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

\articleinfo\rcvdate\rvsdate

Isolated singularities of Toda equations and cyclic Higgs bundles

Qiongling Li  and  Takuro Mochizuki Qiongling Li:
Chern Institute of Mathematics and LPMC, Nankai University,
Tianjin 300071, China
[email protected] Takuro Mochizuki:
Research Institute for Mathematical Sciences, Kyoto University,
Kyoto 606-8512, Japan
[email protected]
Abstract.

This paper is the second part of our study on the Toda equations and the cyclic Higgs bundles associated with rr-differentials over non-compact Riemann surfaces. We classify all the solutions up to boundedness around the isolated singularity of an rr-differential under the assumption that the rr-differential is meromorphic or has some type of essential singularity. As a result, for example, we classify all the solutions on {\mathbb{C}} if the rr-differential is a finite sum of the exponential of polynomials.

Key words and phrases:
Cyclic Higgs bundles, Toda equations, harmonic metrics, essential singularity
2010 Mathematics Subject Classification:
53C07, 58E15

1. Introduction

1.1. Harmonic bundles and Toda equations

1.1.1. Higgs bundles associated with rr-differentials and harmonic metrics

Let XX be any Riemann surface. We fix a line bundle KX1/2K_{X}^{1/2} with an isomorphism KX1/2KX1/2KXK_{X}^{1/2}\otimes K_{X}^{1/2}\simeq K_{X}. Let rr be a positive integer. We set 𝕂X,r:=i=1rKX(r+12i)/2\mathbb{K}_{X,r}:=\bigoplus_{i=1}^{r}K_{X}^{(r+1-2i)/2}. We define the actions of Gr={a|ar=1}G_{r}=\{a\in{\mathbb{C}}\,|\,a^{r}=1\} on KX(r+12i)/2K_{X}^{(r+1-2i)/2} by av=aiva\bullet v=a^{i}v. They induce a GrG_{r}-action on 𝕂X,r\mathbb{K}_{X,r}. For any rr-differential qH0(X,KXr)q\in H^{0}(X,K_{X}^{r}), let θ(q)\theta(q) denote the Higgs field of 𝕂X,r\mathbb{K}_{X,r} induced by the identity map θ(q)i:KX(r+12i)/2KX(r+12(i+1))/2KX\theta(q)_{i}:K_{X}^{(r+1-2i)/2}\rightarrow K_{X}^{(r+1-2(i+1))/2}\otimes K_{X} (i=1,,r1)(i=1,\cdots,r-1) and θ(q)r=q:KX(1r)/2KX(r1)/2KX\theta(q)_{r}=q:K_{X}^{(1-r)/2}\rightarrow K_{X}^{(r-1)/2}\otimes K_{X}. Note that θ(q)\theta(q) is homogeneous with respect to the GrG_{r}-action. It implies that if hh is a harmonic metric of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) then a(h)a^{\ast}(h) is again a harmonic metric of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)).

Let Harm(q)\mathop{\rm Harm}\nolimits(q) denote the set of GrG_{r}-invariant harmonic metrics hh of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) such that det(h)=1\det(h)=1. By the GrG_{r}-invariance, the decomposition 𝕂X,r=i=1rKX(r+12i)/2\mathbb{K}_{X,r}=\bigoplus_{i=1}^{r}K_{X}^{(r+1-2i)/2} is orthogonal with respect to any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), and hence we obtain the decomposition h=h|KX(r+12i)/2h=\bigoplus h_{|K_{X}^{(r+1-2i)/2}}. Note that KX(r+12i)/2K_{X}^{(r+1-2i)/2} and KX(r+12(r+1i))/2=KX(r1+2i)/2K_{X}^{(r+1-2(r+1-i))/2}=K_{X}^{(-r-1+2i)/2} are mutually dual. We say that hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) is real if h|KX(r+12i)/2h_{|K_{X}^{(r+1-2i)/2}} and h|KX(r1+2i)/2h_{|K_{X}^{(-r-1+2i)/2}} are mutually dual. Let Harm(q)\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q) denote the subset of hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) which are real.

Recall that if XX is compact then the classification of harmonic metrics of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) is well known. Indeed, if XX is hyperbolic, the Higgs bundle (𝕂X,r(q),θ(q))(\mathbb{K}_{X,r}(q),\theta(q)) is stable for any qq. Hence, according to the Kobayashi-Hitchin correspondence for Higgs bundles due to Hitchin and Simpson ([19], [38]), (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) has a unique harmonic metric hh such that det(h)=1\det(h)=1. Moreover, as observed by Baraglia [1], the uniqueness implies that hh is GrG_{r}-invariant and real. In other words, for a compact hyperbolic Riemann surface XX, Harm(q)=Harm(q)\mathop{\rm Harm}\nolimits(q)=\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q) consists of a unique element. If XX is an elliptic curve, it is easy to see that Harm(q)\mathop{\rm Harm}\nolimits(q) consists of a unique element if q0q\neq 0, and that Harm(0)\mathop{\rm Harm}\nolimits(0) is empty. If XX is 1\mathbb{P}^{1}, it is easy to see that there is no non-zero holomorphic rr-differential, and that Harm(0)\mathop{\rm Harm}\nolimits(0) is empty. Note that if XX is compact hyperbolic and if q=0q=0, as observed by Hitchin and Simpson, (𝕂X,r,θ(0))(\mathbb{K}_{X,r},\theta(0)) with the harmonic metric is naturally a polarized variation of Hodge structure, which particularly implies the GrG_{r}-invariance.

If XX is non-compact, the uniqueness of harmonic metrics of the Higgs bundle (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) does not necessarily hold, and a harmonic metric of unit determinant is not necessarily GrG_{r}-invariant. But, if XX is the complement of a finite subset in a compact Riemann surface X¯\overline{X}, and if qq is meromorphic on X¯\overline{X}, we may still apply the Kobayashi-Hitchin correspondence for singular Higgs bundles due to Simpson [39] in the tame case and Biquard-Boalch [4] in the wild case. (See also [29] for the extension of a wild harmonic bundle to a filtered Higgs bundle.) Indeed, in [14, 15, 30, 31], Harm(q)\mathop{\rm Harm}\nolimits(q) was classified in the case X=X={\mathbb{C}}^{\ast} with q=zm(dz)rq=z^{m}(dz)^{r} motivated by the relation with the tttt^{\ast}-geometry [7], and the method in [30, 31] owes on the Kobayashi-Hitchin correspondence. (See [14, 15] for a different approach to the same issue.)

In this paper, as a continuation of [24], we investigate a classification of GrG_{r}-invariant harmonic metrics on (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) over a more general non-compact Riemann surface XX with a more general rr-differential qq, which is not covered by the theory of wild harmonic bundles. In particular, we shall closely study the case where qq has some type of essential singularity.

1.1.2. Toda equations associated with rr-differentials

Let gg be any Kähler metric of XX. It induces a GrG_{r}-invariant Hermitian metric h(1)(g)h^{(1)}(g) of 𝕂X,r\mathbb{K}_{X,r}. For any other GrG_{r}-invariant Hermitian metric hh such that det(h)=1\det(h)=1, we obtain a tuple of {\mathbb{R}}-valued functions 𝒘=(w1,,wr){\boldsymbol{w}}=(w_{1},\ldots,w_{r}) such that wi=0\sum w_{i}=0 by the relation

h|KX(r+12i)/2=ewih(1)(g)|KX(r+12i)/2=ewig(r+12i)/2.h_{|K_{X}^{(r+1-2i)/2}}=e^{w_{i}}h^{(1)}(g)_{|K_{X}^{(r+1-2i)/2}}=e^{w_{i}}g^{-(r+1-2i)/2}.

Then, hh is contained in Harm(q)\mathop{\rm Harm}\nolimits(q) if and only if the following type of Toda equation is satisfied:

{1Λg¯w1=ewr+w1|q|g2ew1+w2r12kg1Λg¯wi=ewi1+wiewi+wi+1r+12i2kg(i=2,,r1)1Λg¯wr=ewr1+wrewr+w1|q|g21r2kg\left\{\begin{array}[]{l}\sqrt{-1}\Lambda_{g}\partial\overline{\partial}w_{1}=e^{-w_{r}+w_{1}}|q|_{g}^{2}-e^{-w_{1}+w_{2}}-\frac{r-1}{2}k_{g}\\ \sqrt{-1}\Lambda_{g}\partial\overline{\partial}w_{i}=e^{-w_{i-1}+w_{i}}-e^{-w_{i}+w_{i+1}}-\frac{r+1-2i}{2}k_{g}\quad(i=2,\ldots,r-1)\\ \sqrt{-1}\Lambda_{g}\partial\overline{\partial}w_{r}=e^{-w_{r-1}+w_{r}}-e^{-w_{r}+w_{1}}|q|_{g}^{2}-\frac{1-r}{2}k_{g}\end{array}\right. (1)

Here, Λg\Lambda_{g} denotes the adjoint of the multiplication of the associated Kähler form, and kg=1ΛgR(g)k_{g}=\sqrt{-1}\Lambda_{g}R(g) is the Gaussian curvature of gg. Let Toda1(q,g)\mathop{\rm Toda}\nolimits_{1}(q,g) denote the set of solutions 𝒘{\boldsymbol{w}} of (1) satisfying wi=0\sum w_{i}=0. A solution 𝒘Toda1(q,g){\boldsymbol{w}}\in\mathop{\rm Toda}\nolimits_{1}(q,g) is called real if wi+wr+1i=0w_{i}+w_{r+1-i}=0. Let Toda1(q,g)\mathop{\rm Toda}\nolimits_{1}^{{\mathbb{R}}}(q,g) denote the set of real solutions of (1). As explained, there is a natural bijection Harm(q)Toda1(q,g)\mathop{\rm Harm}\nolimits(q)\simeq\mathop{\rm Toda}\nolimits_{1}(q,g), which induces Harm(q)Toda1(q,g)\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q)\simeq\mathop{\rm Toda}\nolimits_{1}^{{\mathbb{R}}}(q,g).

Remark 1.1.

The equation (1) is equivalent to the Toda equation in [24]. See §1.1.5 below.

Remark 1.2.

If we change the sign in the system (1), we obtain the classical Toda equation studied extensively in integrable system, e.g. see [5, 6]. Geometrically, the classical Toda equation gives rise to harmonic maps from surface to compact flag manifolds.

Remark 1.3.

A solution of the Toda equation gives rise to an equivariant harmonic map ff from a universal covering of XX to the symmetric space SL(r,)/SU(r)\mathop{\rm SL}\nolimits(r,{\mathbb{C}})/{\mathop{\rm SU}}(r) such that Tr(fi)=0\mathop{\rm Tr}\nolimits(\partial f^{\otimes i})=0 (i=1,,r1)(i=1,\cdots,r-1) except for Tr(fr)\mathop{\rm Tr}\nolimits(\partial f^{\otimes r}) is a nonzero constant multiple of qq. In lower rank, the Toda equation is encoded with much richer geometry. If r=2r=2, the Toda equation coincides with the Bochner equation for harmonic maps between surfaces for a given Hopf differential, e.g., see [37, 44, 45, 49]. For the study of solutions for given meromorphic quadratic differentials, one can check [17, 18, 50]. If r=3r=3, the Toda equation for a real solution coincides with Wang’s equation for hyperbolic affine spheres in 3\mathbb{R}^{3} for a given Pick differential, e.g., see [3, 20, 25, 46]. For the study of solutions for given meromorphic cubic differentials, one can check [2, 10, 26, 27, 35]. If r=4r=4, the Toda equation for a real solution coincides with the Gauss-Ricci equation for maximal surfaces in 2,2\mathbb{H}^{2,2}, e.g., see [8, 42].

1.1.3. Existence and uniqueness of complete solutions

A solution 𝒘Toda1(q,g){\boldsymbol{w}}\in\mathop{\rm Toda}\nolimits_{1}(q,g) is called complete if the metrics ewi+1wige^{w_{i+1}-w_{i}}g (i=1,,r1)(i=1,\ldots,r-1) are complete. In terms of harmonic metrics, it is equivalent to the condition that the Kähler metrics g(h)ig(h)_{i} (i=1,,r1)(i=1,\ldots,r-1) induced by h|KX(r+12(i+1))/2h|KX(r+12i)/21h_{|K_{X}^{(r+1-2(i+1))/2}}\otimes h_{|K_{X}^{(r+1-2i)/2}}^{-1} are complete on XX, where we naturally identify the tangent bundle KX1K_{X}^{-1} of XX with KX(r+12(i+1))/2(KX(r+12i)/2)1K_{X}^{(r+1-2(i+1))/2}\otimes(K_{X}^{(r+1-2i)/2})^{-1}. The following fundamental theorem is proved in [24]. (See [24] for more detailed properties of complete solutions.)

Theorem 1.4 ([24]).

Let XX be a non-compact Riemann surface with a holomorphic rr-differential qq. Assume either (i) XX is hyperbolic, or (ii) XX is parabolic and q0q\neq 0. Here, we say that XX is hyperbolic (resp. parabolic) if a universal covering of XX is the upper half plane (resp. {\mathbb{C}}). Then, there uniquely exists a complete solution 𝐰cToda1(q,g){\boldsymbol{w}}^{\mathop{\rm c}\nolimits}\in\mathop{\rm Toda}\nolimits_{1}(q,g). It is real, i.e., 𝐰cToda1(q,g){\boldsymbol{w}}^{\mathop{\rm c}\nolimits}\in\mathop{\rm Toda}\nolimits_{1}^{{\mathbb{R}}}(q,g). Moreover, there exists a complete Kähler metric g~\widetilde{g} of XX such that ewi+1wige^{w_{i+1}-w_{i}}g (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded with g~\widetilde{g}, and that |q|g~|q|_{\widetilde{g}} is bounded. ∎

When we emphasize the dependence on (q,g)(q,g), we use the notation 𝒘c(q,g){\boldsymbol{w}}^{\mathop{\rm c}\nolimits}(q,g). Let hch^{\mathop{\rm c}\nolimits} denote the corresponding harmonic metric, which is independent of the choice of gg. We shall also use the notation hc(q)h^{\mathop{\rm c}\nolimits}(q) when we emphasize the dependence on qq.

1.1.4. Uniqueness and non-uniqueness of general solutions

Suppose that qq has finitely many zeros. In this case, the results in [24] clarify whether general solutions of (1) are uniquely determined or not. Let NN be any relatively compact open neighbourhood of the zero set of qq. On XNX\setminus N, we obtain the Kähler metric |q|2/r:=(qq¯)1/r|q|^{2/r}:=(q\cdot\overline{q})^{1/r}. We proved the following proposition in [24] on the uniqueness. Note that such a uniqueness was first proved in [23] for X=X={\mathbb{C}} with a polynomial rr-differential (r=2,3)(r=2,3).

Proposition 1.5 ([24]).

If |q|2/r|q|^{2/r} induces a complete distance on XNX\setminus N, then the equation (1) has a unique solution, i.e., Toda1(q,g)={𝐰c(q,g)}\mathop{\rm Toda}\nolimits_{1}(q,g)=\{{\boldsymbol{w}}^{\mathop{\rm c}\nolimits}(q,g)\} and Harm(q)={hc(q)}\mathop{\rm Harm}\nolimits(q)=\{h^{\mathop{\rm c}\nolimits}(q)\}. ∎

Remark 1.6.

In fact, the condition |q|2/r|q|^{2/r} induces a complete distance on XNX\setminus N is equivalent to that X=X¯DX=\bar{X}\setminus D where X¯\bar{X} is compact and DD is finite, qq is meromorphic on X¯\bar{X} with poles at each point of DD of pole order at least rr. Here is a brief argument. Since qq has finitely many zeros, one can easily extend |q|2/r|q|^{2/r} to a smooth metric on XX, which is obviously of finite total curvature. By using Huber’s theorem on complete surfaces with finite total curvature, we obtain X=X¯DX=\bar{X}\setminus D where X¯\bar{X} is compact and DD is finite. The rest is proven using a similar argument as in [35, Lemma B.3].

As for the non-uniqueness, the following proposition is proved in [24].

Proposition 1.7 ([24]).

If qq has finitely many zeros, there exists a solution 𝐰Toda1(q,g){\boldsymbol{w}}\in\mathop{\rm Toda}\nolimits_{1}^{{\mathbb{R}}}(q,g) such that outside a compact subset KK,

wir+12irlog|q|g,1ir.w_{i}\sim-\frac{r+1-2i}{r}\log|q|_{g},\quad 1\leq i\leq r.

In particular, solutions of (1) are not unique if |q|2/r|q|^{2/r} does not induce a complete distance of XNX\setminus N. ∎

When the uniqueness does not hold, it is natural to study a classification of general solutions of (1) under a mild assumption for qq. For example, let X=X={\mathbb{C}}^{\ast} and q=zm(dz/z)rq=z^{m}(dz/z)^{r} (m>0)(m>0). According to [14, 15, 30, 31], for any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), we obtain a tuple of real numbers 𝒃(h)=(b(h)1,,b(h)r){\boldsymbol{b}}(h)=(b(h)_{1},\ldots,b(h)_{r}) determined by

b(h)i:=inf{c||z|c+i|(dz)(r+12i)/2|h is bounded around 0}.b(h)_{i}:=\inf\Bigl{\{}c\in{\mathbb{R}}\,\Big{|}\,|z|^{c+i}|(dz)^{(r+1-2i)/2}|_{h}\,\,\mbox{ is bounded around $0$}\Bigr{\}}.

The correspondence h𝒃(h)h\longmapsto{\boldsymbol{b}}(h) induces a bijection

Harm(q){(b1,,br)|b1b2brb1m,bi=r(r+1)/2}.\mathop{\rm Harm}\nolimits(q)\simeq\\ \left\{(b_{1},\ldots,b_{r})\,\left|\,b_{1}\geq b_{2}\geq\cdots\geq b_{r}\geq b_{1}-m,\,\,\sum b_{i}=-r(r+1)/2\right.\right\}. (2)

It is our purpose in this paper to pursue a similar classification of Harm(q)\mathop{\rm Harm}\nolimits(q) in a more general situation, which we shall explain in the following subsections.

