

Re(Cju(z)D(z)𝑑z)=2Re(γju(z)D(z)𝑑z)=0.\text{Re}(\int_{C_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz)=2~{}\text{Re}(\int_{\gamma_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz)=0.

But since C2,C3,,C2nC_{2},C_{3},\cdots,C_{2n} are generators of H1(Σ,)H_{1}(\Sigma,\mathbb{Z}), we know the integral

γu(z)D(z)𝑑z\int_{\gamma}\dfrac{u(z)}{\sqrt{D(z)}}dz

is purely imaginary along any closed loop γ\gamma in Σ\Sigma. So, the integral

z1zju(z)D(z)𝑑z\int_{z_{1}}^{z_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz

has a well-defined real part which does not depend on the path chosen from z1z_{1} to zjz_{j}. 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 u(z)u(z) on \mathbb{C};

  • 2\mathbb{Z}_{2} harmonic functions ff on XX.

The bijection is given by

f(z)=Re(z1zu(z)D(z)𝑑z).()f(z)=\text{Re}(\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{D(z)}}dz).~{}~{}~{}(*)
Proof.
  • f(z)u(z)f(z)\Rightarrow u(z): From theorem 1.4, we know the entire function u(z)u(z) exists. By corollary 1.2,

    0=f(zj)f(z1)=Re(z1zju(z)D(z)𝑑z).0=f(z_{j})-f(z_{1})=\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{u(z)}{\sqrt{D(z)}}dz).
  • u(z)f(z)u(z)\Rightarrow f(z): As shown in the Appendix A, if u(z)u(z) is admissible, then the real part of the integration

    z1zu(z)(D(z))𝑑z\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{(D(z))}}dz

    doesn’t depend on the path chosen from z1z_{1} to zz except its sign (which is also fixed is a sign convention for D(z)\sqrt{D(z)} is chosen). So its real part is a well-defined harmonic function. Call this function ff. We have f(zj)=0f(z_{j})=0 for any j=1,2,,2kj=1,2,\cdots,2k. In order for ff to be a 2\mathbb{Z}_{2} function, we need to check that the monodromy of ff around each point z=zjz=z_{j} is 1-1.

    Near each z=ziz=z_{i}, the Taylor series looks like

    u(z)D(z)=k=0+Ak(zzi)k12.\dfrac{u(z)}{\sqrt{D(z)}}=\sum\limits_{k=0}^{+\infty}A_{k}(z-z_{i})^{k-\frac{1}{2}}.

    So

    f=Re(k=0+Ak(k+12)1(zzi)k+12),f=\text{Re}(\sum\limits_{k=0}^{+\infty}A_{k}(k+\dfrac{1}{2})^{-1}(z-z_{i})^{k+\frac{1}{2}}),

    which indeed has the correct monodromy around z=ziz=z_{i}.

Since u(z)u(z) is uniquely determined by ff, sometimes it is convenient to use the notation uf(z)u_{f}(z) to emphasize its dependence on ff. However, if there is no ambiguity, the subscript ff can be omitted.

Suppose (ρ,θ)(\rho,\theta) are the polar coordinates centered at the origin. Suppose BRB_{R} is a ball with a large radius RR centered at the origin such that KBRK\subset B_{R}. On 2\BR\mathbb{R}^{2}\backslash B_{R}. Using Fourier series,

f=(c0lnρ+Re(n=1+cnρneinθ))+(b0+Re(n=1+bnρneinθ))=U+B,f=(c_{0}\ln\rho+\text{Re}(\sum\limits_{n=1}^{+\infty}c_{n}\rho^{n}e^{in\theta}))+(b_{0}+\text{Re}(\sum\limits_{n=1}^{+\infty}\dfrac{b_{n}}{\rho^{n}}e^{in\theta}))=U+B, (1.8)

where bn,cnb_{n},c_{n} are complex numbers except that b0,c0b_{0},c_{0} must be real. Capital letters UU and BB, representing the first and second parentheses, respectively, are the “unbounded” and “bounded” parts at infinity. We may also call them U(f)U(f) and B(f)B(f) respectively to emphasize their dependency on ff. Note that U(f)U(f) is actually a harmonic function on the whole 2\{O}\mathbb{R}^{2}\backslash\{O\}, where OO is the origin. (But B(f)B(f) is not.)

Theorem 1.9.

Suppose that there are two 2\mathbb{Z}_{2} harmonic functions f1,f2f_{1},f_{2} in XX such that U(f1)=U(f2)U(f_{1})=U(f_{2}), then f1=f2f_{1}=f_{2}. In particular, if U(f)=0U(f)=0, then f=0f=0.

Proof.

Since f=f1f2f=f_{1}-f_{2} is a 2\mathbb{Z}_{2} harmonic function with U(f)=0U(f)=0. So it suffices to prove f=0f=0.

Assume when ρ>R\rho>R,

f=b0+Re(n=1+bnρneinθ).f=b_{0}+\text{Re}(\sum\limits_{n=1}^{+\infty}\dfrac{b_{n}}{\rho^{n}}e^{in\theta}).

We may assume b00b_{0}\geq 0, since otherwise we can change the trivialization of \mathcal{L} and flip the sign of ff. Then b0|f|b_{0}-|f| is a super-harmonic function on the entire 2\mathbb{R}^{2} which approaches 0 at infinity. By maximum principle, it cannot be negative. So in particular, when ρ>R\rho>R, fa0=Re(n=1+anρneinθ)0f-a_{0}=\text{Re}(\sum\limits_{n=1}^{+\infty}\dfrac{a_{n}}{\rho^{n}}e^{in\theta})\geq 0.

Suppose f0f\neq 0. Then let kk is the smallest positive integer such that ak0a_{k}\neq 0. Then

0fa0=Re(akρkeikθ)+O(ρ(k+1)),0\leq f-a_{0}=\text{Re}(\dfrac{a_{k}}{\rho^{k}}e^{ik\theta})+O(\rho^{-(k+1)}),

which is impossible when ρ\rho is large enough.

Corollary 1.10.

If the zero locus of ff is bounded, then U(f)=c0lnρU(f)=c_{0}\ln\rho for some constant c00c_{0}\neq 0. In addition, such ff (if exists) is unique up to a scalar multiplication by a constant.

Proof.

Suppose the zero locus of ff is bounded and in general,

U(f)=c0lnρ+Re(n=1+cnρneinθ).U(f)=c_{0}\ln\rho+\text{Re}(\sum\limits_{n=1}^{+\infty}c_{n}\rho^{n}e^{in\theta}).

We may assume f>0f>0 when ρ>R\rho>R for some large enough RR.

Then U(f)c0lnρU(f)-c_{0}\ln\rho is a harmonic function on the entire 2\mathbb{R}^{2} which is bounded below by c0lnρ-c_{0}\ln\rho when ρ>R\rho>R. It is well known that such a harmonic function with a lower bound must be a constant. But this constant can only be 0 based on the expression of U(f)U(f). So, the only possibility is U(f)=c0lnρU(f)=c_{0}\ln\rho. The uniqueness of ff up to a scalar multiplication follows directly from theorem 1.9. ∎

One may wonder for which choice of U(f)U(f) the 2\mathbb{Z}_{2} harmonic function exists. In fact, we have the following theorem:

Theorem 1.11.

There is a bijection between:

  • The equivalence classes [h][h] of harmonic functions on 2\{O}\mathbb{R}^{2}\backslash\{O\}, where OO is the origin, given by:

    h1h2if and only ifh1h2Cwhen|z|+,h_{1}\sim h_{2}~{}~{}~{}~{}\text{if and only if}~{}~{}~{}h_{1}-h_{2}\rightarrow C~{}~{}\text{when}~{}~{}|z|\rightarrow+\infty,

    where CC is a constant.

  • 2\mathbb{Z}_{2} harmonic functions ff on XX.

The bijection is given by sending ff to U(f)U(f).

We have already shown that the map fU(f)f\rightarrow U(f) 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 z1z_{1} to each zjz_{j} on the double branched cover Σ\Sigma of \mathbb{C} so that the following expression makes sense:

vjk=z1zjzkD(z)𝑑z.v_{jk}=\int_{z_{1}}^{z_{j}}\dfrac{z^{k}}{\sqrt{D(z)}}dz.