1.1.5. Remark on the difference of the conventions for the induced Hermitian metrics

We use a different convention from [24] for the induced Hermitian metric on KXj/2K_{X}^{j/2}. For a Kähler metric g=g0dzdz¯g=g_{0}\,dz\otimes d\overline{z}, we set |(dz)j/2|g2=(g02)j/2|(dz)^{j/2}|^{2}_{g}=(\frac{g_{0}}{2})^{-j/2} in this paper. (For example, see [13].) In [24], we used the norm (|(dz)j/2|g)2=g0j/2(|(dz)^{j/2}|^{\prime}_{g})^{2}=g_{0}^{-j/2}. In particular, |q|g2=2r(|q|g)2|q|_{g}^{2}=2^{r}(|q|^{\prime}_{g})^{2}. Let h(0)(g)h^{(0)}(g) denote the induced Hermitian metric of 𝕂X,r\mathbb{K}_{X,r} obtained by the latter convention. Then, for an r{\mathbb{R}}^{r}-valued function 𝒖{\boldsymbol{u}}, a Hermitian metric i=1reuih(0)(g)|KX(r+12i)/2\bigoplus_{i=1}^{r}e^{u_{i}}h^{(0)}(g)_{|K_{X}^{(r+1-2i)/2}} is a harmonic metric of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) if and only if the following equation is satisfied, which is studied in [24]:

{121Λg¯u1=eur+u1|q|g 2eu1+u2r14kg121Λg¯ui=eui1+uieui+ui+1r+12i4kg(i=2,,r1)121Λg¯ur=eur1+ureur+u1|q|g 21r4kg\left\{\begin{array}[]{l}\frac{1}{2}\sqrt{-1}\Lambda_{g}\partial\overline{\partial}u_{1}=e^{-u_{r}+u_{1}}|q|_{g}^{\prime\,2}-e^{-u_{1}+u_{2}}-\frac{r-1}{4}k_{g}\\ \frac{1}{2}\sqrt{-1}\Lambda_{g}\partial\overline{\partial}u_{i}=e^{-u_{i-1}+u_{i}}-e^{-u_{i}+u_{i+1}}-\frac{r+1-2i}{4}k_{g}\quad(i=2,\ldots,r-1)\\ \frac{1}{2}\sqrt{-1}\Lambda_{g}\partial\overline{\partial}u_{r}=e^{-u_{r-1}+u_{r}}-e^{-u_{r}+u_{1}}|q|_{g}^{\prime\,2}-\frac{1-r}{4}k_{g}\end{array}\right. (3)

An r{\mathbb{R}}^{r}-valued function 𝒘{\boldsymbol{w}} satisfies (1) if and only if the tuple ui=wi+r+12i2log2u_{i}=w_{i}+\frac{r+1-2i}{2}\log 2 (i=1,,r)(i=1,\ldots,r) satisfies (3).

1.2. Isolated singularities

Let DXD\subset X be a finite subset. Let us consider the case where qq is a holomorphic rr-differential on XDX\setminus D, which is not constantly 0. We shall study the classification of the behaviour of hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) up to boundedness around each point PP of DD. Let (XP,zP)(X_{P},z_{P}) denote a holomorphic coordinate neighbourhood around PP with zP(P)=0z_{P}(P)=0. We set XP:=XP{P}X_{P}^{\ast}:=X_{P}\setminus\{P\}.

1.2.1. The case of poles

Let us recall that if PP is a pole of qq, a general theory of harmonic bundles [38, 29, 32] allows us to obtain the classification in terms of parabolic structure up to boundedness. We have the expression

q|XP=zPmPαP(dzP/zP)r,q_{|X_{P}^{\ast}}=z_{P}^{m_{P}}\alpha_{P}\cdot(dz_{P}/z_{P})^{r},

where mPm_{P} denotes an integer, and αP\alpha_{P} induces a holomorphic function on XPX_{P} such that αP(P)0\alpha_{P}(P)\neq 0. If XPX_{P} is sufficiently small, αP\alpha_{P} is nowhere vanishing.

If mP0m_{P}\leq 0, we obtain the following estimate:

log|(dzP)(r+12i)/2|h+r+12i2rlog|zP|mPr=O(1).\log\bigl{|}(dz_{P})^{(r+1-2i)/2}\bigr{|}_{h}+\frac{r+1-2i}{2r}\log|z_{P}|^{m_{P}-r}=O(1).

In particular, any h1h_{1} and h2Harm(q)h_{2}\in\mathop{\rm Harm}\nolimits(q) are mutually bounded around PP. (See §3.5.1 for more details.)

If mP>0m_{P}>0, there exists 𝒃P(h)=(bP,1(h),,bP,r(h))r{\boldsymbol{b}}_{P}(h)=(b_{P,1}(h),\ldots,b_{P,r}(h))\in{\mathbb{R}}^{r} satisfying

{bP,1(h)bP,2(h)bP,r(h)bP,1(h)mP,i=1rbP,i(h)=r(r+1)2,\left\{\begin{array}[]{l}b_{P,1}(h)\geq b_{P,2}(h)\geq\cdots\geq b_{P,r}(h)\geq b_{P,1}(h)-m_{P},\\ \sum_{i=1}^{r}b_{P,i}(h)=-\frac{r(r+1)}{2},\end{array}\right. (4)

for which the following estimates hold:

log|(dzP)(r+12i)/2|h+(bP,i(h)+i)log|zP|ki(𝒃P(h))2log(log|zP|)=O(1).\log\bigl{|}(dz_{P})^{(r+1-2i)/2}\bigr{|}_{h}+(b_{P,i}(h)+i)\log|z_{P}|-\frac{k_{i}({\boldsymbol{b}}_{P}(h))}{2}\log\bigl{(}-\log|z_{P}|\bigr{)}\\ =O(1). (5)

Here, 𝒌(𝒃P(h))=(k1(𝒃P(h)),,kr(𝒃P(h))){\boldsymbol{k}}({\boldsymbol{b}}_{P}(h))=(k_{1}({\boldsymbol{b}}_{P}(h)),\ldots,k_{r}({\boldsymbol{b}}_{P}(h))) denotes the tuple of integers determined by 𝒃P(h){\boldsymbol{b}}_{P}(h) as in §3.5.2. It particularly implies that hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) are mutually bounded around PP if and only if 𝒃P(h1)=𝒃P(h2){\boldsymbol{b}}_{P}(h_{1})={\boldsymbol{b}}_{P}(h_{2}). (See §3.5.2 for more details.)

1.2.2. Holomorphic functions with multiple growth orders

To study more general cases, we introduce a reasonable subclass of essential singularities of holomorphic functions. Let ff denote a holomorphic function on XPX_{P}^{\ast}.

Let ϖ:X~DX\varpi:\widetilde{X}_{D}\longrightarrow X denote the oriented real blowing up. Let Qϖ1(P)Q\in\varpi^{-1}(P). We fix a branch of logzP\log z_{P} around QQ.

  • We say that ff is regularly bounded around QQ if there exist a neighbourhood 𝒰\mathcal{U} of QQ in X~D\widetilde{X}_{D}, non-zero complex numbers αj\alpha_{j} (j=1,,m)(j=1,\ldots,m), and mutually distinct real numbers cjc_{j} (j=1,,m)(j=1,\ldots,m) for which |fj=1mαjzP1cj|0\bigl{|}f-\sum_{j=1}^{m}\alpha_{j}z_{P}^{\sqrt{-1}c_{j}}\bigr{|}\to 0 holds as |zP|0|z_{P}|\to 0 in 𝒰ϖ1(P)\mathcal{U}\setminus\varpi^{-1}(P).

  • We say that ff has a single growth order at QQ if there exists 𝔞(f,Q)b>0zPb\mathfrak{a}(f,Q)\in\bigoplus_{b>0}{\mathbb{C}}z_{P}^{-b}, a(f,Q)a(f,Q)\in{\mathbb{R}} and j(f,Q)0j(f,Q)\in{\mathbb{Z}}_{\geq 0} such that e𝔞(f,Q)zPa(f,Q)(logzP)j(f,Q)fe^{-\mathfrak{a}(f,Q)}z_{P}^{-a(f,Q)}(\log z_{P})^{-j(f,Q)}f is regularly bounded around QQ.

  • We say that ff has multiple growth orders at QQ if ff is expressed as a finite sum of holomorphic functions fif_{i} (i=1,,m)(i=1,\ldots,m) with a single growth order at QQ such that 𝔞(fi,Q)\mathfrak{a}(f_{i},Q) (i=1,,m)(i=1,\ldots,m) are mutually distinct.

  • We say that ff has multiple growth orders at PP if ff has multiple growth orders at any Qϖ1(P)Q\in\varpi^{-1}(P).

Remark 1.8.

If ff satisfies a linear differential equation

zPnf+j=0n1γj(zP)zPjf=0\partial_{z_{P}}^{n}f+\sum_{j=0}^{n-1}\gamma_{j}(z_{P})\partial_{z_{P}}^{j}f=0

for meromorphic functions γj\gamma_{j} on (XP,P)(X_{P},P), then ff has multiple growth orders at PP. (For example, see [28, §II.1].)

Note that when ff has single growth order at Qϖ1(P)Q\in\varpi^{-1}(P),

|𝔞(f,Q)|1Re(𝔞(f,Q))|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q))

induces a continuous function on a neighbourhood of QQ. The function is also denoted by |𝔞(f,Q)|1Re(𝔞(f,Q))|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q)). As the value of the function at QQ, we obtain a real number |𝔞(f,Q)|1Re(𝔞(f,Q))(Q)|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q))(Q).

If ff has multiple growth orders at PP, there exists a finite subset 𝒵(f)ϖ1(P)\mathcal{Z}(f)\subset\varpi^{-1}(P) such that the following holds at Qϖ1(P)𝒵(f)Q\in\varpi^{-1}(P)\setminus\mathcal{Z}(f).

  • ff has a single growth order at QQ.

  • |𝔞(f,Q)|1Re(𝔞(f,Q))(Q)0|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q))(Q)\neq 0,

Note that ϖ1(P)\varpi^{-1}(P) (PD)(P\in D) are identified with S1S^{1}. An interval II of ϖ1(P)\varpi^{-1}(P) is called special with respect to ff if there exists αI\alpha_{I}\in{\mathbb{C}}^{\ast} and ρI>0\rho_{I}>0 such that the following holds.

  • The length of II is π/ρI\pi/\rho_{I}.

  • For any QI𝒵(f)Q\in I\setminus\mathcal{Z}(f), we obtain that

    𝔞(f,Q)αIzPρI0<b<ρIzPb,\mathfrak{a}(f,Q)-\alpha_{I}z_{P}^{-\rho_{I}}\in\bigoplus_{0<b<\rho_{I}}{\mathbb{C}}z_{P}^{-b},

    and that |αIzPρI|1Re(αIzPρI)(Q)<0|\alpha_{I}z_{P}^{-\rho_{I}}|^{-1}\mathop{\rm Re}\nolimits(\alpha_{I}z_{P}^{-\rho_{I}})(Q)<0. Here, we assume that the branches of logzP\log z_{P} at QQ are analytically continued along II.

1.2.3. The case where qq is not meromorphic but has multiple growth orders

For PDP\in D, we have the expression q=fP(dzP)rq=f_{P}(dz_{P})^{r}, where fPf_{P} is a holomorphic function on XPX_{P}^{\ast}. We say that qq has multiple growth orders at PP, if fPf_{P} has multiple growth orders at PP.

Assume that qq is not meromorphic but has multiple growth orders at PP. Let 𝒮(q,P)\mathcal{S}(q,P) denote the (possibly empty) set of intervals in ϖ1(P)\varpi^{-1}(P) which are special with respect to fPf_{P}. Let 𝒫\mathcal{P} denote the set of 𝒂r{\boldsymbol{a}}\in{\mathbb{R}}^{r} such that

a1a2ara11,ai=0.a_{1}\geq a_{2}\geq\cdots\geq a_{r}\geq a_{1}-1,\quad\sum a_{i}=0.
Theorem 1.9 (Theorem 6.1).

  • For any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), there exist 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} (I𝒮(q,P))(I\in\mathcal{S}(q,P)) and ϵ>0\epsilon>0 such that the following estimates hold as |zP|0|z_{P}|\to 0 on {|arg(αIzPρ(I))π|<(1δ)π/2}\bigl{\{}|\arg(\alpha_{I}z_{P}^{-\rho(I)})-\pi|<(1-\delta)\pi/2\bigr{\}} for any δ>0\delta>0:

    log|(dzP)(r+12i)/2|h+ai(h)Re(αIzPρI)=O(|zP|ρI+ϵ).\log\bigl{|}(dz_{P})^{(r+1-2i)/2}\bigr{|}_{h}+a_{i}(h)\mathop{\rm Re}\nolimits\bigl{(}\alpha_{I}z_{P}^{-\rho_{I}}\bigr{)}=O\bigl{(}|z_{P}|^{-\rho_{I}+\epsilon}\bigr{)}. (6)
  • If 𝒂I(h1)=𝒂I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) for any I𝒮(q,P)I\in\mathcal{S}(q,P), there exists a relatively compact neighbourhood XPX_{P} of PP such that h1h_{1} and h2h_{2} are mutually bounded on XPX_{P}^{\ast}.

  • For any 𝒂I𝒫{\boldsymbol{a}}_{I}\in\mathcal{P} (I𝒮(q,P))(I\in\mathcal{S}(q,P)), there exists hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that 𝒂I(h)=𝒂I{\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}_{I} (I𝒮(q,P))(I\in\mathcal{S}(q,P)).

Let us explain rough ideas for the proof. Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). If fPf_{P} has a single growth order at QQ with

|𝔞(fP,Q)|1Re(𝔞(fP,Q))(Q)>0,|\mathfrak{a}(f_{P},Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f_{P},Q))(Q)>0,

then hh should be close to a canonical harmonic metric hcanh_{\mathop{\rm can}\nolimits} around QQ outside of some neighbourhoods of the zeroes of qq, which follows from variants of Simpson’s main estimate. (See Proposition 5.1. See §3.2.1 for hcanh_{\mathop{\rm can}\nolimits}.) Therefore, around such QQ, we may apply the subharmonicity of the difference of two harmonic metrics (see §2.4) to obtain that two harmonic metrics are mutually bounded. Let II be an interval of ϖ1(P)\varpi^{-1}(P) such that

|𝔞(fP,Q)|1Re(𝔞(fP,Q))(Q)<0|\mathfrak{a}(f_{P},Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f_{P},Q))(Q)<0

for any QI𝒵(fP)Q\in I\setminus\mathcal{Z}(f_{P}). Assume that II is maximal among such intervals. There exist ρI>0\rho_{I}\in{\mathbb{R}}_{>0} and αI\alpha_{I}\in{\mathbb{C}}^{\ast} such that 𝔞(fP,Q)αIzPρI0<b<ρIzPb\mathfrak{a}(f_{P},Q)-\alpha_{I}z_{P}^{-\rho_{I}}\in\bigoplus_{0<b<\rho_{I}}{\mathbb{C}}z_{P}^{-b} for any QI𝒵(f)Q\in I\setminus\mathcal{Z}(f). Moreover, the length of II is not strictly greater than π/ρI\pi/\rho_{I}. If the length of II is strictly smaller than π/ρI\pi/\rho_{I}, we can apply Phragmén-Lindelöf theorem (Proposition 4.1) to obtain that any hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) are mutually bounded around the closure I¯\overline{I} of II (Theorem 5.2). If the length of II is π/ρ\pi/\rho, i.e., if II is special, then the Nevanlinna formula (Proposition 4.4) says that there exists the tuple 𝒂I(h){\boldsymbol{a}}_{I}(h) for any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that the estimate (6) holds. (See Theorem 5.3.) We can also apply Phragmén-Lindelöf theorem to obtain that any hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) are mutually bounded around I¯\overline{I} if 𝒂I(h1)=𝒂I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}).

For the construction of hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that 𝒂I(h)=𝒂I{\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}_{I} (I𝒮(q,P))(I\in\mathcal{S}(q,P)), we shall find a neighbourhood 𝒰I\mathcal{U}_{I} of I𝒮(q,P)I\in\mathcal{S}(q,P), a GrG_{r}-invariant Hermitian metric h𝒂h_{{\boldsymbol{a}}} of 𝕂XD,r\mathbb{K}_{X\setminus D,r} with det(h𝒂)=1\det(h_{{\boldsymbol{a}}})=1, and an increasing sequence YiY_{i} of relatively compact open subsets of XDX\setminus D such that the following holds.

  • The estimate (6) holds for h𝒂h_{{\boldsymbol{a}}} and each I𝒮(q,P)I\in\mathcal{S}(q,P).

  • The boundary Yi\partial Y_{i} are smooth, and Yi=XD\bigcup Y_{i}=X\setminus D.

  • For hiHarm(q|Yi)h_{i}\in\mathop{\rm Harm}\nolimits(q_{|Y_{i}}) such that hi|Yi=h𝒂|Yih_{i|\partial Y_{i}}=h_{{\boldsymbol{a}}|\partial Y_{i}}, we obtain logTr(hih𝒂|Yi1)C|zP|ρI+ϵ\log\mathop{\rm Tr}\nolimits(h_{i}\cdot h_{{\boldsymbol{a}}|Y_{i}}^{-1})\leq C|z_{P}|^{-\rho_{I}+\epsilon} on Yi(𝒰Iϖ1(P))Y_{i}\cap(\mathcal{U}_{I}\setminus\varpi^{-1}(P)), where C>0C>0 and 0<ϵ<ρI0<\epsilon<\rho_{I} are independent of ii.