And let vk={v2k,v3k,,v2n,k}\vec{v}_{k}=\{v_{2k},v_{3k},\cdots,v_{2n,k}\}. Then the following 2n12n-1 vectors

Re(v0),Im(v0),Re(v1),Im(v1),,Re(vn2),Im(vn2),Im(vn1)\text{Re}(\vec{v}_{0}),\text{Im}(\vec{v}_{0}),\text{Re}(\vec{v}_{1}),\text{Im}(\vec{v}_{1}),\cdots,\text{Re}(\vec{v}_{n-2}),\text{Im}(\vec{v}_{n-2}),\text{Im}(\vec{v}_{n-1})

are linearly independent over \mathbb{R}.

Proof.

Suppose

a0Re(v0)+b0Im(v0)++an2Re(vn2)+bn2Im(vn2)+bn1Im(vn1)=0.a_{0}\text{Re}(\vec{v}_{0})+b_{0}\text{Im}(\vec{v}_{0})+\cdots+a_{n-2}\text{Re}(\vec{v}_{n-2})+b_{n-2}\text{Im}(\vec{v}_{n-2})+b_{n-1}\text{Im}(\vec{v}_{n-1})=0.

Let

c0=a0ib0,c1=a1ib1,,cn2=an2ibn2,cn1=ibn1.c_{0}=a_{0}-ib_{0},c_{1}=a_{1}-ib_{1},\cdots,c_{n-2}=a_{n-2}-ib_{n-2},c_{n-1}=-ib_{n-1}.

And let u(z)=c0+c1z++cn2zn2+cn1zn1u(z)=c_{0}+c_{1}z+\cdots+c_{n-2}z^{n-2}+c_{n-1}z^{n-1}.

Then

Re(z1zjzkD(z)𝑑z)=0,for allj=2,3,,2n.\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{z^{k}}{\sqrt{D(z)}}dz)=0,~{}~{}~{}~{}\text{for all}~{}~{}j=2,3,\cdots,2n.

So by theorem 1.7,

f=Re(z1zu(z)D(z))dzf=\text{Re}(\int_{z_{1}}^{z}\dfrac{u(z)}{\sqrt{D(z)}})dz

defines a 2\mathbb{Z}_{2} harmonic function.

Note that

u(z)D(z)=ibn1z+O(|z|2)\dfrac{u(z)}{\sqrt{D(z)}}=-\dfrac{ib_{n-1}}{z}+O(|z|^{-2})

So

f=Re(ibn1lnz)+C+O(|z|1),f=\text{Re}(-ib_{n-1}\ln z)+C+O(|z|^{-1}),

where CC is a constant. But in order for Re(ibn1lnz)\text{Re}(-ib_{n-1}\ln z) to be well defined, bn1b_{n-1} must be 0. So f=C+O(|z|1)f=C+O(|z|^{-1}). So U(f)=0U(f)=0. Hence f=0f=0, which implies u(z)=0u(z)=0. ∎

Now we prove that fU(f)f\rightarrow U(f) is surjective.

Choose any

U(f)=c0lnρ+Re(n=1+cnρneinθ).U(f)=c_{0}\ln\rho+\text{Re}(\sum\limits_{n=1}^{+\infty}c_{n}\rho^{n}e^{in\theta}).

Consider the following formal Laurent series

D(z)(c0z+n=1+ncnzn1)=k=+Ckzk.\sqrt{D(z)}(\dfrac{c_{0}}{z}+\sum\limits_{n=1}^{+\infty}nc_{n}z^{n-1})=\sum\limits_{k=-\infty}^{+\infty}C_{k}z^{k}.

Let

u0(z)=k=0+Ckzk.u_{0}(z)=\sum\limits_{k=0}^{+\infty}C_{k}z^{k}.

Then u0(z)u_{0}(z) is an entire function in \mathbb{C} such that

|dU(f)Re(d(u0(z)D(z)))|=O(1|z|n)as|z|+.|dU(f)-\text{Re}(d(\dfrac{u_{0}(z)}{\sqrt{D(z)}}))|=O(\dfrac{1}{|z|^{n}})~{}~{}~{}~{}\text{as}~{}~{}|z|\rightarrow+\infty.

By the previous lemma, there exist real numbers a0,b0,a1,b1,an2,bn2,bn1a_{0},b_{0},a_{1},b_{1},\cdots a_{n-2},b_{n-2},b_{n-1} and a corresponding polynomial

u1(z)=(a0ib0)+(z1ib1)z++(an2ibn2)zn2ibn1zn1,u_{1}(z)=(a_{0}-ib_{0})+(z_{1}-ib_{1})z+\cdots+(a_{n-2}-ib_{n-2})z^{n-2}-ib_{n-1}z^{n-1},

such that

Re(z1zju0(z)D(z)𝑑z)=Re(z1zju1(z)D(z)𝑑z),\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{u_{0}(z)}{\sqrt{D(z)}}dz)=\text{Re}(\int_{z_{1}}^{z_{j}}\dfrac{u_{1}(z)}{\sqrt{D(z)}}dz),

for each j=2,3,,2nj=2,3,\cdots,2n.

Thus

f=Re(z1z(u0(z)u1(z))D(z)𝑑z)f=\mathrm{R}e(\int_{z_{1}}^{z}\dfrac{(u_{0}(z)-u_{1}(z))}{\sqrt{D(z)}}dz)

defines a 2\mathbb{Z}_{2} harmonic function.

Since

df=Re(u0(z)u1(z)D(z)dz)=dU(f)+Re(ibn1z)+O(|z|1)when|z|+,df=\text{Re}(\dfrac{u_{0}(z)-u_{1}(z)}{\sqrt{D(z)}}dz)=dU(f)+\text{Re}(\dfrac{ib_{n-1}}{z})+O(|z|^{-1})~{}~{}\text{when}~{}~{}|z|\rightarrow+\infty,

so

f=U(f)+C+O(|z|1)when|z|+,f=U(f)+C+O(|z|^{-1})~{}~{}~{}\text{when}~{}~{}|z|\rightarrow+\infty,

where CC is a constant. This implies that the unbounded part of ff is indeed the proscribed U(f)U(f) 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 X=n\KX=\mathbb{R}^{n}\backslash K, where n3n\geq 3 and KK is a compact smooth manifold of dimension n2n-2 embedded smoothly into n\mathbb{R}^{n}. Let \mathcal{L} be the unique real bundle in XX with monodromy 1-1 around KK. We can similarly define the 2\mathbb{Z}_{2} harmonic functions ff in XX with monodromy 1-1 around KK.

Suppose ff is a 2\mathbb{Z}_{2} harmonic function in U\KU\backslash K with monodromy 1-1 around KK, where KK is a smooth compactly embedded sub-manifold of codimension 22 in n\mathbb{R}^{n}, and UU is a bounded open subset of n\mathbb{R}^{n} that contains KK. Suppose that the L2L^{2} norm of |f||\nabla f| is finite on UU.

In addition, choose a large enough number R0R_{0} and an open ball BR0B_{R_{0}} of radius R0R_{0} centered at the origin such that KBR0K\subset B_{R_{0}}. Fix a trivialization of \mathcal{L} in n\BR0\mathbb{R}^{n}\backslash B_{R_{0}}. Then ff is viewed as an ordinary function on n\BR0\mathbb{R}^{n}\backslash B_{R_{0}}.

Lemma 1.13.

The function |f||f| is a subharmonic function on U\KU\backslash K. That is to say, for any ball BB centered at PP in UU, where PKP\notin K,

|f(P)|1Vol(B)B\K|f|dn𝐱.|f(P)|\leq\dfrac{1}{\text{Vol}(B)}\int_{B\backslash K}|f|d^{n}\mathbf{x}.

Note that we allow BB to have intersection with KK.

Proof.

If B(U\K)B\subset(U\backslash K), then ff is an ordinary harmonic function in BB up to the sign. So, the inequality is true by the mean value property.

In general, suppose BrB_{r} is the ball of radius rr centered at PP. Since PKP\notin K, the inequality is true when rr is small such that BrK=B_{r}\cap K=\emptyset.

Let

φ(r)=1Vol(Br)Br\K|f|dn𝐱.\varphi(r)=\dfrac{1}{\text{Vol}(\partial B_{r})}\int_{\partial B_{r}\backslash K}|f|d^{n}\mathbf{x}.
Vol(Br)φ(r)=limϵ0(Br\KϵΔ(|f|)dn𝐱+KϵBrfn𝑑σ)lim infϵ0KϵBrfn𝑑σ.\text{Vol}(\partial B_{r})\cdot\varphi^{\prime}(r)=\lim\limits_{\epsilon\rightarrow 0}(\int_{B_{r}\backslash K_{\epsilon}}\Delta(|f|)d^{n}\mathbf{x}+\int_{\partial K_{\epsilon}\cap B_{r}}\dfrac{\partial f}{\partial\vec{n}}d\sigma)\geq\liminf\limits_{\epsilon\rightarrow 0}\int_{\partial K_{\epsilon}\cap B_{r}}\dfrac{\partial f}{\partial\vec{n}}d\sigma.

Here

|KϵBrfn𝑑σ|Kϵ|f|𝑑σVoln1(Kϵ)(Kϵ|f|2𝑑σ)12C1(Kϵ|f|2ϵ𝑑σ)12,|\int_{\partial K_{\epsilon}\cap B_{r}}\dfrac{\partial f}{\partial\vec{n}}d\sigma|\leq\int_{\partial K_{\epsilon}}|\nabla f|d\sigma\leq\sqrt{\text{Vol}_{n-1}(\partial K_{\epsilon})}(\int_{\partial K_{\epsilon}}|\nabla f|^{2}d\sigma)^{\frac{1}{2}}\leq C_{1}(\int_{\partial K_{\epsilon}}|\nabla f|^{2}\epsilon d\sigma)^{\frac{1}{2}},

where C1C_{1} is a constant that doesn’t depend on ϵ\epsilon.

On the other hand, fix any small enough δ\delta, there exists a constant CC which doesn’t depend on δ\delta, such that

K2δ\Kδ|f|2dn𝐱Cδ2δKρ|f|2𝑑σ𝑑ρC2infδ<ϵ<2δKϵ|f|2ϵ𝑑σ.\int_{K_{2\delta}\backslash K_{\delta}}|\nabla f|^{2}d^{n}\mathbf{x}\geq C\int_{\delta}^{2\delta}\int_{\partial K_{\rho}}|\nabla f|^{2}d\sigma d\rho\geq\dfrac{C}{2}\inf\limits_{\delta<\epsilon<2\delta}\int_{\partial K_{\epsilon}}|\nabla f|^{2}\epsilon d\sigma.

Note that |f||\nabla f| has locally finite L2L^{2} norm. The above inequality implies that

lim infϵ0Kϵ|f|2ϵ𝑑σlim infδ0K2δ\Kδ|f|2dn𝐱=0.\liminf\limits_{\epsilon\rightarrow 0}\int_{\partial K_{\epsilon}}|\nabla f|^{2}\epsilon d\sigma\leq\liminf\limits_{\delta\rightarrow 0}\int_{K_{2\delta}\backslash K_{\delta}}|\nabla f|^{2}d^{n}\mathbf{x}=0.

So

lim infϵ0KϵBrfn𝑑σ=0.\liminf\limits_{\epsilon\rightarrow 0}\int_{\partial K_{\epsilon}\cap B_{r}}\dfrac{\partial f}{\partial\vec{n}}d\sigma=0.

So φ(r)0\varphi^{\prime}(r)\geq 0. And |f(P)|φ(r)|f(P)|\leq\varphi(r) for any r>0r>0 such that BrUB_{r}\subset U. ∎

Corollary 1.14.

If we extend |f||f| by setting |f|=0|f|=0 on KK, then |f||f| is a sub-harmonic function on the entire UU.

Note that we haven’t proved that the extension is continuous yet.

Lemma 1.15.

Suppose KϵK_{\epsilon} is a small tubular neighborhood of KK which is a D(ϵ)D(\epsilon) bundle over KK, where D(ϵ)D(\epsilon) is a disk of radius ϵ\epsilon whose local polar coordinates are (ρ,θ)(\rho,\theta). (Note that θ\theta may not be well-defined globally along KK). For a small contractible relatively open subset UU of KK, let UϵU_{\epsilon} be the ϵ\epsilon tubular neighborhood of UU (which is roughly U×D(ϵ)U\times D(\epsilon) with minor modifications on the metric). Then there exists a constant CC which doesn’t depend on ϵ\epsilon, but may depend on UU, such that

1Vol(Uϵ)Uϵ|f|dn𝐱CfL2(Uϵ).\dfrac{1}{\text{Vol}(U_{\epsilon})}\int_{U_{\epsilon}}|f|d^{n}\mathbf{x}\leq C||\nabla f||_{L^{2}(U_{\epsilon})}.

Here Vol(Uϵ)\text{Vol}(U_{\epsilon}) volume of UϵU_{\epsilon}.

In particular

limϵ01Vol(Uϵ)Uϵ|f|dn𝐱=0.\lim\limits_{\epsilon\rightarrow 0}\dfrac{1}{\text{Vol}(U_{\epsilon})}\int_{U_{\epsilon}}|f|d^{n}\mathbf{x}=0.
Proof.

Because ff has modronomy 1-1 around KK, we know that for any θ0\theta_{0},

2|f|02π|θf|𝑑θ.2|f|\leq\int_{0}^{2\pi}|\partial_{\theta}f|d\theta.
12π02π|f|𝑑θ1202π|θf|𝑑θC02π|f|ρ𝑑θ,\dfrac{1}{2\pi}\int_{0}^{2\pi}|f|d\theta\leq\dfrac{1}{2}\int_{0}^{2\pi}|\partial_{\theta}f|d\theta\leq C\int_{0}^{2\pi}|\nabla f|\rho d\theta,

where CC is a constant. So there is a constant CC^{\prime} such that

Uϵ|f|dn𝐱CUϵ|f|ρdn𝐱Cϵ(Vol(Uϵ))fL2(Uϵ).\int_{U_{\epsilon}}|f|d^{n}\mathbf{x}\leq C^{\prime}\int_{U_{\epsilon}}|\nabla f|\rho d^{n}\mathbf{x}\leq C^{\prime}\epsilon(\sqrt{\text{Vol}(U_{\epsilon})})||\nabla f||_{L^{2}(U_{\epsilon})}.

And there is another constant C1C_{1} such that

CϵC1Vol(Uϵ).C^{\prime}\epsilon\leq C_{1}\sqrt{\text{Vol}(U_{\epsilon})}.

So

1Vol(Uϵ)Uϵ|f|dn𝐱C1fL2(Uϵ).\dfrac{1}{\text{Vol}(U_{\epsilon})}\int_{U_{\epsilon}}|f|d^{n}\mathbf{x}\leq C_{1}||\nabla f||_{L^{2}(U_{\epsilon})}.

The final assertion in the lemma is because

limϵ0fL2(Uϵ)=0.\lim\limits_{\epsilon\rightarrow 0}||\nabla f||_{L^{2}(U_{\epsilon})}=0.

Lemma 1.16.

Let C(δ)C(\delta) be a subset of KK such that for each PC(δ)P\in C(\delta), there exists a ball BB centered at PP such that

1Vol(B)B|f|dn𝐱δ.\dfrac{1}{\text{Vol(B)}}\int_{B}|f|d^{n}\mathbf{x}\leq\delta.

Then C(δ)=KC(\delta)=K.

Proof.

Clearly C(δ)C(\delta) is a closed subset of KK. It suffices to prove that C(δ)C(\delta) is dense in KK.

For any small contractible relatively open subset UU of KK, we know from the previous lemma that

1Vol(Uϵ)Uϵ|f|dn𝐱CfL2(Uϵ).\dfrac{1}{\text{Vol}(U_{\epsilon})}\int_{U_{\epsilon}}|f|d^{n}\mathbf{x}\leq C||\nabla f||_{L^{2}(U_{\epsilon})}.

But

limϵ0fL2(Uϵ)=0.\lim\limits_{\epsilon\rightarrow 0}||\nabla f||_{L^{2}(U_{\epsilon})}=0.

So the “average value” of |f||f| on UϵU_{\epsilon} approaches 0 as ϵ\epsilon goes to 0. This implies that there exists at least one point PUP\in U such that PP is also in C(δ)C(\delta). Since this is true for any small contractible relative open subset of KK, we know that C(δ)C(\delta) is dense. ∎

Theorem 1.17.