Note that there always exists hiHarm(q|Yi)h_{i}\in\mathop{\rm Harm}\nolimits(q_{|Y_{i}}) such that hi|Yi=h𝒂|Yih_{i|\partial Y_{i}}=h_{{\boldsymbol{a}}|\partial Y_{i}} by a theorem of Donaldson [9] (see Proposition 2.1). For any compact subset KXDK\subset X\setminus D, there exists CK>0C_{K}>0 such that Tr(hih𝒂|Yi1)<CK\mathop{\rm Tr}\nolimits(h_{i}\cdot h_{{\boldsymbol{a}}|Y_{i}}^{-1})<C_{K} on KK for any ii such that KYiK\subset Y_{i}. By taking a subsequence, we may assume that the sequence hih_{i} is convergent to hHarm(q)h_{\infty}\in\mathop{\rm Harm}\nolimits(q). (See Proposition 3.15.) By the uniform estimate logTr(hih𝒂|Yi1)C|zP|ρI+ϵ\log\mathop{\rm Tr}\nolimits(h_{i}h_{{\boldsymbol{a}}|Y_{i}}^{-1})\leq C|z_{P}|^{-\rho_{I}+\epsilon} on Yi(𝒰Iϖ1(P))Y_{i}\cap(\mathcal{U}_{I}\setminus\varpi^{-1}(P)), we obtain the estimate logTr(hh𝒂1)C|zP|ρI+ϵ\log\mathop{\rm Tr}\nolimits(h_{\infty}h_{{\boldsymbol{a}}}^{-1})\leq C|z_{P}|^{-\rho_{I}+\epsilon} on 𝒰Iϖ1(P)\mathcal{U}_{I}\setminus\varpi^{-1}(P), and hence 𝒂I(h)=𝒂I{\boldsymbol{a}}_{I}(h_{\infty})={\boldsymbol{a}}_{I}.

1.3. Global classification in the case where qq has multiple growth orders

Suppose that qq has multiple growth orders at each point of DD. Let DmeroD_{\mathop{\rm mero}\nolimits} denote the set of PDP\in D such that qq is meromorphic at PP. We divide Dmero=D>0D0D_{\mathop{\rm mero}\nolimits}=D_{>0}\sqcup D_{\leq 0}, where D>0D_{>0} denotes the set of PDP\in D such that mP>0m_{P}>0, and D0:=DD>0D_{\leq 0}:=D\setminus D_{>0}. For PD>0P\in D_{>0}, let 𝒫(q,P)\mathcal{P}(q,P) denote the set of 𝒃r{\boldsymbol{b}}\in{\mathbb{R}}^{r} satisfying the condition (4). We set Dess:=DDmeroD_{\mathop{\rm ess}\nolimits}:=D\setminus D_{\mathop{\rm mero}\nolimits}. We set 𝒮(q):=PDess𝒮(q,P)\mathcal{S}(q):=\coprod_{P\in D_{\mathop{\rm ess}\nolimits}}\mathcal{S}(q,P). Then, we obtain the map

Harm(q)PD>0𝒫(q,P)×I𝒮(q)𝒫.\mathop{\rm Harm}\nolimits(q)\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}(q,P)\times\prod_{I\in\mathcal{S}(q)}\mathcal{P}. (7)

If D>0D_{>0} is empty, PD>0𝒫(q,P)\prod_{P\in D_{>0}}\mathcal{P}(q,P) denotes a set which consists of one element. Similarly, if 𝒮(q)\mathcal{S}(q) is empty, I𝒮(q)𝒫\prod_{I\in\mathcal{S}(q)}\mathcal{P} denotes a set which consists of one element.

Let 𝒫(q,P)\mathcal{P}^{{\mathbb{R}}}(q,P) denote the set of 𝒃𝒫(q,P){\boldsymbol{b}}\in\mathcal{P}(q,P) such that bi+br+1i=r1b_{i}+b_{r+1-i}=-r-1. Let 𝒫𝒫\mathcal{P}^{{\mathbb{R}}}\subset\mathcal{P} denote the set of 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P} such that ai+ar+1i=0a_{i}+a_{r+1-i}=0. The map (7) induces the following:

Harm(q)PD>0𝒫(q,P)×I𝒮(q)𝒫.\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q)\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}^{{\mathbb{R}}}(q,P)\times\prod_{I\in\mathcal{S}(q)}\mathcal{P}^{{\mathbb{R}}}. (8)

The following theorem is a special case of Theorem 1.12 below.

Theorem 1.10.

If XX is compact, then the maps (7) and (8) are bijective.

As a special case of Theorem 1.10, we obtain the following corollary in the meromorphic case.

Corollary 1.11.

Suppose that XX is compact, and that qq is meromorphic on (X,D)(X,D). Then, there exists a natural bijection Harm(q)PD>0𝒫(q,P)\mathop{\rm Harm}\nolimits(q)\simeq\prod_{P\in D_{>0}}\mathcal{P}(q,P). If moreover mP0m_{P}\leq 0 at each PDP\in D, we obtain Harm(q)={hc(q)}\mathop{\rm Harm}\nolimits(q)=\{h^{\mathop{\rm c}\nolimits}(q)\}. ∎

If XX is not compact, it is reasonable to impose an additional boundary condition on the behaviour of harmonic metrics around the infinity of XX. Recall that DD is assumed to be finite. Let Harm(q;D,c)\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) denote the set of hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that for any relatively compact open neighbourhood NN of DD, the Kähler metrics g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are complete. (See §1.1.3 for g(h)ig(h)_{i}.) It turns out that hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) is contained in Harm(q;D,c)\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) if and only if for any relatively compact open neighbourhood NN of DD, h|XNh_{|X\setminus N} and hc(q)|XNh^{\mathop{\rm c}\nolimits}(q)_{|X\setminus N} are mutually bounded. We set Harm(q;D,c):=Harm(q)Harm(q;D,c)\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q;D,\mathop{\rm c}\nolimits):=\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q)\cap\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits).

We obtain the following map as the restriction of (7):

Harm(q;D,c)PD>0𝒫(q,P)×I𝒮(q)𝒫.\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits)\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}(q,P)\times\prod_{I\in\mathcal{S}(q)}\mathcal{P}. (9)

Similarly, we obtain the following map as the restriction of (8):

Harm(q;D,c)PD>0𝒫(q,P)×I𝒮(q)𝒫.\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q;D,\mathop{\rm c}\nolimits)\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}^{{\mathbb{R}}}(q,P)\times\prod_{I\in\mathcal{S}(q)}\mathcal{P}^{{\mathbb{R}}}. (10)
Theorem 1.12 (Theorem 6.6).

The maps (9) and (10) are bijective.

Let us explain rough ideas for the proof of Theorem 1.12. For the uniqueness, the argument in [24] is available even in this situation. According to [38, Lemma 3.1] (see §2.4), for any two hjHarm(q)h_{j}\in\mathop{\rm Harm}\nolimits(q) (j=1,2)(j=1,2), we obtain the inequality (128) on XDX\setminus D. If 𝒃P(h1)=𝒃P(h2){\boldsymbol{b}}_{P}(h_{1})={\boldsymbol{b}}_{P}(h_{2}) (PD>0)(P\in D_{>0}) and 𝒂I(h1)=𝒂I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) (I𝒮(q))(I\in\mathcal{S}(q)), we obtain that h1h_{1} and h2h_{2} are mutually bounded, and hence [39, Lemma 2.2] implies that the inequality (128) weakly holds on XX. Then, as in [24], we may apply Omori-Yau maximum principle to obtain h1=h2h_{1}=h_{2}. Note that the argument can be simplified if XX is compact. Indeed, because XDX\setminus D is potential theoretically parabolic in this case, we obtain h1=h2h_{1}=h_{2} immediately from the mutual boundedness of h1h_{1} and h2h_{2}. (See Corollary 8.19 and Remark 8.20.)

As for the existence, for any (𝒃P)PD>0({\boldsymbol{b}}_{P})_{P\in D_{>0}} and (𝒂I)I𝒮(q)({\boldsymbol{a}}_{I})_{I\in\mathcal{S}(q)}, we can construct a GrG_{r}-invariant Hermitian metric h0h_{0} of 𝕂XD,r\mathbb{K}_{X\setminus D,r} such that the following holds.

  • There exists a relatively compact neighbourhood N1N_{1} of DD such that h0|XN1=hc(q)|XN1h_{0|X\setminus N_{1}}=h^{\mathop{\rm c}\nolimits}(q)_{|X\setminus N_{1}}.

  • There exists a relatively compact neighbourhood N2N1N_{2}\subset N_{1} such that h0|N2DHarm(q|N2D)h_{0|N_{2}\setminus D}\in\mathop{\rm Harm}\nolimits(q_{|N_{2}\setminus D}) and that

    𝒃P(h0|N2D)=𝒃P(PD>0),𝒂I(h0|N2D)=𝒂I(I𝒮(q)).{\boldsymbol{b}}_{P}(h_{0|N_{2}\setminus D})={\boldsymbol{b}}_{P}\quad(P\in D_{>0}),\quad\quad{\boldsymbol{a}}_{I}(h_{0|N_{2}\setminus D})={\boldsymbol{a}}_{I}\quad(I\in\mathcal{S}(q)).

Note that there exists a compact subset KXDK\subset X\setminus D such that h0|XKHarm(q|XK)h_{0|X\setminus K}\in\mathop{\rm Harm}\nolimits(q_{|X\setminus K}). We develop a variant of Kobayashi-Hitchin correspondence for harmonic bundles (Proposition 2.11, Proposition 3.18), which allows us to obtain hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that hh and h0h_{0} are mutually bounded. We remark that if XX is compact, the argument can be simplified, again. Indeed, we immediately obtain the desired metric by applying a theorem of Simpson [38] to h0h_{0}.

1.4. Comparison of the methods and the results

In [24] and this paper, we apply rather different two methods. The both methods have their own advantage. In fact, one of the goals of this work is to present readers as many as tools which can be useful in dealing with cyclic Higgs bundles over non-compact Riemann surfaces.

On one hand, the study in [24] is more p.d.e theoretic, and it is available in a quite general setting. The p.d.e tools like Omori-Yau and Cheng-Yau maximum principles work well for complete Riemann manifolds with bounded curvature. The method is particularly powerful in analyzing real solutions of (1), which allows us to obtain precise estimates of complete solutions. Therefore, we manage to use the maximum principles together with the method of super-subsolution to construct a complete solution for any Riemann surface and any holomorphic rr-differential.

On the other hand, the study in this paper heavily owes to the techniques and tools developed in the theory of Kobayashi-Hitchin correspondence for Higgs bundles, pioneered by Donaldson, Hitchin and Simpson. It is more efficient if the global information on qq is provided. In the case of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)), the Higgs field is generically regular semisimple, and the spectral curve is easily described as the set of rr-th roots of qq. It allows us to use efficiently variants of Simpson’s main estimate (see §3.3) in controlling the behaviour of harmonic metrics on (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) around the isolated singularities by using the information of qq. Therefore, for instance, under the assumptions that XX is compact, that DD is finite, and that qq has at most multiple growth orders at each point of DD, we can classify the set of solutions of the Toda equation associated with qq.

In the case that both methods can be applied, we obtain different features of solutions from these two methods. Suppose that XX is compact, and that qq has multiple growth orders at each point of DD. Using Theorem 1.10, we obtain the full set Harm(q)\mathop{\rm Harm}\nolimits(q) of solutions h,𝒜h^{\mathcal{B},\mathcal{A}}, where (,𝒜)=((𝒃P)PD>0,(𝒂I)I𝒮(q))PD>0𝒫(q,P)×I𝒮(q)𝒫(\mathcal{B},\mathcal{A})=\big{(}({\boldsymbol{b}}_{P})_{P\in D_{>0}},({\boldsymbol{a}}_{I})_{I\in\mathcal{S}(q)}\big{)}\in\prod_{P\in D_{>0}}\mathcal{P}(q,P)\times\prod_{I\in\mathcal{S}(q)}\mathcal{P} satisfies

ibP,i=r(r+1)2,bP,ibP,i+1(i=1,,r1),\displaystyle\sum_{i}b_{P,i}=-\frac{r(r+1)}{2},\quad b_{P,i}\geq b_{P,i+1}\quad(i=1,\cdots,r-1),
bP,rbP,1mP,for PD>0,\displaystyle b_{P,r}\geq b_{P,1}-m_{P},\quad\text{for $P\in D_{>0}$},
iaI,i=0,aI,1aI,2aI,raI,11,for I𝒮(q).\displaystyle\sum_{i}a_{I,i}=0,\quad a_{I,1}\geq a_{I,2}\geq\cdots\geq a_{I,r}\geq a_{I,1}-1,\quad\text{for $I\in\mathcal{S}(q)$}.

Suppose in addition that qq has finitely many zeros. ¿From Theorem 1.4 and Proposition 1.7, two special solutions are singled out. One is the complete solution in Theorem 1.4, which corresponds to

bP,i=r+12,for PD>0,\displaystyle b_{P,i}=-\frac{r+1}{2},\quad\text{for $P\in D_{>0}$},
aI,i=0,for I𝒮(q).\displaystyle a_{I,i}=0,\quad\text{for $I\in\mathcal{S}(q)$}.

The other one is the real solution purely controlled by qq which corresponds to

bP,i=r+12i2rmPr+12,for PD>0,\displaystyle b_{P,i}=\frac{r+1-2i}{2r}m_{P}-\frac{r+1}{2},\quad\text{for $P\in D_{>0}$},
aI,i=r+12i2r,for I𝒮(q).\displaystyle a_{I,i}=\frac{r+1-2i}{2r},\quad\text{for $I\in\mathcal{S}(q)$}.

1.5. A generalization

Assume that XDX\setminus D is non-compact, i.e., XX is non-compact, or XX is compact and DD is non-empty. (See Remark 7.12 for the case that XDX\setminus D is compact, i.e., XX is compact and DD is empty.) Let LiL_{i} (i=1,,r)(i=1,\ldots,r) be holomorphic line bundles on XDX\setminus D. Let ψi:LiLi+1KXD\psi_{i}:L_{i}\longrightarrow L_{i+1}\otimes K_{X\setminus D} (i=1,,r1)(i=1,\cdots,r-1) and ψr:LrL1KXD\psi_{r}:L_{r}\longrightarrow L_{1}\otimes K_{X\setminus D} be non-zero morphisms. We set E:=LiE:=\bigoplus L_{i}. Let θ\theta be the cyclic Higgs field of EE induced by ψi\psi_{i} (i=1,,r)(i=1,\ldots,r).

We obtain a holomorphic section q:=ψrψ1q:=\psi_{r}\circ\cdots\circ\psi_{1} of KXDrK_{X\setminus D}^{\otimes r}, and a holomorphic section qr1:=ψr1ψ1q_{\leq r-1}:=\psi_{r-1}\circ\cdots\circ\psi_{1} of Hom(L1,Lr)KXD(r1)\mathop{\rm Hom}\nolimits(L_{1},L_{r})\otimes K_{X\setminus D}^{\otimes(r-1)}. We assume the following.

  • The zero set of qr1q_{\leq r-1} is finite.

  • qq is not constantly 0. Moreover, qq has multiple growth orders at each PDP\in D.

Let DmeroD_{\mathop{\rm mero}\nolimits} denote the set of PDP\in D such that qq is meromorphic at PP. We put Dess:=DDmeroD_{\mathop{\rm ess}\nolimits}:=D\setminus D_{\mathop{\rm mero}\nolimits}. For PDmeroP\in D_{\mathop{\rm mero}\nolimits}, we describe q|XP=zPmPβP(dzP/zP)rq_{|X_{P}^{\ast}}=z_{P}^{m_{P}}\beta_{P}(dz_{P}/z_{P})^{r}, where βP\beta_{P} is a nowhere vanishing holomorphic function on XPX_{P}. Let D>0D_{>0} denote the set of PDmeroP\in D_{\mathop{\rm mero}\nolimits} such that mP>0m_{P}>0. We set D0:=DmeroD>0D_{\leq 0}:=D_{\mathop{\rm mero}\nolimits}\setminus D_{>0}.

Let hdet(E)h_{\det(E)} be a flat metric of det(E)\det(E). For each PDP\in D, there exist a tuple of frames 𝒗P=(vP,i|i=1,,r){\boldsymbol{v}}_{P}=(v_{P,i}\,|\,i=1,\ldots,r) of Li|XPL_{i|X_{P}^{\ast}} and a real number c(𝒗P)c({\boldsymbol{v}}_{P}) such that θ(vP,i)=vP,i+1dzP/zP\theta(v_{P,i})=v_{P,i+1}dz_{P}/z_{P} (i=1,,r1)(i=1,\ldots,r-1) and

|vP,1vP,r|hdet(E)=|zP|c(𝒗P).\bigl{|}v_{P,1}\wedge\cdots\wedge v_{P,r}\bigr{|}_{h_{\det(E)}}=|z_{P}|^{-c({\boldsymbol{v}}_{P})}.

Let 𝒫(q,P,𝒗P)\mathcal{P}(q,P,{\boldsymbol{v}}_{P}) be the set of 𝒃=(bi)r{\boldsymbol{b}}=(b_{i})\in{\mathbb{R}}^{r} such that

i=1rbi=c(𝒗P),bibi+1(i=1,,r1),brb1mP.\sum_{i=1}^{r}b_{i}=c({\boldsymbol{v}}_{P}),\quad b_{i}\geq b_{i+1}\,\,(i=1,\ldots,r-1),\quad b_{r}\geq b_{1}-m_{P}.
Remark 1.13.

If (E,θ)=(𝕂XD,r,θ(q))(E,\theta)=(\mathbb{K}_{X\setminus D,r},\theta(q)), we canonically choose

vP,i=zPi(dzP)(r+12i)/2.v_{P,i}=z_{P}^{i}(dz_{P})^{(r+1-2i)/2}.

Because det(𝕂XD,r)=𝒪XD\det(\mathbb{K}_{X\setminus D,r})=\mathcal{O}_{X\setminus D}, we choose hdet(E)=1h_{\det(E)}=1. Then, c(𝐯P)=r(r+1)/2c({\boldsymbol{v}}_{P})=-r(r+1)/2.

We consider the GrG_{r}-action on LiL_{i} by aui=aiuia\bullet u_{i}=a^{i}u_{i}, which induces a GrG_{r}-action on EE. Let HarmGr(E,θ,hdet(E))\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) denote the set of GrG_{r}-invariant harmonic metrics hh of (E,θ)(E,\theta) such that det(h)=hdet(E)\det(h)=h_{\det(E)}. We recall that DD is assumed to be a finite subset of XX.

Proposition 1.14 (Proposition 7.5, Proposition 7.6).

The following holds for any hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}).