The function |f||f| can be extended continuously in all UU setting |f|=0|f|=0 in KK. This extended version of |f||f| is a subharmonic function in UU.

Proof.

We have already proved that the extended version of |f||f| is subharmonic. So, it suffices to prove that it is continuous at any point PP in KK.

From the previous lemma, for any small enough δ>0\delta>0, PC(δ)P\in C(\delta). So there exists a ball BB centered at PP such that

1Vol(B)B|f|dn𝐱δ.\dfrac{1}{\text{Vol}(B)}\int_{B}|f|d^{n}\mathbf{x}\leq\delta.

Let B0B_{0} be the ball centered at PP whose radius is half of the radius of BB. For any PB0P^{\prime}\in B_{0}, let B(P)B(P^{\prime}) be the ball centered at PP^{\prime} whose radius is also half of the radius of BB. Then B(P)BB(P^{\prime})\subset B. And Vol(B(P))=12nVol(B)\text{Vol}(B(P^{\prime}))=\dfrac{1}{2^{n}}\text{Vol}(B). So

1Vol(B(P))B(P)|f|dn𝐱1Vol(B(P))B|f|dn𝐱2nδ.\dfrac{1}{\text{Vol}(B(P^{\prime}))}\int_{B(P^{\prime})}|f|d^{n}\mathbf{x}\leq\dfrac{1}{\text{Vol}(B(P^{\prime}))}\int_{B}|f|d^{n}\mathbf{x}\leq 2^{n}\delta.

Since |f||f| is subharmonic, for any PB0P^{\prime}\in B_{0},

|f(P)|2nδ.|f(P^{\prime})|\leq 2^{n}\delta.

Letting δ0\delta\rightarrow 0 implies that

lim supPP|f(P)|=|f(P)|=0.\limsup\limits_{P^{\prime}\rightarrow P}|f(P^{\prime})|=|f(P)|=0.

Thus |f||f| is continuous at PP.

Before we prove theorem 0.4, we introduce an inspiring special case of theorem 0.4 due to Taubes:

Corollary 1.18.

There is a unique 2\mathbb{Z}_{2} harmonic function on n\K\mathbb{R}^{n}\backslash K such that

f10when|𝐱|+.f-1\rightarrow 0~{}~{}~{}~{}\text{when}~{}~{}~{}|\mathbf{x}|\rightarrow+\infty.

Taubes’ proof of corollary 1.18

Consider all smooth 2\mathbb{Z}_{2} function that equals 11 when 𝐱\mathbf{x} is large enough. Let the Dirichlet energy be

(f)=X|df|2𝑑𝐱.\mathcal{E}(f)=\int_{X}|df|^{2}d\mathbf{x}.

Consider the following space:

Suppose ff is a smooth 2\mathbb{Z}_{2} function such that 1|f|=01-|f|=0 far away from the origin and that |f||f| extends continuously to KK by setting |f|=0|f|=0 on KK. Then we define the MM “norm” of ff as its Dirichlet energy:

fM2=X|df|2𝑑𝐱.||f||^{2}_{M}=\int_{X}|df|^{2}d\mathbf{x}.

The space of completion of the above type of smooth 2\mathbb{Z}_{2} functions under the MM norm is still denoted as MM.

Note that if fMf\in M, then by Sobolev embedding, the L2nn2(X)L^{\frac{2n}{n-2}}(X) norm of 1|f|1-|f| is bounded up to a constant by fM||f||_{M}. As a corollary, ff also has finite L2(B)L^{2}(B) norm for any ball BB whose closure is in XX.

Choose a sequence of smooth functions fnMf_{n}\in M whose Dirichlet energy approaches the infimum in MM. One may assume, by possibly choosing a subsequence, that

  • fnf_{n} weakly converges to ff in MM.

  • 1|fn|1-|f_{n}| weakly converges to 1|f|1-|f| in L2nn2(X)L^{\frac{2n}{n-2}}(X).

The second bullet implies fMf\in M. The first bullets implies

lim infn+fnM=infgMCgMfM.\liminf\limits_{n\rightarrow+\infty}||f_{n}||_{M}=\inf\limits_{g\in M_{C}}||g||_{M}\geq||f||_{M}.

Since fMf\in M, by the way that the sequence fnf_{n} was constructed, ff achieves the minimal Dirichlet energy in MM.

By the method of variation, under any local trivialization of \mathcal{L}, one sees that ff is a weak solution of the Laplacian equation on any small ball BB whose closure is in XX. Since ff is also in L2(B)L^{2}(B), by a standard elliptic regularity argument, ff is harmonic on BB. Thus ff is a 2\mathbb{Z}_{2} harmonic function on the entire XX.

Now we return to theorem 0.4. We first need a lemma.

Lemma 1.19.

Let UU be a connected, bounded open subset of n\mathbb{R}^{n} such that KUK\subset U. Suppose that the boundary U\partial U is smooth. Suppose that a trivialization of \mathcal{L} on U\partial U is given. Suppose ff is a smooth function on U¯\bar{U}. Then there exists a unique 2\mathbb{Z}_{2} harmonic function uu on U¯\K\bar{U}\backslash K (that is to say, harmonic on U\KU\backslash K and continuous on U¯\K\bar{U}\backslash K) with monodromy 1-1 around KK (we just call it 2\mathbb{Z}_{2} harmonic function when there is no ambiguity) such that

v=fon U.v=f~{}~{}~{}\text{on $\partial U$}.
Proof.

Consider all smooth 2\mathbb{Z}_{2} functions vv on U\KU\backslash K that equals ff on the boundary.

Consider the energy function

EU(v)=U|v|2dn𝐱.E_{U}(v)=\int_{U}|\nabla v|^{2}d^{n}\mathbf{x}.

The energy function is bounded above by the H1(U)H^{1}(U) norm. By Poincare inequality, the H1(U)H^{1}(U) norm is also bounded above by a constant times EU(v)E_{U}(v) plus a constant which only depends on UU and ff. Thus we can choose a sequence of 2\mathbb{Z}_{2} smooth functions vnv_{n} on U¯\K\bar{U}\backslash K that converges to the infimum of the energy among all such 2\mathbb{Z}_{2} functions and weakly converges to a weak solution of the Laplacian equation in H1(BR)H^{1}(B_{R}), which by choosing a further sub-sequence strongly converges in L2(BR)L^{2}(B_{R}). By a standard regularity theorem, this converging 2\mathbb{Z}_{2} function as a weak solution of the Laplacian equation is a 2\mathbb{Z}_{2} harmonic function.

Suppose there are two 2\mathbb{Z}_{2} harmonic functions on U¯\K\bar{U}\backslash K which has the same boundary values on U\partial U. Then their difference v~\tilde{v} is a 2\mathbb{Z}_{2} harmonic function which equals 0 on U\partial U. But if v~0\tilde{v}\neq 0, then let V0V_{0} be any non-empty open connected component of U\(KZ)U\backslash(K\cup Z), where ZZ is the zero locus of v~\tilde{v}. Then such V0V_{0} exists and |v~||\tilde{v}| is an ordinary non-zero harmonic function on V0V_{0} whose boundary equals zero. This violates the maximal principle. Hence, the 2\mathbb{Z}_{2} harmonic function is unique. ∎

Proof of theorem 0.4.

  • hfh\Rightarrow f

Let fRf_{R} be the 2\mathbb{Z}_{2} harmonic function on BR\KB_{R}\backslash K defined in the previous lemma. We may extend fRf_{R} continuously to the entire XX by setting fR=hf_{R}=h outside of BRB_{R}, still denoted as fRf_{R}. Suppose R0R1RR2R_{0}\ll R_{1}\ll R\ll R_{2}. Let ERE_{R} be the short for EBRE_{B_{R}} which is the energy defined in the previous lemma.

Since fR=hf_{R}=h outside of BRB_{R}, we have

ER2(fR)ER2(h)=ER(fR)ER(h).E_{R_{2}}(f_{R})-E_{R_{2}}(h)=E_{R}(f_{R})-E_{R}(h).

Let E(fR)=limR2+(ER2(fR)ER2(h))=ER(fR)ER(h).E(f_{R})=\lim\limits_{R_{2}\rightarrow+\infty}(E_{R_{2}}(f_{R})-E_{R_{2}}(h))=E_{R}(f_{R})-E_{R}(h).

Note that by the way fRf_{R} was chosen, fRf_{R} has the minimal ERE_{R} energy among all 2\mathbb{Z}_{2} functions in BR\KB_{R}\backslash K equal to uu in BR\partial B_{R}. And fR1f_{R_{1}} is an example of such 2\mathbb{Z}_{2} function. Thus, ER(fR)ER(fR1).E_{R}(f_{R})\leq E_{R}(f_{R_{1}}).

So E(fR)E(fR1)E(f_{R})\leq E(f_{R_{1}}). That is, E(fR)E(f_{R}) does not increase in RR. Since it is bounded below by 0, it must converge. We may assume E=limR+E(fR)E=\lim\limits_{R\rightarrow+\infty}E(f_{R}).

Suppose R1R_{1} is large enough such that E(fR1)E<ϵE(f_{R_{1}})-E<\epsilon. Then for any R>R1R>R_{1}, E(fR1)E(fR)<ϵE(f_{R_{1}})-E(f_{R})<\epsilon. In particular, |ER(fR)ER(fR1)|<ϵ|E_{R}(f_{R})-E_{R}(f_{R_{1}})|<\epsilon. On the other hand, fRf_{R} has the minimal ERE_{R} energy among all 2\mathbb{Z}_{2} harmonic functions on BR\KB_{R}\backslash K that is equal to uu on BR\partial B_{R}. And fR+fR12\dfrac{f_{R}+f_{R_{1}}}{2} is one of such kind of 2\mathbb{Z}_{2} functions. So

ER(fR+fR1)=4ER(fR+fR12)4ER(fR).E_{R}(f_{R}+f_{R_{1}})=4E_{R}(\dfrac{f_{R}+f_{R_{1}}}{2})\geq 4E_{R}(f_{R}).

And

ER(fRfR1)=2(ER(fR))+ER(fR1))ER(fR+fR1)<2ϵ.E_{R}(f_{R}-f_{R_{1}})=2(E_{R}(f_{R}))+E_{R}(f_{R_{1}}))-E_{R}(f_{R}+f_{R_{1}})<2\epsilon.

Note that fRfR1f_{R}-f_{R_{1}} has bounded support in n\K\mathbb{R}^{n}\backslash K. By Sobolev embedding, there exists a constant CC which does not depend on RR such that

fRfR1L2nn2(X)Cϵ.||f_{R}-f_{R_{1}}||_{L^{\frac{2n}{n-2}}(X)}\leq C\epsilon.

This implies that fRfR1f_{R}-f_{R_{1}} is a Cauchy sequence (as RR goes to infinity) in L2nn2(X)L^{\frac{2n}{n-2}}(X). So, it converges to some 2\mathbb{Z}_{2} function ffR1f-f_{R_{1}} as R+R\rightarrow+\infty. This function ff is locally a weak solution of the Laplacian equation and locally in L2L^{2}, so it is a 2\mathbb{Z}_{2} harmonic function.

On the other hand,

fRhL2nn2(n\BR1)Cϵ.||f_{R}-h||_{L^{\frac{2n}{n-2}}(\mathbb{R}^{n}\backslash B_{R_{1}})}\leq C\epsilon.

So

fhL2nn2(n\BR1)2Cϵ.||f-h||_{L^{\frac{2n}{n-2}}(\mathbb{R}^{n}\backslash B_{R_{1}})}\leq 2C\epsilon.

So for any ball BB of radius 22 in n\BR1\mathbb{R}^{n}\backslash B_{R_{1}},

fhL2(B)C1ϵ,||f-h||_{L^{2}(B)}\leq C_{1}\epsilon,

where C1C_{1} is a large constant.

But fhf-h is a harmonic function in BB. So, by a regularity argument, assuming that BB^{\prime} is the ball of radius 11 with the same center as BB, we have another constant C2C_{2} such that

fhL(B)C2ϵ.||f-h||_{L^{\infty}(B^{\prime})}\leq C_{2}\epsilon.

Note that C2C_{2} does not depend on the particular ball BB chosen. So, in fact, for some R2R1R_{2}\gg R_{1}, we have

fhL(n\BR2)C2ϵ.||f-h||_{L^{\infty}(\mathbb{R}^{n}\backslash B_{R_{2}})}\leq C_{2}\epsilon.

Letting ϵ0\epsilon\rightarrow 0 implies that ff converges to hh uniformly as R+R\rightarrow+\infty.

  • fhf\Rightarrow h

We may use the same argument as in the hfh\Rightarrow f case. However, we introduce a different argument here.

In the spherical coordinates, the Laplacian equation is written as

Δf=r2f+(n1)rrf+1r2ΔSn1f=0.\Delta f=\partial^{2}_{r}f+\dfrac{(n-1)}{r}\partial_{r}f+\dfrac{1}{r^{2}}\Delta_{S^{n-1}}f=0.

And when RR is large enough,

f=A0+B0R(n2)+m=1+AmRmΦm+m=1+BmR(m+n2)Ψm,f=A_{0}+B_{0}R^{-(n-2)}+\sum\limits_{m=1}^{+\infty}A_{m}R^{m}\Phi_{m}+\sum\limits_{m=1}^{+\infty}B_{m}R^{-(m+n-2)}\Psi_{m},

where Φm\Phi_{m} and Ψm\Psi_{m} are spherical harmonics with eigenvalue (m+n2)m(m+n-2)m, AmA_{m} and BmB_{m} are constants.

Let

h=A0+m=1+AmRmΦm.h=A_{0}+\sum\limits_{m=1}^{+\infty}A_{m}R^{m}\Phi_{m}.

Then hh is the harmonic function on n\mathbb{R}^{n} that we want.

Remark 1.20.

Like in the case 2\mathbb{R}^{2}, one sees from the proof that the 2\mathbb{Z}_{2} function uu depends on hh and KK continuously in a suitable sense.

Now we prove the corollary 0.6. We may assume that our S1S^{1} embeds in 3\mathbb{R}^{3} as follows:

S1={(coss,sins,0)|s[0,2π)}.S^{1}=\{(\cos s,\sin s,0)~{}|~{}s\in[0,2\pi)\}.

Clearly there is an S1S^{1} action of 3\mathbb{R}^{3} that restricts to a rotation on the embedded S1S^{1}.

We may also choose local coordinates (s,ρ,θ)(s,\rho,\theta) in a neighborhood of S1S^{1} to represent the point: ((coss)(1+ρcosθ),(sins)(1+ρcosθ),ρsinθ)((\cos s)\cdot(1+\rho\cos\theta),(\sin s)\cdot(1+\rho\cos\theta),\rho\sin\theta).

Lemma 1.21.

There is a sequence of linearly independent harmonic functions in 3\mathbb{R}^{3} that are invariant under the S1S^{1} action.

Proof.

In fact, these harmonic functions can be chosen to have polynomial growth. Otherwise, there are even uncountable such harmonic functions.

Examples are: f=|x|l(l+1)Pl(coss)f=|x|^{l(l+1)}P_{l}(\cos s), where |x|=x12+x22+x32|x|=\sqrt{x_{1}^{2}+x_{2}^{2}+x_{3}^{2}} is the distance to the origin, and PlP_{l} are Lendgre polynomials. It is well known that Pl(coss)P_{l}(\cos s) are spherical harmonics, and hence ff 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 2\mathbb{Z}_{2} harmonic functions in 2\S1\mathbb{R}^{2}\backslash S^{1} with monodromy 1-1 around S1S^{1} that are invariant under the S1S^{1} action.

Here is a sketch of the proof of corollary 0.6:

Proof.

In the local coordinates,

Δf=0ρ(ρ(1+ρcosθ)1ρf)+θ(ρ1(1+ρcosθ)θf)+s(ρ(1+ρcosθ)sf)=0.\Delta f=0\iff\partial_{\rho}(\rho(1+\rho\cos\theta)^{-1}\partial_{\rho}f)+\partial_{\theta}(\rho^{-1}(1+\rho\cos\theta)\partial_{\theta}f)+\partial_{s}(\rho(1+\rho\cos\theta)\partial_{s}f)=0.