  • For any PD>0P\in D_{>0}, there exists 𝒃P(h)𝒫(q,P,𝒗P){\boldsymbol{b}}_{P}(h)\in\mathcal{P}(q,P,{\boldsymbol{v}}_{P}) determined by the following condition.

    bP,i(h)=inf{b||zP|b|vP,i|h is bounded}.b_{P,i}(h)=\inf\Bigl{\{}b\in{\mathbb{R}}\,\Big{|}\,\mbox{\rm$|z_{P}|^{b}|v_{P,i}|_{h}$ is bounded}\Bigr{\}}.
  • For any PDessP\in D_{\mathop{\rm ess}\nolimits} and any I𝒮(q,P)I\in\mathcal{S}(q,P), there exist 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} and ϵ>0\epsilon>0 such that the following estimates hold as |zP|0|z_{P}|\to 0 on {|arg(αIzPρ(I))π|<(1δ)π/2}\bigl{\{}|\arg(\alpha_{I}z_{P}^{-\rho(I)})-\pi|<(1-\delta)\pi/2\bigr{\}} for any δ>0\delta>0:

    log|vi|h+aI,i(h)Re(αIzPρ(I))=O(|zP|ρ(I)+ϵ)\log|v_{i}|_{h}+a_{I,i}(h)\mathop{\rm Re}\nolimits(\alpha_{I}z_{P}^{-\rho(I)})=O\bigl{(}|z_{P}|^{-\rho(I)+\epsilon}\bigr{)}

    (See §1.2.2 for αI\alpha_{I} and ρ(I)\rho(I) for a special interval II.)

Moreover, if hiHarmGr(E,θ,hdet(E))h_{i}\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) satisfy 𝐛P(h1)=𝐛P(h2){\boldsymbol{b}}_{P}(h_{1})={\boldsymbol{b}}_{P}(h_{2}) (PD>0)(P\in D_{>0}) and 𝐚I(h1)=𝐚I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) (IPDess𝒮(q,P))(I\in\coprod_{P\in D_{\mathop{\rm ess}\nolimits}}\mathcal{S}(q,P)), then h1h_{1} and h2h_{2} are mutually bounded on NDN\setminus D for any relatively compact neighbourhood NN of DD.

Let Z(qr1)Z(q_{\leq r-1}) denote the zero set of qr1q_{\leq r-1}, which is assumed to be finite. We set D~=DZ(qr1)\widetilde{D}=D\cup Z(q_{\leq r-1}). Note that ψi\psi_{i} (i=1,,r1)(i=1,\ldots,r-1) induce isomorphisms KXD~1(Li+1/Li)|XD~K_{X\setminus\widetilde{D}}^{-1}\simeq(L_{i+1}/L_{i})_{|X\setminus\widetilde{D}}. Hence, for any hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}), we obtain Kähler metrics g(h)ig(h)_{i} (i=1,,r1)(i=1,\ldots,r-1) induced by the restrictions of hh to Li|XD~L_{i|X\setminus\widetilde{D}} and Li+1|XD~L_{i+1|X\setminus\widetilde{D}}. Let

HarmGr(E,θ,hdet(E);D,c)\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits)

denote the set of hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) such that for any relatively compact open neighbourhood NN of D~\widetilde{D}, the metrics g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are complete.

By Proposition 1.14, we obtain the map

HarmGr(E,θ,hdet(E);D,c)PD>0𝒫(q,P,𝒗P)×PDessI𝒮(q,P)𝒫.\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits)\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}(q,P,{\boldsymbol{v}}_{P})\times\prod_{P\in D_{\mathop{\rm ess}\nolimits}}\prod_{I\in\mathcal{S}(q,P)}\mathcal{P}. (11)
Theorem 1.15 (Theorem 7.11).

The map (11) is a bijection.

1.6. Examples on {\mathbb{C}}

Let γ(z)\gamma(z) be any non-zero polynomial. We set q=γ(z)e1z(dz)rq=\gamma(z)e^{\sqrt{-1}z}(dz)^{r}. It is easy to see that {}×{0<arg(z)<π}\{\infty\}\times\{0<\arg(z)<\pi\} is the unique special interval with respect to qq in this case. Hence, we obtain the following proposition as a corollary of Theorem 1.9.

Proposition 1.16 (A special case of Proposition 6.12).

We have the bijection Harm(q)𝒫\mathop{\rm Harm}\nolimits(q)\simeq\mathcal{P}. ∎

More generally, let γi(z)\gamma_{i}(z) (i=1,,m)(i=1,\ldots,m) be non-zero polynomials. Let αi\alpha_{i} (i=1,,m)(i=1,\ldots,m) be mutually distinct complex numbers. We set q=i=1mγi(z)e1αiz(dz)rq=\sum_{i=1}^{m}\gamma_{i}(z)e^{\sqrt{-1}\alpha_{i}z}(dz)^{r}.

Proposition 1.17 (A special case of Proposition 6.14).

If there exists a non-zero complex number α\alpha such that αi/α>0\alpha_{i}/\alpha\in{\mathbb{R}}_{>0} (i=1,,m)(i=1,\ldots,m), then there exists a bijection Harm(q)𝒫\mathop{\rm Harm}\nolimits(q)\simeq\mathcal{P}. Otherwise, Harm(q)={hc}\mathop{\rm Harm}\nolimits(q)=\{h^{\mathop{\rm c}\nolimits}\}. ∎

Corollary 1.18.

For any non-zero polynomial γ(z)\gamma(z), we obtain that Harm(γ(z)cos(z)(dz)r)={hc}\mathop{\rm Harm}\nolimits\bigl{(}\gamma(z)\cos(z)(dz)^{r}\bigr{)}=\{h^{\mathop{\rm c}\nolimits}\}. ∎

We may apply Theorem 1.10 to the case q=f(dz)rq=f(dz)^{r} in which ff is the product of a non-zero polynomial and a non-zero solution of a linear differential equation with polynomial coefficients. For instance, let Ai(z)\mathop{\rm Ai}\nolimits(z) be the Airy function given as in (129) below, which is a solution of the differential equation z2uzu=0\partial_{z}^{2}u-zu=0. Let γ(z)\gamma(z) be any non-zero polynomial. For the rr-differential q=γ(z)Ai(z)(dz)rq=\gamma(z)\mathop{\rm Ai}\nolimits(z)(dz)^{r}, {π/3<arg(z)<π/3}\{-\pi/3<\arg(z)<\pi/3\} is the unique special interval in this case. (See (130). See [47, §23] for a more detailed asymptotic expansion of Ai(z)\mathop{\rm Ai}\nolimits(z).) Hence, we obtain the following.

Proposition 1.19 (Proposition 6.16).

If q=γ(z)Ai(z)(dz)rq=\gamma(z)\mathop{\rm Ai}\nolimits(z)(dz)^{r}, we have a bijection Harm(q)𝒫\mathop{\rm Harm}\nolimits(q)\simeq\mathcal{P}. ∎

1.7. Acknowledgement

This study starts from a discussion during the workshop “Higgs bundles and related topics” in University of Nice, in 2017. The first author is partially supported by the National Key R&\&D Program of China No. 2022YFA1006600, the Fundamental Research Funds for the Central Universities and Nankai Zhide foundation. The second author is grateful to Martin Guest and Claus Hertling for discussions on harmonic bundles and Toda equations. The second author is partially supported by the Grant-in-Aid for Scientific Research (S) (No. 16H06335), the Grant-in-Aid for Scientific Research (S) (No. 17H06127), the Grant-in-Aid for Scientific Research (A) (No. 21H04429), the Grant-in-Aid for Scientific Research (A) (No. 22H00094), the Grant-in-Aid for Scientific Research (A) (No. 23H00083), the Grant-in-Aid for Scientific Research (C) (No. 15K04843), and the Grant-in-Aid for Scientific Research (C) (No. 20K03609), Japan Society for the Promotion of Science. The second author is also partially supported by the Research Institute for Mathematical Sciences, an International Joint Usage/Research Center located in Kyoto University.

2. Some existence results of harmonic metrics

2.1. Dirichlet problem

Let XX be any Riemann surface. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on XX. Let YXY\subset X be a relatively compact connected open subset with smooth boundary Y\partial Y. Assume that Y\partial Y is non-empty. Let hYh_{\partial Y} be any Hermitian metric of E|YE_{|\partial Y}. The following proposition is essentially due to Donaldson [9].

Proposition 2.1 (Donaldson).

There exists a unique harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that h|Y=hYh_{|\partial Y}=h_{\partial Y}.

Proof   Donaldson proved this theorem in the case where YY is a disc in [9]. The general case is similar. We shall give a proof for the convenience of the readers. We may assume that XX is an open Riemann surface. According to [16], there exists a nowhere vanishing holomorphic 11-form τ\tau om XX. Let ff be the automorphism of EE determined by θ=fτ\theta=f\,\tau. We consider the Kähler metric gX=ττ¯g_{X}=\tau\,\overline{\tau} of XX.

Let Γ\Gamma be a lattice of {\mathbb{C}} and let TT be a real 22-dimensional torus obtained as /Γ{\mathbb{C}}/\Gamma. We set gT=dzdz¯g_{T}=dz\,d\overline{z}. We set X~:=X×T\widetilde{X}:=X\times T with the projection p:X~Xp:\widetilde{X}\longrightarrow X. It is equipped with the flat Kähler metric gX~g_{\widetilde{X}} induced by gTg_{T} and gXg_{X}. We set Y~:=p1(Y)\widetilde{Y}:=p^{-1}(Y).

Let E~\widetilde{E} be the pull back of EE with the holomorphic structure p(¯E)+p(f)dz¯p^{\ast}(\overline{\partial}_{E})+p^{\ast}(f)\,d\overline{z}. According to the dimensional reduction of Hitchin, a Hermitian metric hh of E|YE_{|Y} is a harmonic metric of (E,¯E,θ)|Y(E,\overline{\partial}_{E},\theta)_{|Y} if and only if ΛY~R(ph)=0\Lambda_{\widetilde{Y}}R(p^{\ast}h)=0. According to a theorem of Donaldson [9], there exists a unique Hermitian metric h~\widetilde{h} of E~\widetilde{E} such that ΛY~R(h~)=0\Lambda_{\widetilde{Y}}R(\widetilde{h})=0 and that h~|Y=p(hY)\widetilde{h}_{|\partial Y}=p^{\ast}(h_{\partial Y}). By the uniqueness, h~\widetilde{h} is TT-invariant. Hence, there uniquely exists a harmonic metric hh of (E,¯E,θ)|Y(E,\overline{\partial}_{E},\theta)_{|Y} which induces h~\widetilde{h}. It satisfies h|Y=hYh_{|\partial Y}=h_{\partial Y}. ∎

Let h0h_{0} be a Hermitian metric of EE. Assume that det(h0)\det(h_{0}) is flat.

Corollary 2.2.

There exists a unique harmonic metric hh of E|YE_{|Y} such that h|Y=h0|Yh_{|\partial Y}=h_{0|\partial Y} and that det(h)=det(h0)|Y\det(h)=\det(h_{0})_{|Y}.

Proof   There exists a unique harmonic metric hh such that h|Y=h0|Yh_{|\partial Y}=h_{0|\partial Y}. We obtain det(h)|Y=det(h0)|Y\det(h)_{|\partial Y}=\det(h_{0})_{|\partial Y}. Note that both det(h)\det(h) and det(h0)|Y\det(h_{0})_{|Y} are flat. By the uniqueness in Proposition 2.1, we obtain det(h)=det(h0)|Y\det(h)=\det(h_{0})_{|Y}. ∎

2.1.1. Homogeneous case

Let GG be a compact Lie group. Let κ:GS1\kappa:G\longrightarrow S^{1} be a character. Suppose that XX is equipped with a GG-action, which can be trivial. A Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is called (G,κ)(G,\kappa)-homogeneous if (E,¯E)(E,\overline{\partial}_{E}) is GG-equivariant, and g(θ)=κ(g)θg^{\ast}(\theta)=\kappa(g)\theta for any gGg\in G.

Lemma 2.3.

If YY and hYh_{\partial Y} are GG-invariant in Proposition 2.1, the harmonic metric hh is also GG-invariant.

Proof   For any gGg\in G, ghg^{\ast}h is a harmonic metric of (E,¯E,κ(g)θ)(E,\overline{\partial}_{E},\kappa(g)\theta) such that (gh)|Y=hY(g^{\ast}h)_{|\partial Y}=h_{\partial Y}. Because |κ(g)|=1|\kappa(g)|=1, g(h)g^{\ast}(h) is also a harmonic metric of (E,¯E,θ)(E,\overline{\partial}_{E},\theta). By the uniqueness, we obtain g(h)=hg^{\ast}(h)=h. ∎

2.1.2. Donaldson functional

Let (E|Y,hY)\mathcal{H}(E_{|Y},h_{\partial Y}) be the space of CC^{\infty}-Hermitian metrics hh of E|YE_{|Y} such that h|Y=hYh_{|\partial Y}=h_{\partial Y}. For two metrics h1,h2(E|Y,hY)h_{1},h_{2}\in\mathcal{H}(E_{|Y},h_{\partial Y}), let uu be the automorphism of EE which is self-adjoint with respect to both hih_{i}, determined by h2=h1euh_{2}=h_{1}e^{u}. Then, we define

M(h1,h2):=1YTr(uΛF(h1))+Yh1(Ψ(u)((¯E+θ)u),(¯E+θ)u).M(h_{1},h_{2}):=\sqrt{-1}\int_{Y}\mathop{\rm Tr}\nolimits(u\Lambda F(h_{1}))+\int_{Y}h_{1}\Bigl{(}\Psi(u)\bigl{(}(\overline{\partial}_{E}+\theta)u\bigr{)},(\overline{\partial}_{E}+\theta)u\Bigr{)}.

(See §2.4 below for F(h1)F(h_{1}). See [38, §4] for Ψ(u)\Psi(u).)

Lemma 2.4.

Let h0h_{0} be any element of (E|Y,hY)\mathcal{H}(E_{|Y},h_{\partial Y}). Let hh be the harmonic metric as in Proposition 2.1. Then, we obtain M(h0,h)0M(h_{0},h)\leq 0.

Proof   By the dimensional reduction as in the proof of Proposition 2.1, the claim is reduced to [33, Proposition 2.18]. ∎

2.2. Convergence

Let XX be an open Riemann surface. Let GG be a compact Lie group acting on XX. Let κ:GS1\kappa:G\longrightarrow S^{1} be a character. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a (G,κ)(G,\kappa)-homogeneous Higgs bundle on XX. Let h0h_{0} be any GG-invariant Hermitian metric of EE.

Definition 2.5.

An exhaustive family {Xi}\{X_{i}\} of a Riemann surface XX means an increasing sequence of relatively compact GG-invariant open subsets X1X2X_{1}\subset X_{2}\subset\cdots of XX such that X=XiX=\bigcup X_{i}. The family is called smooth if Xi\partial X_{i} are smooth.

Let {Xi}\{X_{i}\} be a smooth exhaustive family of XX. The restriction h0|Xih_{0|X_{i}} is denoted by h0,ih_{0,i}. Let hih_{i} (i=1,2,)(i=1,2,\ldots) be GG-invariant harmonic metrics of (E,¯E,θ)|Xi(E,\overline{\partial}_{E},\theta)_{|X_{i}}. Let sis_{i} be the automorphism of E|XiE_{|X_{i}} determined by hi=h0,isih_{i}=h_{0,i}\cdot s_{i}. Let ff be an >0{\mathbb{R}}_{>0}-valued function on XX such that each f|Xif_{|X_{i}} is bounded.

Proposition 2.6.

Assume that |si|h0,i+|si1|h0,if|Xi|s_{i}|_{h_{0,i}}+|s^{-1}_{i}|_{h_{0,i}}\leq f_{|X_{i}} for any ii. Then, there exists a subsequence si(j)s_{i(j)} which is convergent to a GG-invariant automorphism ss_{\infty} of EE on any relatively compact subset of XX in the CC^{\infty}-sense. As a result, we obtain a GG-invariant harmonic metric h=h0sh_{\infty}=h_{0}s_{\infty} of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) as the limit of the subsequence hi(j)h_{i(j)}. Moreover, we obtain |s|h0+|s1|h0f|s_{\infty}|_{h_{0}}+|s_{\infty}^{-1}|_{h_{0}}\leq f. In particular, if ff is bounded, h0h_{0} and hh_{\infty} are mutually bounded.

Proof   We explain an outline of the proof. Let gXg_{X} be a Kähler metric of XX. According to a general formula (19) below, the following holds on any XiX_{i}:

1Λ¯Tr(si)=Tr(siΛF(h0,i))|(¯+θ)(si)si1/2|h0,i,g2.\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s_{i})=-\mathop{\rm Tr}\nolimits\bigl{(}s_{i}\Lambda F(h_{0,i})\bigr{)}-\bigl{|}(\overline{\partial}+\theta)(s_{i})\cdot s_{i}^{-1/2}\bigr{|}^{2}_{h_{0,i},g}. (12)

Let KK be any compact subset of XX. Let NN be a relatively compact neighbourhood of KK in XX. Let χ:X0\chi:X\longrightarrow{\mathbb{R}}_{\geq 0} be a CC^{\infty}-function such that (i) χ|K=1\chi_{|K}=1, (ii) χ|XN=0\chi_{|X\setminus N}=0, (iii) χ1/2χ\chi^{-1/2}\partial\chi and χ1/2¯χ\chi^{-1/2}\overline{\partial}\chi on {PX|χ(P)>0}\{P\in X\,|\,\chi(P)>0\} induces a CC^{\infty}-function on XX.