Suppose ff is invariant under S1S^{1} action, that is to say, independent with ss. Then sf=0\partial_{s}f=0 and

ρ2f+1ρρf+1ρ2θ2f+(higher order terms)=0.\partial_{\rho}^{2}f+\dfrac{1}{\rho}\partial_{\rho}f+\dfrac{1}{\rho^{2}}\partial_{\theta}^{2}f+(\text{higher order terms})=0.

We can construct the Fourier series of ff as in the proof of lemma 1.1.
Suppose

f=Re(m=0+am(ρ)e(2m+1)iθ2),f=\text{Re}(\sum\limits_{m=0}^{+\infty}a_{m}(\rho)e^{\frac{(2m+1)i\theta}{2}}),

where am(ρ)a_{m}(\rho) are complex-valued functions in ρ\rho.

am′′(ρ)+1ρam(ρ)(2m+1)24ρ2am(ρ)+(higher order terms)=0.a_{m}^{\prime\prime}(\rho)+\dfrac{1}{\rho}a_{m}^{\prime}(\rho)-\dfrac{(2m+1)^{2}}{4\rho^{2}}a_{m}(\rho)+(\text{higher order terms})=0.

From corollary 1.22, there are infinitely many choices of linearly independent 2\mathbb{Z}_{2} harmonic functions ff that clearly satisfy |f|=O(ρ1)|f|=O(\rho^{-1}). We use induction on the power.

By induction, suppose that there are infinitely many choices of linearly independent ff such that f=O(ρm)f=O(\rho^{m}) as ρ0\rho\rightarrow 0. All such ff form an infinite-dimensional vector space VmV_{m}. For each fVmf\in V_{m}, we have ak(ρ)=O(ρm)a_{k}(\rho)=O(\rho^{m}) for all kk. From the standard regular singular point theory in ODE, we know that ak(ρ)=O(ρm+1)a_{k}(\rho)=O(\rho^{m+1}) for all kmk\neq m and that am(ρ)=Aρm+12+O(ρm+1)a_{m}(\rho)=A\rho^{m+\frac{1}{2}}+O(\rho^{m+1}). The map from VmV_{m} to \mathbb{R} sending ff to AA has an infinite-dimensional kernel. Clearly, any 2\mathbb{Z}_{2} harmonic function in the kernel behaves like O(ρm+1)O(\rho^{m+1}) as ρ0\rho\rightarrow 0. ∎

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 ρ0\rho\rightarrow 0. Moreover, this boundary behavior is uniform for different Fourier components am(ρ)a_{m}(\rho). However, these are all easy to check and are omitted here. In fact, there are many much stronger conclusions about what a 2\mathbb{Z}_{2} 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 D(z)D(z) is Pellian or not. We work in the two-dimensional case. We assume that KK and D(z)D(z) are given.

By theorem 0.2, there exists a unique 2\mathbb{Z}_{2} harmonic function, denoted as GD(z)G_{D}(z) (or G(z)G(z) for short), such that the unbounded part at infinity U(G)U(G) has only a term ln|z|\ln|z|. By corollary 1.10, it is also the unique 2\mathbb{Z}_{2} 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 G(z)G(z) is due to the following observation of Taubes: if one uses a conformal map to send \mathbb{C} to 1\{S}\mathbb{C}\mathbb{P}^{1}\backslash\{S\}, where SS is the south pole (corresponding to the point of infinity), then a 2\mathbb{Z}_{2} harmonic function on \mathbb{C} becomes a 2\mathbb{Z}_{2} harmonic function on 1\{S}\mathbb{C}\mathbb{P}^{1}\backslash\{S\}. And G(z)G(z) becomes the 2\mathbb{Z}_{2} version of the Green function on 1\mathbb{C}\mathbb{P}^{1} evaluated at SS. More details of this Green function can be found implicitly in [1] [2] or [3].

Theorem 2.2.

When f(z)=G(z)f(z)=G(z), the entire function corresponding to u(z)u(z) in theorem 1.4 is a polynomial of degree n1n-1, and vice versa (up to a multiplication of a constant of real value). We call this polynomial U(z)U(z). Then U(z)U(z) is uniquely determined by D(z)D(z) (up to a multiplication by a constant of real value as always).

Proof.

Once the first statement is proved, the uniqueness is a consequence of theorem 1.9 and corollary 1.10. The first statement (to be more precise, the first sentence in the above statement of the theorem) is a special case of the upcoming theorem 2.4 whose proof is deferred until there. ∎

Definition 2.3.

For each even degree polynomial D(z)D(z) (may not have distinct roots), we define a polynomial UD(z)U_{D}(z) (up to a multiplication by a real number) in the following way:

  • If all roots of D(z)D(z) are distinct, then UD(z)U_{D}(z) is the unique polynomial U(z)U(z) from the previous theorem.

  • Suppose D(z)=D1(z)2D0(z)D(z)=D_{1}(z)^{2}D_{0}(z), such that all roots of D0(z)D_{0}(z) are distinct. Then UD(z)=D1(z)UD0(z)U_{D}(z)=D_{1}(z)U_{D_{0}}(z).

If there is no ambiguity, we may omit the subscript DD and simply write it as U(z)U(z).

Theorem 2.4.

Suppose ff is a 2\mathbb{Z}_{2} harmonic function. Suppose the unbounded part of ff at infinity is

U(f)=c0lnρ+Re(n=1+cnρneinθ).U(f)=c_{0}\ln\rho+Re(\sum\limits_{n=1}^{+\infty}c_{n}\rho^{n}e^{in\theta}).

Then u(z)u(z) is a polynomial of degree k+n1k+n-1 if and only if kk is the largest integer such that ckc_{k} is not zero.

Proof.

In fact, when ρ\rho is large enough, we have expression (1.8)

f=(c0lnρ+Re(n=1+cnρneinθ))+(b0+Re(n=1+bnρneinθ))f=(c_{0}\ln\rho+\text{Re}(\sum\limits_{n=1}^{+\infty}c_{n}\rho^{n}e^{in\theta}))+(b_{0}+\text{Re}(\sum\limits_{n=1}^{+\infty}\dfrac{b_{n}}{\rho^{n}}e^{in\theta}))
=b0+Re(c0lnz+n=1+(cnzn+bnzn)).=b_{0}+\text{Re}(c_{0}\ln z+\sum\limits_{n=1}^{+\infty}(c_{n}z^{n}+b_{n}z^{-n})).

So

df=Re(u(z)dzD(z))=Re(c0zdz+n=1+(ncnzn1nbnzn1)dz).df=\text{Re}(\dfrac{u(z)dz}{\sqrt{D(z)}})=\text{Re}(\dfrac{c_{0}}{z}dz+\sum\limits_{n=1}^{+\infty}(nc_{n}z^{n-1}-nb_{n}z^{-n-1})dz).

So

u(z)=D(z)(c0z+n=1+(ncnzn1nbnzn1)).u(z)=\sqrt{D(z)}(\dfrac{c_{0}}{z}+\sum\limits_{n=1}^{+\infty}(nc_{n}z^{n-1}-nb_{n}z^{-n-1})).

Therefore, u(z)u(z) has a pole of order n+k1n+k-1 at infinity if and only if kk is the largest integer such that ckc_{k} is non-zero. But u(z)u(z) is an entire function. So, it has a pole of order n+k1n+k-1 at infinity if and only if it is a polynomial of degree n+k1n+k-1. ∎

Note that there is a corollary: the leading coefficient of U(z)U(z) is real.

Remark: The 2\mathbb{Z}_{2} harmonic function ff depends on [h][h] and D(z)D(z) continuously in a suitable sense. In fact, in a suitable sense (for example, the C0(B)C^{0}(B) topology, where BB is a large enough disk that contains all the roots of D(z)D(z)), the following expressions for D(z)D(z):

z1zjzkD(z)𝑑z\int_{z_{1}}^{z_{j}}\dfrac{z^{k}}{\sqrt{D(z)}}dz

and the following expressions for [h][h]

z1zju0(z)D(z)𝑑z\int_{z_{1}}^{z_{j}}\dfrac{u_{0}(z)}{\sqrt{D(z)}}dz

are all continuous, where u0(z)u_{0}(z) is defined in the proof of theorem 0.4.