There exist i0i_{0} such that NN is a relatively compact open subset of XiX_{i} for any ii0i\geq i_{0}. We obtain the following:

1Λ¯(χTr(si))=χ1Λ¯Tr(si)+(1Λ¯χ)Tr(si)+1Λ(¯χTr(si))1Λ(χ¯Tr(si)).\sqrt{-1}\Lambda\overline{\partial}\partial(\chi\mathop{\rm Tr}\nolimits(s_{i}))=\chi\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s_{i})+(\sqrt{-1}\Lambda\overline{\partial}\partial\chi)\cdot\mathop{\rm Tr}\nolimits(s_{i})\\ +\sqrt{-1}\Lambda(\overline{\partial}\chi\partial\mathop{\rm Tr}\nolimits(s_{i}))-\sqrt{-1}\Lambda(\partial\chi\overline{\partial}\mathop{\rm Tr}\nolimits(s_{i})). (13)

Note that |¯Esi|hi,gX=|E,hisi|hi,gX|\overline{\partial}_{E}s_{i}|_{h_{i},g_{X}}=|\partial_{E,h_{i}}s_{i}|_{h_{i},g_{X}}, and that

|X1Λ(¯χTr(si))|(X|χ1/2¯χ|2)1/2(Xχ|E,hisi|h0,gX2)1/2.\left|\int_{X}\sqrt{-1}\Lambda(\overline{\partial}\chi\partial\mathop{\rm Tr}\nolimits(s_{i}))\right|\leq\left(\int_{X}|\chi^{-1/2}\overline{\partial}\chi|^{2}\right)^{1/2}\left(\int_{X}\chi|\partial_{E,h_{i}}s_{i}|^{2}_{h_{0},g_{X}}\right)^{1/2}\!\!\!\!. (14)

Note that there exists C0>0C_{0}>0 such that |si|h0+|si1|h0C0|s_{i}|_{h_{0}}+|s_{i}^{-1}|_{h_{0}}\leq C_{0} on NN for any ii. By (12), (13) and (14), there exist Cj>0C_{j}>0 (j=1,2)(j=1,2) such that the following holds for any sufficiently large ii:

χ|¯Esi|h0,gX2+χ|[θ,si]|h0,gX2C1+C2(χ|¯Esi|h0,gX2+χ|[θ,si]|h0,gX2)1/2.\int\chi\bigl{|}\overline{\partial}_{E}s_{i}\bigr{|}^{2}_{h_{0},g_{X}}+\int\chi\bigl{|}[\theta,s_{i}]\bigr{|}^{2}_{h_{0},g_{X}}\leq\\ C_{1}+C_{2}\left(\int\chi\bigl{|}\overline{\partial}_{E}s_{i}\bigr{|}^{2}_{h_{0},g_{X}}+\int\chi\bigl{|}[\theta,s_{i}]\bigr{|}^{2}_{h_{0},g_{X}}\right)^{1/2}. (15)

Therefore, there exists C3>0C_{3}>0 such that the following holds for any sufficiently large ii:

χ|¯Esi|h0,gX2+χ|[θ,si]|h0,gX2C3.\int\chi\bigl{|}\overline{\partial}_{E}s_{i}\bigr{|}^{2}_{h_{0},g_{X}}+\int\chi\bigl{|}[\theta,s_{i}]\bigr{|}^{2}_{h_{0},g_{X}}\leq C_{3}.

We obtain the boundedness of the L2L^{2}-norms of ¯Esi\overline{\partial}_{E}s_{i} and E,hisi\partial_{E,h_{i}}s_{i} (ii0)(i\geq i_{0}) on KK with respect to h0h_{0} and gXg_{X}. By a variant of Simpson’s main estimate (see [32, Proposition 2.1]), we obtain the boundedness of the sup norms of θ\theta on NN with respect to hih_{i} and gXg_{X}. By the Hitchin equation, we obtain the boundedness of the sup norms of ¯E(si1E,hisi)\overline{\partial}_{E}(s_{i}^{-1}\partial_{E,h_{i}}s_{i}) on NN with respect to hih_{i} and gXg_{X}. By using the elliptic regularity, we obtain that the L1pL_{1}^{p}-norms of si1E,hi(si)s_{i}^{-1}\partial_{E,h_{i}}(s_{i}) on a relatively compact neighbourhood of KK are bounded for any p>1p>1. It follows that L2pL_{2}^{p}-norms of sis_{i} on a relatively compact neighbourhood of KK are bounded for any pp. Hence, a subsequence of sis_{i} is weakly convergent in L2pL_{2}^{p} on a relatively compact neighbourhood of KK. By the bootstrapping argument using a general formula (18) below, we obtain that the sequence is convergent on a relatively compact neighbourhood of KK in the CC^{\infty}-sense. By using the diagonal argument, we obtain that a subsequence of sis_{i} is weakly convergent in CC^{\infty}-sense on any compact subset. ∎

Let us give a complement to Proposition 2.6.

Proposition 2.7.

Suppose that hi|Xi=h0|Xih_{i|\partial X_{i}}=h_{0|\partial X_{i}} and that det(h0)\det(h_{0}) is flat. Then, in Proposition 2.6, we obtain det(s)=1\det(s_{\infty})=1. Moreover, if X|F(h0)|h0,gX<\int_{X}|F(h_{0})|_{h_{0},g_{X}}<\infty, and if ff is bounded, then |(¯E+θ)(s)|h0,gX|(\overline{\partial}_{E}+\theta)(s_{\infty})|_{h_{0},g_{X}} is L2L^{2}, where gXg_{X} denotes any Kähler metric of XX.

Proof   Because det(si)=1\det(s_{i})=1, the first claim is clear. Let gXg_{X} be any Kähler metric of XX. Note that Tr(si)rank(E)\mathop{\rm Tr}\nolimits(s_{i})\geq\mathop{\rm rank}\nolimits(E). We obtain the following by using Green formula:

Xi1Λ¯Tr(si)0.\int_{X_{i}}\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s_{i})\geq 0. (16)

We obtain the following from (12) and (16):

Xi|(¯E+θ)(si)si1/2|h0,gX2Xi1Tr(siΛF(h0)).\int_{X_{i}}\bigl{|}(\overline{\partial}_{E}+\theta)(s_{i})\cdot s_{i}^{-1/2}\bigr{|}^{2}_{h_{0},g_{X}}\leq-\int_{X_{i}}\sqrt{-1}\mathop{\rm Tr}\nolimits\bigl{(}s_{i}\Lambda F(h_{0})\bigr{)}.

We obtain the claim of the corollary by Fatou’s lemma. ∎

Remark 2.8.

In [34] and [33], a Hermitian-Einstein metric of a holomorphic vector bundle (E,¯E)(E,\overline{\partial}_{E}) on a Kähler manifold YY is constructed as a limit of Hermitian-Einstein metrics (E,¯E)|Yi(E,\overline{\partial}_{E})_{|Y_{i}} for an exhaustive family YiY_{i} of YY. For the proof of Proposition 2.6, we may also apply an argument in [34] and [33, §2.8] by using the dimensional reduction.

2.3. A Kobayashi-Hitchin correspondence

Let XX be an open Riemann surface. Let gXg_{X} be any Kähler metric of XX. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a (G,κ)(G,\kappa)-homogeneous Higgs bundle on XX. Let h0h_{0} be a GG-invariant Hermitian metric of EE such that det(h0)\det(h_{0}) is flat.

Condition 2.9.

The support of F(h0)F(h_{0}) is compact.

The condition clearly implies that |F(h0)||F(h_{0})| is integrable on XX. For any Higgs subbundle EE^{\prime} of EE, let h0h_{0}^{\prime} be the induced metric of EE^{\prime}, and we set

degan(E,h0):=1ΛTrF(h0).\deg^{\mathop{\rm an}\nolimits}(E^{\prime},h_{0}):=\int\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits F(h_{0}^{\prime}).

Recall the Chern-Weil formula [38, Lemma 3.2]:

degan(E,h0)=1Tr(pEΛF(h0))|(¯E+θ)pE|h0,gX2,\deg^{\mathop{\rm an}\nolimits}(E^{\prime},h_{0})=\int\sqrt{-1}\mathop{\rm Tr}\nolimits(p_{E^{\prime}}\Lambda F(h_{0}))-\int\bigl{|}(\overline{\partial}_{E}+\theta)p_{E^{\prime}}\bigr{|}^{2}_{h_{0},g_{X}},

where pEp_{E^{\prime}} denote the orthogonal projection of EE onto EE^{\prime}. By Condition 2.9, degan(E,h0)\deg^{\mathop{\rm an}\nolimits}(E^{\prime},h_{0}) is well defined in {}{\mathbb{R}}\cup\{-\infty\}. Because det(h0)\det(h_{0}) is assumed to be flat, we obtain Tr(F(h0))=0\mathop{\rm Tr}\nolimits(F(h_{0}))=0, and hence degan(E,h0)=0\deg^{\mathop{\rm an}\nolimits}(E,h_{0})=0.

Definition 2.10.

(E,¯E,θ,h0)(E,\overline{\partial}_{E},\theta,h_{0}) is analytically stable with respect to the GG-action if degan(E,h0)<0\deg^{\mathop{\rm an}\nolimits}(E^{\prime},h_{0})<0 for any proper GG-equivariant Higgs subbundle EE^{\prime}.

Proposition 2.11.

If (E,¯E,θ,h0)(E,\overline{\partial}_{E},\theta,h_{0}) is analytically stable with respect to the GG-action, there exists a GG-invariant harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that (i) h0h_{0} and hh are mutually bounded, (ii) det(h)=det(h0)\det(h)=\det(h_{0}), (iii) (¯E+θ)(hh01)(\overline{\partial}_{E}+\theta)(h\cdot h_{0}^{-1}) is L2L^{2}.

Proof   Let N0N_{0} be a relatively compact neighbourhood of the support of F(h0)F(h_{0}). Let N1N_{1} be a relatively compact neighbourhood of N¯0\overline{N}_{0}. Let X1X2X_{1}\subset X_{2}\subset\cdots be a smooth exhaustive sequence of XX such that N¯1X1\overline{N}_{1}\subset X_{1}. We set h0,i:=h0|Xih_{0,i}:=h_{0|X_{i}}.

There exists a harmonic metric hih_{i} of (E,¯E,θ)|Xi(E,\overline{\partial}_{E},\theta)_{|X_{i}} such that hi|Xi=h0|Xih_{i|\partial X_{i}}=h_{0|\partial X_{i}}. Let sis_{i} be the automorphism of E|XiE_{|X_{i}} determined by hi=h0,isih_{i}=h_{0,i}\cdot s_{i}. Note that det(si)=1\det(s_{i})=1. According to the inequality (21) below, we obtain

1Λ¯logTr(si)|F(h0,i)|h0,gX.\sqrt{-1}\Lambda\overline{\partial}\partial\log\mathop{\rm Tr}\nolimits(s_{i})\leq\bigl{|}F(h_{0,i})\bigr{|}_{h_{0},g_{X}}. (17)

Note that Trsi|Xi=rank(E)\mathop{\rm Tr}\nolimits s_{i|\partial X_{i}}=\mathop{\rm rank}\nolimits(E) and that F(h0,i)=0F(h_{0,i})=0 on XiN0X_{i}\setminus N_{0}. By (17), we obtain the following

maxXilogTr(si)max{maxN0logTr(si),rank(E)}.\max_{X_{i}}\log\mathop{\rm Tr}\nolimits(s_{i})\leq\max\Bigl{\{}\max_{N_{0}}\log\mathop{\rm Tr}\nolimits(s_{i}),\mathop{\rm rank}\nolimits(E)\Bigr{\}}.

Let logTr(si)h0,N1\|\log\mathop{\rm Tr}\nolimits(s_{i})\|_{h_{0},N_{1}} denote the L1L^{1}-norm of logTr(si)|N1\log\mathop{\rm Tr}\nolimits(s_{i})_{|N_{1}} with respect to h0h_{0} and gXg_{X}. By using the argument in the proof of [38, Proposition 2.1] together with (17), we obtain Cj>0C_{j}>0 (j=0,1)(j=0,1) depending only on |F(h0)|h0,gX|F(h_{0})|_{h_{0},g_{X}} such that

maxN0|logTr(si)|h0C0logTr(si)h0,N1+C1.\max_{N_{0}}|\log\mathop{\rm Tr}\nolimits(s_{i})|_{h_{0}}\leq C_{0}\|\log\mathop{\rm Tr}\nolimits(s_{i})\|_{h_{0},N_{1}}+C_{1}.

Let uiu_{i} be the endomorphism of E|XiE_{|X_{i}} which is self-adjoint with respect to h0,ih_{0,i} and hih_{i} determined by si=euis_{i}=e^{u_{i}}. There exists CjC_{j} (j=2,3)(j=2,3) depending only on rank(E)\mathop{\rm rank}\nolimits(E) such that

|ui|h0C2(log|si|h0+1),log|si|h0C3(|ui|h0+1).|u_{i}|_{h_{0}}\leq C_{2}(\log|s_{i}|_{h_{0}}+1),\quad\quad\log|s_{i}|_{h_{0}}\leq C_{3}(|u_{i}|_{h_{0}}+1).

Let uih0,N1\|u_{i}\|_{h_{0},N_{1}} denote the L1L^{1}-norm of ui|N1u_{i|N_{1}} with respect to h0h_{0} and gXg_{X}. There exists CjC_{j} (j=4,5)(j=4,5), depending only on rank(E)\mathop{\rm rank}\nolimits(E) and |F(h0)||F(h_{0})| such that

maxXi|ui|h0C4uih0,N1+C5.\max_{X_{i}}|u_{i}|_{h_{0}}\leq C_{4}\|u_{i}\|_{h_{0},N_{1}}+C_{5}.

The rest is almost the same as the proof of [33, Theorem 2.5], particularly [33, Proposition 2.32] and [33, §2.7.4]. We explain only an outline.

Suppose that supXi|si|h0,i\sup_{X_{i}}|s_{i}|_{h_{0,i}}\to\infty as ii\to\infty. It implies uih0,N1\|u_{i}\|_{h_{0},N_{1}}\to\infty. We set vi:=ui/uih0,N1v_{i}:=u_{i}/\|u_{i}\|_{h_{0},N_{1}}. They are endomorphisms of E|XiE_{|X_{i}} which are self adjoint with respect to hih_{i} and h0,ih_{0,i}. We can prove the following lemma by using the argument in [38, Lemma 5.4].

Lemma 2.12.

There exists a GG-invariant L12L_{1}^{2}-section vv_{\infty} of End(E)\mathop{\rm End}\nolimits(E) on XX such that the following holds.

  • v0v_{\infty}\neq 0.

  • A subsequence of {vi}\{v_{i}\} is weakly convergent to vv_{\infty} in L12L_{1}^{2} on any compact subset of XX.

  • Let Φ:×>0\Phi:{\mathbb{R}}\times{\mathbb{R}}\longrightarrow{\mathbb{R}}_{>0} be a CC^{\infty}-function such that Φ(y1,y2)<(y1y2)1\Phi(y_{1},y_{2})<(y_{1}-y_{2})^{-1} if y1>y2y_{1}>y_{2}. Then, the following holds:

    1XTr(vΛF(h0))+Xh0(Φ(v)((¯E+θ)v),(¯E+θ)v)0.\sqrt{-1}\int_{X}\mathop{\rm Tr}\nolimits\bigl{(}v_{\infty}\Lambda F(h_{0})\bigr{)}+\int_{X}h_{0}\Bigl{(}\Phi(v_{\infty})\bigl{(}(\overline{\partial}_{E}+\theta)v_{\infty}\bigr{)},(\overline{\partial}_{E}+\theta)v_{\infty}\Bigr{)}\leq 0.

    (See [38, §4] for Φ(v)\Phi(v_{\infty}).) ∎

By applying Lemma 2.12 with the argument in the proof of [38, Lemma 5.5], we obtain that the eigenvalues of vv_{\infty} are constant. Let λ1<<λ\lambda_{1}<\cdots<\lambda_{\ell} be the eigenvalues of vv_{\infty}. As the orthogonal projections onto the sum of the eigen spaces associated with λ1,,λi\lambda_{1},\ldots,\lambda_{i}, we obtain an L12L_{1}^{2}-section πi\pi_{i} of End(E)\mathop{\rm End}\nolimits(E) for which πi2=πi\pi_{i}^{2}=\pi_{i} and (idEπi)(¯E+θ)πi=0(\mathop{\rm id}\nolimits_{E}-\pi_{i})\circ(\overline{\partial}_{E}+\theta)\pi_{i}=0. (See [38, §4] for more precise construction of πi\pi_{i}.) According to the regularity of Higgs L12L_{1}^{2}-subbundle [38, Proposition 5.8], there exists a Higgs subbundle EiEE_{i}\subset E such that the orthogonal projection onto EiE_{i} is equal to πi\pi_{i}. By the argument in the proof of [38, Proposition 5.3], we obtain deg(Ei,h0)>0\deg(E_{i},h_{0})>0 for one of ii, which contradicts with the stability condition. Hence, there exists C>0C>0, which is independent of ii, such that |si|h0,i<C|s_{i}|_{h_{0,i}}<C. Then, the claim of Proposition 2.11 follows from Proposition 2.6. ∎

2.4. Appendix

We recall some fundamental formulas due to Simpson [38, Lemma 3.1] for the convenience of the readers. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on a Riemann surface XX. For a Hermitian metric hh of EE, we obtain the Chern connection h=¯E+E,h\nabla_{h}=\overline{\partial}_{E}+\partial_{E,h} of (E,¯E,h)(E,\overline{\partial}_{E},h). The curvature of h\nabla_{h} is denoted by R(h)R(h). We also obtain the adjoint θh\theta^{\dagger}_{h} of θ\theta with respect to hh. The curvature of h+θ+θh\nabla_{h}+\theta+\theta_{h}^{\dagger} is denoted by F(h)F(h), i.e., F(h)=R(h)+[θ,θh]F(h)=R(h)+[\theta,\theta^{\dagger}_{h}].

Let hih_{i} (i=1,2)(i=1,2) be Hermitian metrics of EE. We obtain the automorphism ss of EE determined by h2=h1sh_{2}=h_{1}\cdot s. Let gg be a Kähler metric of XX, let Λ\Lambda denote the adjoint of the multiplication of the associated Kähler form. Then, according to [38, Lemma 3.1 (a)], we obtain the following on XX:

1Λ(¯E+θ)(E,h1+θh1)s=s1Λ(F(h2)F(h1))+1Λ((¯E+θ)(s)s1(E,h1+θh1)(s)).\sqrt{-1}\Lambda\bigl{(}\overline{\partial}_{E}+\theta\bigr{)}\circ\bigl{(}\partial_{E,h_{1}}+\theta^{\dagger}_{h_{1}}\bigr{)}s=\\ s\sqrt{-1}\Lambda\bigl{(}F(h_{2})-F(h_{1})\bigr{)}+\sqrt{-1}\Lambda\Bigl{(}\bigl{(}\overline{\partial}_{E}+\theta\bigr{)}(s)s^{-1}\bigl{(}\partial_{E,h_{1}}+\theta^{\dagger}_{h_{1}}\bigr{)}(s)\Bigr{)}. (18)

By taking the trace, and by using [38, Lemma 3.1 (b)], we obtain

1Λ¯Tr(s)=1Tr(sΛ(F(h2)F(h1)))|(¯E+θ)(s)s1/2|h1,g2.\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s)=\sqrt{-1}\mathop{\rm Tr}\nolimits\Bigl{(}s\Lambda\bigl{(}F(h_{2})-F(h_{1})\bigr{)}\Bigr{)}-\Bigl{|}\bigl{(}\overline{\partial}_{E}+\theta\bigr{)}(s)s^{-1/2}\Bigr{|}^{2}_{h_{1},g}. (19)

Note that (¯E+θ)(s)=¯E(s)+[θ,s](\overline{\partial}_{E}+\theta)(s)=\overline{\partial}_{E}(s)+[\theta,s]. Moreover, ¯E(s)\overline{\partial}_{E}(s) is a (0,1)(0,1)-form, and [θ,s][\theta,s] is a (1,0)(1,0)-form. Hence, (19) is also rewritten as follows:

1Λ¯Tr(s)=1Tr(sΛ(F(h2)F(h1)))|[θ,s]s1/2|h1,g2|¯E(s)s1/2|h1,g.\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s)=\sqrt{-1}\mathop{\rm Tr}\nolimits\Bigl{(}s\Lambda\bigl{(}F(h_{2})-F(h_{1})\bigr{)}\Bigr{)}\\ -\bigl{|}[\theta,s]s^{-1/2}\bigr{|}^{2}_{h_{1},g}-\bigl{|}\overline{\partial}_{E}(s)s^{-1/2}\bigr{|}_{h_{1},g}. (20)

We also recall the following inequality [38, Lemma 3.1 (d)]:

1Λ¯logTr(s)|ΛF(h1)|h1+|ΛF(h2)|h2.\sqrt{-1}\Lambda\overline{\partial}\partial\log\mathop{\rm Tr}\nolimits(s)\leq\bigl{|}\Lambda F(h_{1})\bigr{|}_{h_{1}}+\bigl{|}\Lambda F(h_{2})\bigr{|}_{h_{2}}. (21)

In particular, if both hih_{i} are harmonic, the functions Tr(s)\mathop{\rm Tr}\nolimits(s) and logTr(s)\log\mathop{\rm Tr}\nolimits(s) are subharmonic:

1Λ¯Tr(s)=|(¯E+θ)(s)s1/2|h,g20,1Λ¯logTr(s)0.\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s)=-\bigl{|}(\overline{\partial}_{E}+\theta)(s)s^{-1/2}\bigr{|}^{2}_{h,g}\leq 0,\quad\sqrt{-1}\Lambda\overline{\partial}\partial\log\mathop{\rm Tr}\nolimits(s)\leq 0. (22)

3. Preliminaries for harmonic metrics of cyclic Higgs bundles

3.1. Cyclic Higgs bundles associated with rr-differentials

Let XX be any Riemann surface equipped with a line bundle KX1/2K_{X}^{1/2} and an isomorphism (KX1/2)2=KX(K_{X}^{1/2})^{2}=K_{X}. We set 𝕂X,r:=i=1rKX(r+12i)/2\mathbb{K}_{X,r}:=\bigoplus_{i=1}^{r}K_{X}^{(r+1-2i)/2}. We set Gr:={a|ar=1}G_{r}:=\{a\in{\mathbb{C}}\,|\,a^{r}=1\}. We define the GrG_{r}-actions on KX(r+12i)/2K_{X}^{(r+1-2i)/2} by av=aiva\bullet v=a^{i}v. They induce a GrG_{r}-action on 𝕂X,r\mathbb{K}_{X,r}.

For any holomorphic rr-differential qq on XX, Let θ(q)\theta(q) be obtained as θ(q)=i=1rθ(q)i\theta(q)=\sum_{i=1}^{r}\theta(q)_{i}, where θ(q)i:KX(r+12i)/2KX(r+12(i+1))/2KX\theta(q)_{i}:K_{X}^{(r+1-2i)/2}\longrightarrow K_{X}^{(r+1-2(i+1))/2}\otimes K_{X} (i=1,,r1)(i=1,\ldots,r-1) are induced by the identity, and θ(q)r:KX(r+1)/2KX(r1)/2KX\theta(q)_{r}:K_{X}^{(-r+1)/2}\longrightarrow K_{X}^{(r-1)/2}\otimes K_{X} is induced by the multiplication of qq. Let Harm(q)\mathop{\rm Harm}\nolimits(q) denote the set of GrG_{r}-invariant harmonic metrics hh of (𝕂X,r,θ(q))(\mathbb{K}_{X,r},\theta(q)) such that det(h)=1\det(h)=1. By the GrG_{r}-invariance, the decomposition 𝕂X,r=i=1rKX(r+12i)/2\mathbb{K}_{X,r}=\bigoplus_{i=1}^{r}K_{X}^{(r+1-2i)/2} is orthogonal with respect to any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), and hence we obtain the decomposition h=h|KX(r+12i)/2h=\bigoplus h_{|K_{X}^{(r+1-2i)/2}}.

Note that KX(r+12i)/2K_{X}^{(r+1-2i)/2} and KX(r+12(r+1i))/2=KX(r1+2i)/2K_{X}^{(r+1-2(r+1-i))/2}=K_{X}^{(-r-1+2i)/2} are mutually dual. We say that hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) is real if h|KX(r+12i)/2h_{|K_{X}^{(r+1-2i)/2}} and h|KX(r1+2i)/2h_{|K_{X}^{(-r-1+2i)/2}} are mutually dual. Let Harm(q)\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q) denote the subset of hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) which are real.

Note that (KX(r+12i)/2)1KX(r+12(i+1))/2\bigl{(}K_{X}^{(r+1-2i)/2}\bigr{)}^{-1}\otimes K_{X}^{(r+1-2(i+1))/2} is naturally isomorphic to the tangent bundle of XX. Hence,

h|KX(r+12i)/21h|KX(r+12(i+1))/2(i=1,,r1)h_{|K_{X}^{(r+1-2i)/2}}^{-1}\otimes h_{|K_{X}^{(r+1-2(i+1))/2}}\quad(i=1,\ldots,r-1) (23)

induce Kähler metrics of XX. If the metrics (23) induce complete distances on XX, hh is called complete.

Remark 3.1.

According to [24], there uniquely exists hcHarm(q)h^{\mathop{\rm c}\nolimits}\in\mathop{\rm Harm}\nolimits(q) which is real and complete. If qq is nowhere vanishing on XX, there exists a harmonic metric hcanh_{\mathop{\rm can}\nolimits} as in [24] (see also §3.3 below).

3.2. Hermitian metrics on a vector space with a cyclic automorphism

Let rr be a positive integer. Let VV be an rr-dimensional {\mathbb{C}}-vector space with a base 𝒆=(e0,,er1){\boldsymbol{e}}=(e_{0},\ldots,e_{r-1}). Let α\alpha be a non-zero complex number. Let ff be the automorphism of VV determined by f(ei)=ei+1f(e_{i})=e_{i+1} (i=0,1,,r2)(i=0,1,\ldots,r-2) and f(er1)=αre0f(e_{r-1})=\alpha^{r}e_{0}. We put Gr:={a|ar=1}G_{r}:=\{a\in{\mathbb{C}}\,|\,a^{r}=1\}. We consider the action of GrG_{r} on VV determined by aei=aieia\bullet e_{i}=a^{i}e_{i}.

Let (f,𝒆)\mathcal{H}(f,{\boldsymbol{e}}) be the set of GrG_{r}-invariant Hermitian metrics hh of VV such that i=0r1h(ei,ei)=1\prod_{i=0}^{r-1}h(e_{i},e_{i})=1. The GrG_{r}-invariance is equivalent to the orthogonality h(ei,ej)=0h(e_{i},e_{j})=0 (ij)(i\neq j).

We put τ:=exp(2π1/r)Gr\tau:=\exp(2\pi\sqrt{-1}/r)\in G_{r}. We set vi:=j=0r1τijαjejv_{i}:=\sum_{j=0}^{r-1}\tau^{-ij}\alpha^{-j}e_{j} (i=0,1,,r1)(i=0,1,\ldots,r-1). Then, 𝒗=(v0,,vr1){\boldsymbol{v}}=(v_{0},\ldots,v_{r-1}) is a base of VV such that f(vi)=τiαvif(v_{i})=\tau^{i}\alpha v_{i}. Note that τvi=vi1\tau\bullet v_{i}=v_{i-1} (i=1,,r1)(i=1,\ldots,r-1) and τv0=vr1\tau\bullet v_{0}=v_{r-1}. Let us remark the following well known lemma.

Lemma 3.2.

Let VVV^{\prime}\subset V be a GrG_{r}-invariant {\mathbb{C}}-subspace such that f(V)Vf(V^{\prime})\subset V^{\prime}. Then, either V=0V^{\prime}=0 or V=VV^{\prime}=V hold.

Proof   Because f(V)Vf(V^{\prime})\subset V^{\prime}, there is a subset I{0,,r1}I\subset\{0,\ldots,r-1\} such that V=iIviV^{\prime}=\bigoplus_{i\in I}{\mathbb{C}}v_{i}. Because VV^{\prime} is GrG_{r}-invariant, and because GrG_{r} acts on {v0,,vr1}\{v_{0},\ldots,v_{r-1}\} in a cyclic way, we obtain I=I=\emptyset or I={0,,r1}I=\{0,\ldots,r-1\}. ∎

Let h(f,𝒆)h\in\mathcal{H}(f,{\boldsymbol{e}}). By the GrG_{r}-invariance, b(h):=h(vi,vi)b(h):=h(v_{i},v_{i}) is independent of ii. We obtain

h(ej,ej)=b(h)r|α|2j+1r2i|α|2jτj(i)h(vi,v).h(e_{j},e_{j})=\frac{b(h)}{r}|\alpha|^{2j}+\frac{1}{r^{2}}\sum_{i\neq\ell}|\alpha|^{2j}\tau^{j(i-\ell)}h(v_{i},v_{\ell}). (24)

3.2.1. Canonical metric

Let hcan(f,𝒆)h_{\mathop{\rm can}\nolimits}\in\mathcal{H}(f,{\boldsymbol{e}}) be determined by

hcan(ej,ej)=|α|(r1)+2j(j=0,,r1).h_{\mathop{\rm can}\nolimits}(e_{j},e_{j})=|\alpha|^{-(r-1)+2j}\quad(j=0,\ldots,r-1). (25)
Lemma 3.3.

hcanh_{\mathop{\rm can}\nolimits} is the unique Hermitian metric contained in (f,𝐞)\mathcal{H}(f,{\boldsymbol{e}}) such that the base 𝐯{\boldsymbol{v}} is orthogonal with respect to hcanh_{\mathop{\rm can}\nolimits}.

Proof   We can check the orthogonality of 𝒗{\boldsymbol{v}} by a direct computation. Suppose that 𝒗{\boldsymbol{v}} is orthogonal with respect to h1(f,𝒆)h_{1}\in\mathcal{H}(f,{\boldsymbol{e}}). By (24), we obtain

h1(ej,ej)=b(h1)r|α|2j.h_{1}(e_{j},e_{j})=\frac{b(h_{1})}{r}|\alpha|^{2j}.

By the condition 1=j=0r1h1(ej,ej)1=\prod_{j=0}^{r-1}h_{1}(e_{j},e_{j}), we obtain b(h1)=r|α|(r1)b(h_{1})=r|\alpha|^{-(r-1)}, and hence h1=hcanh_{1}=h_{\mathop{\rm can}\nolimits}. ∎

3.2.2. ϵ\epsilon-orthogonality

Suppose r>1r>1, and let 0ϵ10\leq\epsilon\leq 1. Suppose that the following holds for a metric h(f,𝒆)h\in\mathcal{H}(f,{\boldsymbol{e}}):

|h(vi,vj)|ϵb(h)=ϵ|vi|h|vj|h(ij).|h(v_{i},v_{j})|\leq\epsilon b(h)=\epsilon\cdot|v_{i}|_{h}\cdot|v_{j}|_{h}\quad(i\neq j).

The second term in the right hand side of (24) is dominated by

b(h)|α|2jr(r1)ϵ.\frac{b(h)|\alpha|^{2j}}{r}(r-1)\epsilon.

Hence, we obtain the following description

h(ej,ej)=b(h)|α|2jr(1+cj),h(e_{j},e_{j})=\frac{b(h)|\alpha|^{2j}}{r}(1+c_{j}),

where |cj|(r1)ϵ|c_{j}|\leq(r-1)\epsilon and 1+cj>01+c_{j}>0. By the condition h(ej,ej)=1\prod h(e_{j},e_{j})=1, we obtain

1=(b(h)r)r|α|r(r1)j=0r1(1+cj).1=\left(\frac{b(h)}{r}\right)^{r}|\alpha|^{r(r-1)}\prod_{j=0}^{r-1}(1+c_{j}).

In particular, we obtain 1=(b(hcan)r)r|α|r(r1)1=\left(\frac{b(h_{\mathop{\rm can}\nolimits})}{r}\right)^{r}|\alpha|^{r(r-1)}. Hence, we obtain

b(hcan)=b(h)j=0r1(1+cj)1/r.b(h_{\mathop{\rm can}\nolimits})=b(h)\cdot\prod_{j=0}^{r-1}(1+c_{j})^{1/r}.

For 0<δ<10<\delta<1, we set Cδ:=(1δ)1(r1)C_{\delta}:=(1-\delta)^{-1}(r-1). If (r1)ϵ<δ(r-1)\epsilon<\delta, we obtain |log(1+cj)|Cδϵ|\log(1+c_{j})|\leq C_{\delta}\epsilon. For such ϵ\epsilon, we obtain

|log(b(h)/b(hcan))|=|1rj=0r1log(1+cj)|Cδϵ.\left|\log\bigl{(}b(h)/b(h_{\mathop{\rm can}\nolimits})\bigr{)}\right|=\left|\frac{1}{r}\sum_{j=0}^{r-1}\log(1+c_{j})\right|\leq C_{\delta}\epsilon.

We also obtain

|log(h(ej,ej)/hcan(ej,ej))|=|log(b(h)/b(hcan))+log(1+cj)|2Cδϵ.\left|\log\Bigl{(}h(e_{j},e_{j})/h_{\mathop{\rm can}\nolimits}(e_{j},e_{j})\Bigr{)}\right|=\Bigl{|}\log\bigl{(}b(h)/b(h_{\mathop{\rm can}\nolimits})\bigr{)}+\log(1+c_{j})\Bigr{|}\leq 2C_{\delta}\epsilon. (26)

3.2.3. Norm of the automorphism

For any h(f,𝒆)h\in\mathcal{H}(f,{\boldsymbol{e}}), we obtain |f|hr|α||f|_{h}\geq\sqrt{r}|\alpha|. Let CC be any positive number such that Cr|α|C\geq\sqrt{r}|\alpha|.

Lemma 3.4.

For any h(f,𝐞)h\in\mathcal{H}(f,{\boldsymbol{e}}) such that |f|hC|f|_{h}\leq C, we obtain

|ei|h|α|rC(r+1)/2+i(i=0,,r2),|er1|hC(r1)/2.|e_{i}|_{h}\leq|\alpha|^{-r}C^{(r+1)/2+i}\quad(i=0,\ldots,r-2),\quad|e_{r-1}|_{h}\leq C^{(r-1)/2}.

Proof   Because ei+1=f(ei)e_{i+1}=f(e_{i}), we obtain |ei+1|hC|ei|h|e_{i+1}|_{h}\leq C|e_{i}|_{h}. We obtain |er1|hCr1i|ei|h|e_{r-1}|_{h}\leq C^{r-1-i}|e_{i}|_{h}. Because |ei|h=1\prod|e_{i}|_{h}=1, we obtain

|er1|hrCr(r1)/21,i.e., |er1|hC(r1)/2.|e_{r-1}|_{h}^{r}C^{-r(r-1)/2}\leq 1,\quad\mbox{\rm i.e., }|e_{r-1}|_{h}\leq C^{(r-1)/2}.

Because f(er1)=αre0f(e_{r-1})=\alpha^{r}e_{0}, we obtain |e0|hC|α|r|er1|h|e_{0}|_{h}\leq C|\alpha|^{-r}|e_{r-1}|_{h}, and hence |e0|h|α|rC(r+1)/2|e_{0}|_{h}\leq|\alpha|^{-r}C^{(r+1)/2}. Then, we obtain |ei|h|α|rC(r+1)/2+i|e_{i}|_{h}\leq|\alpha|^{-r}C^{(r+1)/2+i}. ∎

Corollary 3.5.

We obtain

|α|r(C+1)2r|ei|h|α|r(C+1)2r.|\alpha|^{r}(C+1)^{-2r}\leq|e_{i}|_{h}\leq|\alpha|^{-r}(C+1)^{2r}.

Proof   We obtain the second inequality from Lemma 3.4. By using the duality, we obtain the first inequality.

Corollary 3.6.

For any hi(f,𝐞)h_{i}\in\mathcal{H}(f,{\boldsymbol{e}}) (i=1,2)(i=1,2) such that |f|hiC|f|_{h_{i}}\leq C, we obtain

|α|2r(C+1)4r|ei|h1|ei|h2|α|2r(C+1)4r.|\alpha|^{2r}(C+1)^{-4r}\leq\frac{|e_{i}|_{h_{1}}}{|e_{i}|_{h_{2}}}\leq|\alpha|^{-2r}(C+1)^{4r}.

3.3. Harmonic metrics of a cyclic Higgs bundle on a disc

We shall recall variants of Simpson’s main estimate for harmonic bundles, which was pioneered in [39], and further developed in [29] and [32].