Recall that suppose D(z)D(z) is a polynomial with complex coefficients. The following equation is called the polynomial Pell equation, or the Pell equation for short:

p(z)2q(z)2D(z)=1,p(z)^{2}-q(z)^{2}D(z)=1,

where p(z)p(z) and q(z)q(z) are unknown polynomials with complex coefficients.

If the Pell equation for D(z)D(z) has a nontrivial solution (that is, q(z)0q(z)\neq 0), then D(z)D(z) is called Pellian. Ohterwise, it is called non-Pellian.

One obvious observation is that, D(z)D(z) is non-Pellian when the degree is odd. And when the degree of D(z)D(z) is 22, it is Pellian if and only if it is square free. It is believed that when the degree of D(z)D(z) is at least 44, 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 D(z)D(z) 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 (p1(z),q1(z))(p_{1}(z),q_{1}(z)) is a solution to the Pell equation for D(z)D(z) with minimal degree. Let

pk(z)+qk(z)D(z)=(p1(z)+q1(z)D(z))k.p_{k}(z)+q_{k}(z)\sqrt{D(z)}=(p_{1}(z)+q_{1}(z)\sqrt{D(z)})^{k}.

Then (±pk,±qk)(\pm p_{k},\pm q_{k}) are all the solutions to the Pell equation for D(z)D(z).

Corollary 2.6.

Suppose (p,q)(p,q) is a solution to the Pell equation for D(z)D(z). Let U~(z)=p(z)q(z)\tilde{U}(z)=\dfrac{p^{\prime}(z)}{q(z)}. Then different solutions will give the same U~(z)\tilde{U}(z) up to a multiplication by a real-valued constant.

Corollary 2.7.

If D(z)D(z) is Pellian and if (p(z),q(z))(p(z),q(z)) is a solution to the Pell equation. Let U~(z)=p(z)q(z)\tilde{U}(z)=\dfrac{p^{\prime}(z)}{q(z)}. Then U~(z)\tilde{U}(z) is a polynomial and the following Abel integration has an algebraic expression:

U~(z)D(z)𝑑z=ln(p(z)+q(z)D(z)).\int\dfrac{\tilde{U}(z)}{\sqrt{D(z)}}dz=\ln(p(z)+q(z)\sqrt{D(z)}).

Thus in order to prove theorem 0.8, the only task is to verify that U~(z)\tilde{U}(z) given above is the same as our UD(z)U_{D}(z) 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 Re(ln(p(z)+q(z)D(z)))\text{Re}(\ln(p(z)+q(z)\sqrt{D(z)})) defines a 2\mathbb{Z}_{2} harmonic function. Then by uniqueness given by theorem 0.2, UD(z)U_{D}(z) must be the same as U~(z)\tilde{U}(z).

First of all,

(p(z)+q(z)D(z))(p(z)q(z)D(z))=1.(p(z)+q(z)\sqrt{D(z)})(p(z)-q(z)\sqrt{D(z)})=1.

So

p(z)+q(z)D(z)0for allz.p(z)+q(z)\sqrt{D(z)}\neq 0~{}~{}~{}\text{for all}~{}~{}~{}z\in\mathbb{C}.

Thus ln(p(z)+q(z)D(z))\ln(p(z)+q(z)\sqrt{D(z)}) has a well-defined real part on the entire \mathbb{C} except a sign ambiguity for D(z)\sqrt{D(z)}. But a different choice of the sign of D(z)\sqrt{D(z)} also changes only the sign of Re(ln(p(z)+q(z)D(z)))\text{Re}(\ln(p(z)+q(z)\sqrt{D(z)})). So, this sign ambiguity corresponds to the correct monodromy for a 2\mathbb{Z}_{2} function. Thus, Re(ln(p(z)+q(z)D(z)))\text{Re}(\ln(p(z)+q(z)\sqrt{D(z)})) is a well-defined 2\mathbb{Z}_{2} harmonic function on XX.

Note that this above theorem partially explains why we choose to have the second bullet of the definition 2.3. The author wonders whether U(z)U(z) can be computed from D(z)D(z) in a purely algebraic way (that is, not doing the integrals).

When the coefficients of DD are real numbers, then the coefficients of UU are also real numbers. In fact, it is not hard to check that UD¯=U¯DU_{\bar{D}}=\bar{U}_{D}, where D¯\bar{D} means the polynomial whose coefficients are conjugates of the coefficients of DD.

Finally, inspired by theorem 0.2, we have the following general method to check whether a polynomial D(z)D(z) is Pellian or not.

Theorem 2.8.

Suppose D(z)[z]D(z)\in\mathbb{C}[z] is a monic polynomial of even degree with distinct roots. Then D(z)D(z) is Pellian if and only if all the following integrals

z1zjU(z)D(z)𝑑z\int_{z_{1}}^{z_{j}}\dfrac{U(z)}{\sqrt{D(z)}}dz

are integer multiplications of πi\pi i.

Proof.

If D(z)D(z) is Pellian, suppose

p(z)2q(z)2D(z)=1.p(z)^{2}-q(z)^{2}D(z)=1.

Then U(z)=p(z)q(z)U(z)=\dfrac{p^{\prime}(z)}{q(z)} and

z1zU(z)D(z)𝑑z=ln(p(z)+q(z)D(z)).\int_{z_{1}}^{z}\dfrac{U(z)}{\sqrt{D(z)}}dz=\ln(p(z)+q(z)\sqrt{D(z)}).

But from Pell equation, p(zj)=±1p(z_{j})=\pm 1 for any jj, so

z1zjU(z)D(z)𝑑z\int_{z_{1}}^{z_{j}}\dfrac{U(z)}{\sqrt{D(z)}}dz

is an integer multiplication of πi\pi i.

The other direction: if all

z1zjU(z)D(z)𝑑z\int_{z_{1}}^{z_{j}}\dfrac{U(z)}{\sqrt{D(z)}}dz

are integer multiplications of πi\pi i, then

exp(zizU(z)D(z)𝑑z)\exp(\int_{z_{i}}^{z}\dfrac{U(z)}{\sqrt{D(z)}}dz)

defines a holomorphic function on Σ={(w,z)2|w2=D(z)}\Sigma=\{(w,z)\in\mathbb{C}^{2}|w^{2}=D(z)\} for the following reason:

We know that the topology of Σ\Sigma is a surface of genus n1n-1 minus two points. The first homology of Σ\Sigma is generated by the following cycles:

For each zjz1z_{j}\neq z_{1}, one goes from z1z_{1} to zjz_{j} in one branch and go back from zjz_{j} to z1z_{1} using the same path but in another branch. This loop on Σ\Sigma is denoted as Cj(j1)C_{j}(j\neq 1).

We have assumed that the path integration of U(z)dzD(z)\dfrac{U(z)dz}{\sqrt{D(z)}} is a multiplication of 2πi2\pi i along any other loop CjC_{j}. So the exp of this integral is well-defined holomorphic function on Σ\Sigma without ambiguity.

Finally, by GAGA (see for example, [9]), a holomorphic function on Σ\Sigma with polynomial growth must be a polynomial in ww and zz. Since w2=D(z)w^{2}=D(z), we may write this polynomial as

p(z)+q(z)w=p(z)+q(z)D(z).p(z)+q(z)w=p(z)+q(z)\sqrt{D(z)}.

And the unique 2\mathbb{Z}_{2} harmonic function with bounded zero locus is

G(z)=ln|p(z)+q(z)D(z)|.G(z)=\ln|p(z)+q(z)\sqrt{D(z)}|.

Since G(z)G(z) has monodromy 1-1 around any point z=ziz=z_{i}, a change of the sign of D(z)\sqrt{D(z)} leads to a change of sign in G(z)G(z) as well. We have

ln|p(z)+q(z)D(z)|=ln|p(z)q(a)D(z)|.\ln|p(z)+q(z)\sqrt{D(z)}|=-\ln|p(z)-q(a)\sqrt{D(z)}|.