We set U(R):={z||z|<R}U(R):=\bigl{\{}z\in{\mathbb{C}}\,\big{|}\,|z|<R\bigr{\}}. We set E:=i=0r1𝒪U(R)eiE:=\bigoplus_{i=0}^{r-1}\mathcal{O}_{U(R)}\,e_{i}. Let β\beta be a holomorphic function on U(R)U(R). Let ff be the endomorphism of EE determined by f(ei)=ei+1f(e_{i})=e_{i+1} (i=0,,r2)(i=0,\ldots,r-2) and f(er1)=βe0f(e_{r-1})=\beta e_{0}. We set θ:=fdz\theta:=f\,dz. We set Gr:={a|ar=1}G_{r}:=\{a\in{\mathbb{C}}\,|\,a^{r}=1\} as in §3.2. We define the GrG_{r}-action on EE by aei=aieia\bullet e_{i}=a^{i}e_{i}. Note that a(θ)=a1θa^{\ast}(\theta)=a^{-1}\theta. Let Harm(β)\mathop{\rm Harm}\nolimits(\beta) denote the set of GrG_{r}-invariant harmonic metrics of the Higgs bundle (E,θ)(E,\theta) such that i=0r1h(ei,ei)=1\prod_{i=0}^{r-1}h(e_{i},e_{i})=1.

Proposition 3.7 ([32, Proposition 2.1]).

Fix 0<R1<R0<R_{1}<R. Then, there exist C1,C2>0C_{1},C_{2}>0 depending only on rr, R1R_{1}, and RR such that

supU(R1)|f|hC1+C2maxU(R)|β|1/r\sup_{U(R_{1})}|f|_{h}\leq C_{1}+C_{2}\max_{U(R)}|\beta|^{1/r}

for any hHarm(β)h\in\mathop{\rm Harm}\nolimits(\beta). ∎

Corollary 3.8.

Let h0h_{0} be a GrG_{r}-invariant Hermitian metric of EE such that i=0r1h0(ei,ei)=1\prod_{i=0}^{r-1}h_{0}(e_{i},e_{i})=1. For hHarm(β)h\in\mathop{\rm Harm}\nolimits(\beta), we obtain the automorphism ss of EE which is self-adjoint with respect to both hh and h0h_{0}, determined by h=h0sh=h_{0}s. Let C1C_{1} and C2C_{2} be constants as in Proposition 3.7. Let C3>0C_{3}>0 be a positive constant such that |f|h0C3|f|_{h_{0}}\leq C_{3}. We set

A1:=max{C3,C1+C2maxzU(R)|β(z)|1/r}.A_{1}:=\max\Bigl{\{}C_{3},C_{1}+C_{2}\max_{z\in U(R)}|\beta(z)|^{1/r}\Bigr{\}}.

Then, the following holds at zU(R1)z\in U(R_{1}) such that β(z)0\beta(z)\neq 0:

|log|s|h0(z)|log(r|β(z)|2(A1+1)4r).\bigl{|}\log|s|_{h_{0}}(z)\bigr{|}\leq\log\Bigl{(}\sqrt{r}|\beta(z)|^{-2}(A_{1}+1)^{4r}\Bigr{)}.

Proof   It follows from Corollary 3.6 and Proposition 3.7. ∎

Suppose moreover that β\beta is nowhere vanishing on U(R)U(R). We fix an rr-th root α\alpha of β\beta, i.e., αr=β\alpha^{r}=\beta. We put τ:=exp(2π1/r)\tau:=\exp(2\pi\sqrt{-1}/r). We set vi:=j=0r1τijαjejv_{i}:=\sum_{j=0}^{r-1}\tau^{-ij}\alpha^{-j}e_{j} (i=0,1,,r1)(i=0,1,\ldots,r-1). We obtain the decomposition E=i=0r1𝒪U(R)viE=\bigoplus_{i=0}^{r-1}\mathcal{O}_{U(R)}v_{i} and f(vi)=τiαvif(v_{i})=\tau^{i}\alpha v_{i}.

Proposition 3.9 ([32, Corollary 2.6]).

Assume that |α(0)α(z)|1100|α(0)||\alpha(0)-\alpha(z)|\leq\frac{1}{100}|\alpha(0)| on U(R)U(R). Fix 0<R2<R10<R_{2}<R_{1}. Then, there exist C10>0C_{10}>0 and ϵ10>0\epsilon_{10}>0 depending only on rr, RR, R1R_{1} and R2R_{2} such that the following holds on U(R2)U(R_{2}) for any hHarm(β)h\in\mathop{\rm Harm}\nolimits(\beta) and for iji\neq j:

|h(vi,vj)|C10exp(ϵ10|α(0)|)|vi|h|vj|h.|h(v_{i},v_{j})|\leq C_{10}\cdot\exp\Bigl{(}-\epsilon_{10}|\alpha(0)|\Bigr{)}\cdot|v_{i}|_{h}\cdot|v_{j}|_{h}.

Note that |vi|h=|vj|h|v_{i}|_{h}=|v_{j}|_{h} by the GrG_{r}-invariance of hh. ∎

Let hcanh_{\mathop{\rm can}\nolimits} be the GrG_{r}-invariant Hermitian metric determined by

hcan(ej,ej)=|α|(r1)+2j.h_{\mathop{\rm can}\nolimits}(e_{j},e_{j})=|\alpha|^{-(r-1)+2j}.

Then, the frame (v0,,vr1)(v_{0},\ldots,v_{r-1}) is orthogonal, and we have

hcan(vi,vi)=r|α|(r1).h_{\mathop{\rm can}\nolimits}(v_{i},v_{i})=r|\alpha|^{-(r-1)}.

Hence, the curvature of the Chern connection of hcanh_{\mathop{\rm can}\nolimits} is 0. Because [θ,θhcan]=0[\theta,\theta^{\dagger}_{h_{\mathop{\rm can}\nolimits}}]=0, hcanh_{\mathop{\rm can}\nolimits} is a harmonic metric of the Higgs bundle (E,θ)(E,\theta).

Corollary 3.10.

Suppose that the assumption of Proposition 3.9 is satisfied. Moreover, we assume that

(r1)C10exp(ϵ10|α(0)|)1/2.(r-1)C_{10}\cdot\exp\Bigl{(}-\epsilon_{10}|\alpha(0)|\Bigr{)}\leq 1/2.

There exist ϵi>0\epsilon_{i}>0 and Ci>0C_{i}>0 (i=11,12,13)(i=11,12,13) depending only on rr, RR, R1R_{1} and R2R_{2} such that the following holds for any hHarm(β)h\in\mathop{\rm Harm}\nolimits(\beta) on U(R2)U(R_{2}):

|log(h(ej,ej)/hcan(ej,ej))|C11exp(ϵ11|α(0)|).\Bigl{|}\log\Bigl{(}h(e_{j},e_{j})/h_{\mathop{\rm can}\nolimits}(e_{j},e_{j})\Bigr{)}\Bigr{|}\leq C_{11}\exp\bigl{(}-\epsilon_{11}|\alpha(0)|\bigr{)}. (27)
|zlog(h(ej,ej)/hcan(ej,ej))|C12exp(ϵ12|α(0)|).\Bigl{|}\partial_{z}\log\Bigl{(}h(e_{j},e_{j})/h_{\mathop{\rm can}\nolimits}(e_{j},e_{j})\Bigr{)}\Bigr{|}\leq C_{12}\exp\bigl{(}-\epsilon_{12}|\alpha(0)|\bigr{)}. (28)
|z¯zlog(h(ej,ej)/hcan(ej,ej))|C13exp(ϵ13|α(0)|).\Bigl{|}\partial_{\overline{z}}\partial_{z}\log\Bigl{(}h(e_{j},e_{j})/h_{\mathop{\rm can}\nolimits}(e_{j},e_{j})\Bigr{)}\Bigr{|}\leq C_{13}\exp\bigl{(}-\epsilon_{13}|\alpha(0)|\bigr{)}. (29)

Proof   We obtain (27) from (26) and Proposition 3.9. Note that the Chern connection of the curvature of hcanh_{\mathop{\rm can}\nolimits} is 0, and that the decomposition E=𝒪eiE=\bigoplus\mathcal{O}e_{i} is orthogonal with respect to hh and hcanh_{\mathop{\rm can}\nolimits}. Then, we obtain (29) from Proposition 3.7 and Proposition 3.9. We obtain (28) from (27) and (29) by using the elliptic regularity. ∎

3.3.1. Estimates near boundary

Let YY be an open subset in U(R)U(R). Let Y¯\overline{Y} denote the closure of YY in U¯(R)\overline{U}(R). Let hHarm(β|Y)h\in\mathop{\rm Harm}\nolimits(\beta_{|Y}) such that hh extends to a continuous metric on Y¯\overline{Y}.

Proposition 3.11.

Fix 0<R10<R0<R_{10}<R. There exists CiC_{i} (i=1,2,3)(i=1,2,3), depending only on R10R_{10}, RR and rank(E)\mathop{\rm rank}\nolimits(E), such that the following holds on U(R10)YU(R_{10})\cap Y:

|f|hC1+C2maxzU(R)|β(z)|1/r+C3maxU(R)Y|f|h.|f|_{h}\leq C_{1}+C_{2}\max_{z\in U(R)}|\beta(z)|^{1/r}+C_{3}\max_{U(R)\cap\partial Y}|f|_{h}.

Proof   It follows from Proposition 3.13 below. ∎

Proposition 3.12.

Let h0h_{0} be a GrG_{r}-invariant Hermitian metric of EE such that i=0r1h0(ei,ei)=1\prod_{i=0}^{r-1}h_{0}(e_{i},e_{i})=1. For hHarm(β)h\in\mathop{\rm Harm}\nolimits(\beta), let s(h)s(h) be the automorphism of E|YE_{|Y} which is self-adjoint with respect to hh and h0h_{0}, determined by h=h0s(h)h=h_{0}\cdot s(h). Let CiC_{i} (i=1,2,3)(i=1,2,3) be as in Proposition 3.11, and let C4C_{4} be a positive constant such that |f|h0<C4|f|_{h_{0}}<C_{4}. We set

A1:=max{C4,C1+C2maxzU(R)|β(z)|1/r+C3maxU(R)Y|f|h}.A_{1}:=\max\Bigl{\{}C_{4},\,C_{1}+C_{2}\max_{z\in U(R)}|\beta(z)|^{1/r}+C_{3}\max_{U(R)\cap\partial Y}|f|_{h}\Bigr{\}}.

Then, the following holds at any zU(R10)Yz\in U(R_{10})\cap Y such that β(z)0\beta(z)\neq 0:

|log|s(h)|h0(z)|log(r|β(z)|2(A1+1)4r).\bigl{|}\log|s(h)|_{h_{0}}(z)\bigr{|}\leq\log\Bigl{(}\sqrt{r}|\beta(z)|^{-2}(A_{1}+1)^{4r}\Bigr{)}.

Proof   It follows from Proposition 3.11 and Corollary 3.6. ∎

3.3.2. Appendix: A variant of Simpson’s main estimate near boundary

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle defined on a neighbourhood of U¯(R)\overline{U}(R) in {\mathbb{C}}. Let YY be an open subset in U(R)U(R). Let Y¯\overline{Y} denote the closure of YY in U¯(R)\overline{U}(R). Let hh be a harmonic metric of (E,¯E,θ)|Y(E,\overline{\partial}_{E},\theta)_{|Y} which extends to a continuous metric of E|Y¯E_{|\overline{Y}}.

Let ff be the endomorphism of EE determined by θ=fdz\theta=f\,dz. Let MM be a positive constant such that the following holds.

  • Let α\alpha be any eigenvalue of f|Pf_{|P} for any PU¯(R)P\in\overline{U}(R). Then, |α|M|\alpha|\leq M.

Proposition 3.13.

Let R1<RR_{1}<R. There exist positive constant CiC_{i} (i=1,2,3)(i=1,2,3), depending only on RR, R1R_{1} and rank(E)\mathop{\rm rank}\nolimits(E), such that the following holds on YU(R1)Y\cap U(R_{1})

|f|hC1+C2M+C3maxU(R)Y|f|h.|f|_{h}\leq C_{1}+C_{2}M+C_{3}\max_{U(R)\cap\partial Y}|f|_{h}.

Proof   Recall the following inequality on YY:

zz¯log|f|h2|[f,fh]|2|f|h2.-\partial_{z}\partial_{\overline{z}}\log|f|^{2}_{h}\leq-\frac{\bigl{|}[f,f^{\dagger}_{h}]\bigr{|}^{2}}{|f|_{h}^{2}}.

For any PU¯(R)P\in\overline{U}(R), let α1,,αrankE\alpha_{1},\ldots,\alpha_{\mathop{\rm rank}\nolimits E} be the eigenvalues of f|Pf_{|P}, and we set g(P):=|αi|2g(P):=\sum|\alpha_{i}|^{2}. There exists C10>0C_{10}>0 such that the following holds for any PYP\in Y:

|[f|P,f|P]|hC10(|f|P|h2g(P)).|[f_{|P},f_{|P}^{\dagger}]|_{h}\geq C_{10}(|f_{|P}|_{h}^{2}-g(P)).

We obtain the following on YY:

zz¯log|f|h2(P)C102(|f|P|h2g(P))2|f|P|h2.-\partial_{z}\partial_{\overline{z}}\log|f|^{2}_{h}(P)\leq-C_{10}^{2}\frac{(|f_{|P}|_{h}^{2}-g(P))^{2}}{|f_{|P}|_{h}^{2}}.

If |f|P|22g(P)|f_{|P}|^{2}\geq 2g(P), then we obtain

zz¯log|f|h2(P)C1024|f|P|h2.-\partial_{z}\partial_{\overline{z}}\log|f|^{2}_{h}(P)\leq-\frac{C_{10}^{2}}{4}|f_{|P}|^{2}_{h}. (30)

We consider the following function on U(R)U(R):

A(z):=2R2(R2|z|2)22R2.A(z):=\frac{2R^{2}}{(R^{2}-|z|^{2})^{2}}\geq 2R^{-2}.

Note that zz¯logA=A-\partial_{z}\partial_{\overline{z}}\log A=-A. Let C11C_{11} be a constant such that C11>(C102/4)1C_{11}>(C_{10}^{2}/4)^{-1}. We set

C12:=C11+R2supU(R)Y|f|h2+10rank(E)R2M2.C_{12}:=C_{11}+R^{2}\sup_{U(R)\cap\partial Y}|f|_{h}^{2}+10\mathop{\rm rank}\nolimits(E)R^{2}M^{2}.

Note that the following holds on U(R)U(R):

zz¯log(C12A)=A=C121(C12A)C1024(C12A).-\partial_{z}\partial_{\overline{z}}\log(C_{12}A)=-A=-C_{12}^{-1}(C_{12}A)\geq-\frac{C_{10}^{2}}{4}(C_{12}A).

Let ZZ be the set of PYP\in Y such that |f|P|h2>C12A(P)|f_{|P}|_{h}^{2}>C_{12}A(P). We assume that ZZ is non-empty, and we shall derive a contradiction. Note that |f|h2<C12A|f|_{h}^{2}<C_{12}A on YU(R)\partial Y\cap U(R). We also note that |f|h|f|_{h} is bounded around any point of YU(R)\partial Y\cap\partial U(R), and that A(z)A(z)\to\infty as |z|R|z|\to R. Hence, ZZ is a relatively compact open subset of YY, which implies |f|h2=C12A|f|_{h}^{2}=C_{12}A on Z\partial Z. For PZP\in Z, we obtain |f|P|h22g(P)|f_{|P}|_{h}^{2}\geq 2g(P), and hence (30) holds at PP. We obtain the following at PZP\in Z:

zz¯(log|f|h2log(C12A))(P)C1024(|f|P|2C12A(P))<0.-\partial_{z}\partial_{\overline{z}}\Bigl{(}\log|f|_{h}^{2}-\log(C_{12}A)\Bigr{)}(P)\leq-\frac{C_{10}^{2}}{4}(|f_{|P}|^{2}-C_{12}A(P))<0. (31)

Together with |f|h2=C12A|f|_{h}^{2}=C_{12}A on Z\partial Z, we obtain |f|h2C12A|f|_{h}^{2}\leq C_{12}A on ZZ, which contradicts the construction of ZZ. Hence, we obtain that Z=Z=\emptyset. ∎

3.4. Some existence results of harmonic metrics on cyclic Higgs bundles

Let XX be a Riemann surface. Let E=i=0r1EiE=\bigoplus_{i=0}^{r-1}E_{i} be a graded holomorphic vector bundle on XX. Let θ\theta be a Higgs field of EE such that θ(Ei)Ei+1KX\theta(E_{i})\subset E_{i+1}\otimes K_{X} (i=0,,r2)(i=0,\ldots,r-2) and θ(Er1)E0KX\theta(E_{r-1})\subset E_{0}\otimes K_{X}. We assume that det(θ)\det(\theta) is not constantly 0. Assume that there exists a flat metric hdet(E)h_{\det(E)} of det(E)\det(E). We consider the GrG_{r}-action on EiE_{i} given by avi=aivia\bullet v_{i}=a^{i}v_{i}, which induces a GrG_{r}-action on EE. For any open subset YY, let HarmGr((E,θ)|Y,hdet(E))\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta)_{|Y},h_{\det(E)}) denote the set of GrG_{r}-invariant harmonic metrics hh of (E,θ)|Y(E,\theta)_{|Y} such that det(h)=hdet(E)|Y\det(h)=h_{\det(E)|Y}.

Remark 3.14.

For any rr-differential qq on XX, we obtain Harm(q)=HarmGr((𝕂X,r,θ(q)),1)\mathop{\rm Harm}\nolimits(q)=\mathop{\rm Harm}\nolimits^{G_{r}}\bigl{(}(\mathbb{K}_{X,r},\theta(q)),1\bigr{)} by definition.

3.4.1. Convergence

Let X1X2X_{1}\subset X_{2}\subset\cdots be a smooth exhaustive family of XX. Suppose that hiHarmGr((E,θ)|Xi,hdet(E))h_{i}\in\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta)_{|X_{i}},h_{\det(E)}). Let h0h_{0} be any GrG_{r}-invariant Hermitian metric of EE such that det(h0)=hdet(E)\det(h_{0})=h_{\det(E)}. Let sis_{i} be the automorphism of E|XiE_{|X_{i}} determined by hi=h0|Xisih_{i}=h_{0|X_{i}}s_{i}.

Proposition 3.15.