So |p(z)2q(z)2D(z)|=1.|p(z)^{2}-q(z)^{2}D(z)|=1. We may assume p(z)2q(z)2D(z)=ϵp(z)^{2}-q(z)^{2}D(z)=\epsilon, where ϵ\epsilon is a complex-valued constant with norm 11. Then (ϵ12p(z),ϵ12q(z))(\epsilon^{-\frac{1}{2}}p(z),\epsilon^{-\frac{1}{2}}q(z)) is a nontrivial solution to the Pell equation for D(z)D(z). ∎

Remark: Although we assumed that D(z)D(z) 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 U(z)U(z) has a common factor (zzi)(z-z_{i}) with D(z)D(z). Let k1k_{1} be the order of (zzi)(z-z_{i}) in D(z)D(z) and k2k_{2} be the order of (zzi)(z-z_{i}) in U(z)U(z). Then k2k11k_{2}\geq k_{1}-1. In addition, (zzi)2D(z)(z-z_{i})^{2}D(z) is Pellian if and only if k2k1k_{2}\geq k_{1}.

To finish this section, we write down a concrete example of a 2\mathbb{Z}_{2} 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 2\mathbb{Z}_{2} harmonic functions.

Example 2.10.

Let D(z)=z2n1D(z)=z^{2n}-1, then

G(z)=ln|zn+z2n1|.G(z)=\ln|z^{n}+\sqrt{z^{2n}-1}|.

3 The zero locus

This short section studies the zero locus of the 2\mathbb{Z}_{2} harmonic functions ff in 2\K\mathbb{R}^{2}\backslash K and its relationship to u(z)u(z) in the correspondence given by theorem 0.2.

For an ordinary harmonic function in 2\mathbb{R}^{2}, 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 2\mathbb{Z}_{2} harmonic functions on XX.

Theorem 3.1.

Suppose ff is a 2\mathbb{Z}_{2} harmonic function on XX. Then the zero locus of ff is the union of locally finite smooth curves whose boundaries are empty or belong to KK. 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 ff does not form a loop.

Proof.

Here is the sketch. Suppose PP is a zero point of a 2\mathbb{Z}_{2} harmonic function ff.

If PP does not belong to KK, then it is the same as the usual situation and a proof can be found in the textbook [10]. Otherwise, PP belongs to KK. Near PP,

f=Re(n=0+anz(2n+1)2)f=\text{Re}(\sum\limits_{n=0}^{+\infty}a_{n}z^{\frac{(2n+1)}{2}})

Suppose k0k\geq 0 is the smallest number such that ak0a_{k}\neq 0. We may assume aka_{k} is real, then

f=akρ(2k+1)2cos((2k+1)2θ)+Re(n=k+1+anz(2n+1)2).f=a_{k}\rho^{\frac{(2k+1)}{2}}\cos(\frac{(2k+1)}{2}\theta)+\text{Re}(\sum\limits_{n=k+1}^{+\infty}a_{n}z^{\frac{(2n+1)}{2}}).

So

f=akρ(2k+1)2cos((2k+1)2θ)+O(ρ(2k+3)2).f=a_{k}\rho^{\frac{(2k+1)}{2}}\cos(\frac{(2k+1)}{2}\theta)+O(\rho^{\frac{(2k+3)}{2}}).
df=akρ(2k1)2cos((2k+1)2θ)dρakρ(2k+1)2sin((2k1)2θ)dθ+O(ρ(2k+1)2).df=a_{k}\rho^{\frac{(2k-1)}{2}}\cos(\frac{(2k+1)}{2}\theta)d\rho-a_{k}\rho^{\frac{(2k+1)}{2}}\sin(\frac{(2k-1)}{2}\theta)d\theta+O(\rho^{\frac{(2k+1)}{2}}).

The above two expressions, together with the implicit function theorem, are enough to show that in a neighborhood of PP, the zero locus of ff are 2k+12k+1 smooth curves that intersect at their common boundary PP 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 ZZ of |f||f| has a closed loop as a subset. Then at least one connected component of 2\Z\mathbb{R}^{2}\backslash Z is bounded, denoted as UU. Note that setting f>0f>0 defines a trivialization of \mathcal{L} on UU. And ff is an ordinary harmonic function on UU without zero points in it. In addition, |f||f| can be extended as a continuous function to the closure of UU by setting |f|=0|f|=0 on the boundary. This implies that ff has a positive local maximum in UU, which is impossible. ∎

Definition 3.2.

Suppose z0z_{0} is a zero point of a 2\mathbb{Z}_{2} harmonic function ff. Suppose that there are a non-zero number aa and a half integer kk such that f=Re(a(zz0)k)+O(|zz0|k+1)f=\text{Re}(a(z-z_{0})^{k})+O(|z-z_{0}|^{k+1}) in a neighborhood of zz, then kk is called the degree of ff at z0z_{0}. From the proof of the previous theorem, we know that the zero locus of ff near kk is an intersection of 2k2k curves that share a common boundary point at z0z_{0} and intersect at z0z_{0} equiangularly.

Theorem 3.3.

Suppose z0z_{0} is a zero point of a 2\mathbb{Z}_{2} harmonic function ff. Suppose u(z)u(z) is the entire function defined in section 1. If the degree of ff at z0z_{0} is larger than 11, then z0z_{0} is a zero point of u(z)u(z) with degree k12\lfloor k-\dfrac{1}{2}\rfloor (the largest integer that is no greater than k12k-\dfrac{1}{2}).

Proof.

If z0z_{0} is a zero point of ff whose degree k>1k>1, then z0z_{0} is also a zero point of df+idfdf+i*df whose degree is k1k-1. Since u(z)=D(z)(df+idf)u(z)=\sqrt{D(z)}(df+i*df), z0z_{0} is a zero point of u(z)u(z) with degree k12\lfloor k-\dfrac{1}{2}\rfloor. ∎

Here are some easy examples.

Example 3.4.

Suppose D(z)D(z) is Pellian, and suppose

p(z)2q(z)2D(z)=1.p(z)^{2}-q(z)^{2}D(z)=1.

Then the zero locus of ff is given by the set

{z||p(z)+q(z)D(z)|=1}.\{z|~{}|p(z)+q(z)\sqrt{D(z)}|=1\}.

This implies at a point in the zero locus,

Im(p(z))=Re(q(z)D(z))=0and(Re(p(z)))2+(Im(q(z)D(z)))2=1.\text{Im}(p(z))=\text{Re}(q(z)\sqrt{D(z)})=0~{}~{}~{}\text{and}~{}~{}(\text{Re}(p(z)))^{2}+(\text{Im}(q(z)\sqrt{D(z)}))^{2}=1.

In particular,

  • If D(z)D(z) has only two roots and if they are distinct, then

    D(z)=(zz1)(zz2)=(zk)2m2.D(z)=(z-z_{1})(z-z_{2})=(z-k)^{2}-m^{2}.

    Then we can choose p(z)=zkmp(z)=\dfrac{z-k}{m}, q(z)=1mq(z)=\dfrac{1}{m} and the zero locus is the straight line segment connecting z1z_{1} and z2z_{2}. In this situation, U(z)=1U(z)=1.

  • If D(z)=p(z)21D(z)=p(z)^{2}-1, and q(z)=1q(z)=1, then the zero locus is

    {z|p(z)is real and|p(z)|1}=p1([1,1]).\{z|~{}~{}p(z)~{}\text{is real and}~{}|p(z)|\leq 1\}=p^{-1}([-1,1]).

    An typical example of this kind is: D(z)=z2n1D(z)=z^{2n}-1.

Example 3.5.

If D(z)D(z) is a polynomial of degree 44 with distinct roots and suppose D(z)=(zz1)(zz2)(zz3)(zz4)D(z)=(z-z_{1})(z-z_{2})(z-z_{3})(z-z_{4}), then the degree of U(z)U(z) is one. We may assume U(z)=zz0U(z)=z-z_{0}. (Note that D(z)D(z) 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 z1z_{1} and their other ends are z2z_{2}, z3z_{3}, z4z_{4} respectively. They intersect equiangularly at z1z_{1} and do not intersect at any other point. In this situation, z0=z1z_{0}=z_{1}. An typical example of this kind is: D(z)=z4zD(z)=z^{4}-z.

  • The zero locus is a combination of two curve segments and we may assume that one of them connects z1z_{1} and z2z_{2}, the other one connects z3z_{3} and z4z_{4}. If they intersect at a point in an orthogonal way, then this point must be z0z_{0}. Otherwise, they do not intersect.

The author is very curious how the four points z1,z2,z3z_{1},z_{2},z_{3} and z4z_{4} determine z0z_{0} 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.