There exists a locally bounded function ff on XX such that |si|h0,i+|si1|h0,if|Xi|s_{i}|_{h_{0,i}}+|s_{i}^{-1}|_{h_{0,i}}\leq f_{|X_{i}} for any ii. As a result, there exists a convergent subsequence of the sequence {si}\{s_{i}\} with the limit ss_{\infty}, and we obtain h=h0sHarmGr((E,θ),hdet(E))h_{\infty}=h_{0}s_{\infty}\in\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta),h_{\det(E)}).

Proof   Let ZZ denote the zero set of det(θ)\det(\theta). It is discrete in XX. By Corollary 3.8, for any PXZP\in X\setminus Z, there exists a relatively compact neighbourhood XPX_{P} of PP in XZX\setminus Z with a constant CP>1C_{P}>1 such that CP1|si|h0CPC_{P}^{-1}\leq|s_{i}|_{h_{0}}\leq C_{P} for any large ii. By using the subharmonicity of Tr(si)\mathop{\rm Tr}\nolimits(s_{i}) (see §2.4), we also obtain a similar estimate locally around any point of ZZ. Then, the claim follows from Proposition 2.6. ∎

Corollary 3.16.

Suppose that there exists a subset YXY\subset X and an >0{\mathbb{R}}_{>0}-valued locally bounded function fYf_{Y} on YY such that |si|YXi|h0fY|YXi|s_{i|Y\cap X_{i}}|_{h_{0}}\leq f_{Y|Y\cap X_{i}}. Then, we obtain |s|Y|h0fY|s_{\infty|Y}|_{h_{0}}\leq f_{Y} for ss_{\infty} in Proposition 3.15. ∎

3.4.2. Control of growth order

Let gXg_{X} be any Kähler metric of XX. Let h0h_{0} be a GrG_{r}-invariant Hermitian metric of EE such that det(h0)=hdet(E)\det(h_{0})=h_{\det(E)}. Let {Xi}\{X_{i}\} be a smooth exhaustive family of XX. Let hiHarmGr((E,θ)|Xi,hdet(E))h_{i}\in\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta)_{|X_{i}},h_{\det(E)}) such that hi|Xi=h0|Xih_{i|\partial X_{i}}=h_{0|\partial X_{i}}.

Let YY be an open subset of XX. Suppose the following.

  • There exists an 0{\mathbb{R}}_{\geq 0}-valued function A0A_{0} on a neighbourhood NN of Y¯\overline{Y} such that the following holds on NN:

    1Λ¯A0|F(h0)|h0,gX.\sqrt{-1}\Lambda\overline{\partial}\partial A_{0}\geq|F(h_{0})|_{h_{0},g_{X}}.
  • There exists an 0{\mathbb{R}}_{\geq 0}-valued harmonic function A1A_{1} on NN such that logTr(hih01)A1+logr\log\mathop{\rm Tr}\nolimits(h_{i}\cdot h_{0}^{-1})\leq A_{1}+\log r on XiYX_{i}\cap\partial Y.

Proposition 3.17.

There exists hHarmGr((E,θ),hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta),h_{\det(E)}) such that logTr(hh01)A0+A1+logr\log\mathop{\rm Tr}\nolimits(h\cdot h_{0}^{-1})\leq A_{0}+A_{1}+\log r on YY.

Proof   By Proposition 3.15, we may assume that {hi}\{h_{i}\} is convergent to hh_{\infty}. Let sis_{i} be the automorphism of E|XiE_{|X_{i}} obtained as si=hih0|Xi1s_{i}=h_{i}\cdot h_{0|X_{i}}^{-1}. By our choice, we obtain

1Λ¯(logTr(si)A0A1logr)0.\sqrt{-1}\Lambda\overline{\partial}\partial\bigl{(}\log\mathop{\rm Tr}\nolimits(s_{i})-A_{0}-A_{1}-\log r\bigr{)}\leq 0.

On YXiY\cap\partial X_{i}, we obtain

logTr(si)A0A1logrlogTr(si)logr=0.\log\mathop{\rm Tr}\nolimits(s_{i})-A_{0}-A_{1}-\log r\leq\log\mathop{\rm Tr}\nolimits(s_{i})-\log r=0.

On XiYX_{i}\cap\partial Y, we obtain

logTr(si)A0A1logrlogTr(si)A1logr0.\log\mathop{\rm Tr}\nolimits(s_{i})-A_{0}-A_{1}-\log r\leq\log\mathop{\rm Tr}\nolimits(s_{i})-A_{1}-\log r\leq 0.

Hence, we obtain logTr(si)A0A1logr0\log\mathop{\rm Tr}\nolimits(s_{i})-A_{0}-A_{1}-\log r\leq 0 on YXiY\cap X_{i}. We obtain logTr(hh01)A0A1logr0\log\mathop{\rm Tr}\nolimits(h_{\infty}h_{0}^{-1})-A_{0}-A_{1}-\log r\leq 0 on YY as in Corollary 3.16. ∎

3.4.3. Compact subsets

Proposition 3.18.

Let KXK\subset X be a compact subset. For any h1HarmGr((E,θ)|XK,hdet(E))h_{1}\in\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta)_{|X\setminus K},h_{\det(E)}), there exists hHarmGr((E,θ),hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta),h_{\det(E)}) such that h|XNh_{|X\setminus N} and h1|XNh_{1|X\setminus N} are mutually bounded for any relatively compact neighbourhood NN of KK in XX.

Proof   Let h2h_{2} be a GrG_{r}-invariant Hermitian metric of EE such that det(h2)=hdet(E)\det(h_{2})=h_{\det(E)} and that h2|XN=h1|XNh_{2|X\setminus N}=h_{1|X\setminus N} for a relatively compact neighbourhood NN of KK. Note that (E,θ,h2)(E,\theta,h_{2}) is analytically stable with respect to the GrG_{r}-action in the sense of §2.3 because there is no proper GrG_{r}-invariant Higgs subbundle by Lemma 3.2. (Recall that det(θ)\det(\theta) is not constantly 0.) Then, the claim follows from Proposition 2.11. ∎

3.5. Asymptotic behaviour around poles

Let UU be a neighbourhood of 0 in {\mathbb{C}}. We set U:=U{0}U^{\circ}:=U\setminus\{0\}. Let qq be a nowhere vanishing holomorphic rr-differential on UU^{\circ} which is meromorphic at 0. We have the expression q=zmα(dz/z)rq=z^{m}\,\alpha\,(dz/z)^{r}, where α\alpha induces a nowhere vanishing holomorphic function on UU.

3.5.1. The case m0m\leq 0

For a positive number ϵ>0\epsilon>0, we set

ρϵ(z):={exp(ϵ|z|m/r)(m<0)|z|ϵ(m=0).\rho_{\epsilon}(z):=\left\{\begin{array}[]{ll}\exp(-\epsilon|z|^{m/r})&(m<0)\\ |z|^{\epsilon}&(m=0).\end{array}\right.

There exists a relatively compact neighbourhood U1U_{1} of 0 in UU such that qq is nowhere vanishing on U1:=U1UU_{1}^{\circ}:=U_{1}\cap U^{\circ}. Recall that there exists the canonical harmonic metric hcanHarm(q|U1)h_{\mathop{\rm can}\nolimits}\in\mathop{\rm Harm}\nolimits(q_{|U_{1}^{\circ}}).

Proposition 3.19.

Any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) is mutually bounded with hcanh_{\mathop{\rm can}\nolimits} on U1U_{1}^{\circ}. More strongly, there exist C>0C>0 and ϵ>0\epsilon>0 such that the following holds on U1U_{1}^{\circ}:

|log|(dz)(r+12i)/2|hlog|(dz)(r+12i)/2|hcan|Cρϵ.\left|\log|(dz)^{(r+1-2i)/2}|_{h}-\log|(dz)^{(r+1-2i)/2}|_{h_{\mathop{\rm can}\nolimits}}\right|\leq C\rho_{\epsilon}.

Proof   We may assume that {|z|1}U1\{|z|\leq 1\}\subset U_{1}. Let Ψ:\Psi:{\mathbb{C}}\longrightarrow{\mathbb{C}}^{\ast} be the map given by Ψ(ζ)=e1ζ\Psi(\zeta)=e^{\sqrt{-1}\zeta}. We have the expression

Ψ(q)=e1mζΨ(α)(1dζ)r.\Psi^{\ast}(q)=e^{\sqrt{-1}m\zeta}\Psi^{\ast}(\alpha)(\sqrt{-1}d\zeta)^{r}.

There exists C1>1C_{1}>1 such that C11|Ψ(α)|C1C_{1}^{-1}\leq|\Psi^{\ast}(\alpha)|\leq C_{1} on {Im(ζ)0}\{\mathop{\rm Im}\nolimits(\zeta)\geq 0\}.

Let us consider the case m=0m=0. For ζ0\zeta_{0} with y0=Im(ζ0)>0y_{0}=\mathop{\rm Im}\nolimits(\zeta_{0})>0, let Fζ0:{|w|<1}{Im(ζ)>0}F_{\zeta_{0}}:\{|w|<1\}\longrightarrow\{\mathop{\rm Im}\nolimits(\zeta)>0\} given by Fζ0(w)=y0(w+1)F_{\zeta_{0}}(w)=y_{0}(w+\sqrt{-1}). We set h~:=(ΨFζ0)(h)\widetilde{h}:=(\Psi\circ F_{\zeta_{0}})^{\ast}(h) and h~can:=(ΨFζ0)(hcan)\widetilde{h}_{\mathop{\rm can}\nolimits}:=(\Psi\circ F_{\zeta_{0}})^{\ast}(h_{\mathop{\rm can}\nolimits}). By Proposition 3.7 and Corollary 3.10, there exists C2>0C_{2}>0 and ϵ2>0\epsilon_{2}>0 such that

|log|(dw)(r+12i)/2|h~log|(dw)(r+12i)/2|h~can|C2exp(ϵ2y0).\Bigl{|}\log|(dw)^{(r+1-2i)/2}|_{\widetilde{h}}-\log|(dw)^{(r+1-2i)/2}|_{\widetilde{h}_{\mathop{\rm can}\nolimits}}\Bigr{|}\leq C_{2}\exp(-\epsilon_{2}y_{0}).

It implies the desired estimate in the case m=0m=0.

Let us study the case m<0m<0. Take a large T>0T>0. For ζ0\zeta_{0} with y0=Im(ζ0)>0y_{0}=\mathop{\rm Im}\nolimits(\zeta_{0})>0, we consider Fζ0:{|w|<1}{Im(ζ)>0}F_{\zeta_{0}}:\{|w|<1\}\longrightarrow\{\mathop{\rm Im}\nolimits(\zeta)>0\} determined by Fζ0(w)=T(w+ζ0)F_{\zeta_{0}}(w)=T(w+\zeta_{0}). By applying Proposition 3.7 and Corollary 3.10 to the pull back by ΨFζ0\Psi\circ F_{\zeta_{0}}, we obtain the desired estimate in the case m<0m<0. ∎

Remark 3.20.

Proposition 3.19 can be also obtained as a consequence of the classification of GrG_{r}-equivariant good filtered Higgs bundles over a cyclic Higgs bundle on a punctured disc. See [30, §3.1.3].

3.5.2. The case m>0m>0

Let 𝒫r,m\mathcal{P}_{r,m} denote the set of 𝒃=(bi)r{\boldsymbol{b}}=(b_{i})\in{\mathbb{R}}^{r} satisfying

b1b2brb1m,bi=r(r+1)2.b_{1}\geq b_{2}\geq\cdots\geq b_{r}\geq b_{1}-m,\quad\sum b_{i}=-\frac{r(r+1)}{2}.

Let 𝒫r,m𝒫r,m\mathcal{P}^{{\mathbb{R}}}_{r,m}\subset\mathcal{P}_{r,m} denote the subset of 𝒃𝒫r,m{\boldsymbol{b}}\in\mathcal{P}_{r,m} such that bi+br+1i=(r+1)b_{i}+b_{r+1-i}=-(r+1). We obtain the following proposition from [39] (see also [30]).

Proposition 3.21.

  • For any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), there exist 𝒃(h)=(bi(h))𝒫r,m{\boldsymbol{b}}(h)=(b_{i}(h))\in\mathcal{P}_{r,m} determined by the following condition around z=0z=0:

    log|(dz)(r+12i)/2|h+(bi(h)+i)log|z|=O(log(log|z|))(i=1,,r).\log|(dz)^{(r+1-2i)/2}|_{h}+(b_{i}(h)+i)\log|z|\\ =O\Bigl{(}\log\bigl{(}-\log|z|\bigr{)}\Bigr{)}\quad(i=1,\ldots,r). (32)

    If hHarm(q)h\in\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q), then 𝒃(h)𝒫r,m{\boldsymbol{b}}(h)\in\mathcal{P}^{{\mathbb{R}}}_{r,m}.

  • Let h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q) such that 𝒃(h1)=𝒃(h2){\boldsymbol{b}}(h_{1})={\boldsymbol{b}}(h_{2}). Then, for any relatively compact neighbourhood UU^{\prime} of 0 in UU, the restrictions h1|U{0}h_{1|U^{\prime}\setminus\{0\}} and h2|U{0}h_{2|U^{\prime}\setminus\{0\}} are mutually bounded.

Proof   Let us explain an outline of the proof (see also [30, §3.1.2, §3.2.8]). Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), i.e., hh is a GrG_{r}-invariant harmonic metric of (𝕂U,r,θ(q))(\mathbb{K}_{U^{\circ},r},\theta(q)) such that det(h)=1\det(h)=1. Let ff be the endomorphism of 𝕂U,r\mathbb{K}_{U^{\circ},r} determined by θ(q)=fdz/z\theta(q)=f\,dz/z. Because m>0m>0, the harmonic bundle (𝕂U,r,θ(q),h)(\mathbb{K}_{U^{\circ},r},\theta(q),h) is tame. For any aa\in{\mathbb{R}}, and for any open subset 𝒰0\mathcal{U}\ni 0, let 𝒫ah𝕂U,r(𝒰)\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r}(\mathcal{U}) denote the space of sections ss of 𝒰{0}\mathcal{U}\setminus\{0\} such that |s|h=O(|z|aϵ)|s|_{h}=O\bigl{(}|z|^{-a-\epsilon}\bigr{)} for any ϵ>0\epsilon>0. For any open subset 𝒰∌0\mathcal{U}\not\ni 0, let 𝒫ah𝕂U,r(𝒰)\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r}(\mathcal{U}) denote the space of holomorphic sections of 𝕂U,r(𝒰)\mathbb{K}_{U^{\circ},r}(\mathcal{U}). Thus, we obtain 𝒪U\mathcal{O}_{U}-modules 𝒫ah𝕂U,r\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r} (a)(a\in{\mathbb{R}}). According to [39, Theorem 2], 𝒫ah𝕂U,r\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r} are locally free 𝒪U\mathcal{O}_{U}-modules, and f(𝒫ah𝕂U,r)𝒫ah𝕂U,rf(\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r})\subset\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r} for any aa. Because hh is GrG_{r}-invariant, 𝒫ah𝕂U,r\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r} is naturally GrG_{r}-equivariant. Hence, we obtain the decomposition

𝒫ah𝕂U,r=i=1r𝒫ahKU(r+12i)/2.\mathcal{P}^{h}_{a}\mathbb{K}_{U^{\circ},r}=\bigoplus_{i=1}^{r}\mathcal{P}^{h}_{a}K_{U^{\circ}}^{(r+1-2i)/2}. (33)

Here, the locally free 𝒪U\mathcal{O}_{U}-modules 𝒫ahKU(r+12i)/2\mathcal{P}^{h}_{a}K_{U^{\circ}}^{(r+1-2i)/2} are obtained from KU(r+12i)/2K_{U^{\circ}}^{(r+1-2i)/2} with hh as above.

Because m>0m>0, the eigenvalues of f|zf_{|z} goes to 0 as |z|0|z|\to 0. Hence, according to Simpson’s main estimate [39, Theorem 1], we obtain |f|h=O((log|z|)1)|f|_{h}=O\bigl{(}(-\log|z|)^{-1}\bigr{)}.

We set vi:=zi(dz)(r+12i)/2v_{i}:=z^{i}(dz)^{(r+1-2i)/2}. Then, we obtain f(vi)=vi+1f(v_{i})=v_{i+1} (i=1,,r1)(i=1,\ldots,r-1), and hence |vi+1|h|vi|h(log|z|)1|v_{i+1}|_{h}\leq|v_{i}|_{h}(-\log|z|)^{-1}. Because det(h)=1\det(h)=1, we obtain

|vr|r(log|z|)r(r1)/2i=1h|vi|h=|z|r(r+1)/2.|v_{r}|^{r}\cdot(-\log|z|)^{r(r-1)/2}\leq\prod_{i=1}^{h}|v_{i}|_{h}=|z|^{r(r+1)/2}.

Hence, there exists aa\in{\mathbb{R}} such that vrv_{r} is a section of 𝒫(r+1)/2h𝕂U,r\mathcal{P}^{h}_{(r+1)/2}\mathbb{K}_{U^{\circ},r}. Because f(vr)=zmαv1f(v_{r})=z^{m}\alpha v_{1}, we obtain |z|m|α||v1|h|vr|h(log|z|)1|z|^{m}\cdot|\alpha|\cdot|v_{1}|_{h}\leq|v_{r}|_{h}(-\log|z|)^{-1}. Therefore, there exists bb\in{\mathbb{R}} such that vi𝒫bh𝕂U,rv_{i}\in\mathcal{P}^{h}_{b}\mathbb{K}_{U^{\circ},r} (i=1,,r)(i=1,\ldots,r). We set

bi(h):=min{b|vi𝒫bh𝕂U,r}.b_{i}(h):=\min\bigl{\{}b\in{\mathbb{R}}\,\big{|}\,v_{i}\in\mathcal{P}^{h}_{b}\mathbb{K}_{U^{\circ},r}\bigr{\}}.

Then, by the norm estimate [39, §7] (Proposition 3.22 below), we obtain (32). Because f(vi)=vi+1f(v_{i})=v_{i+1} and f(vr)=zmαv1f(v_{r})=z^{m}\alpha v_{1}, we obtain bibi+1b_{i}\geq b_{i+1} and brb1mb_{r}\geq b_{1}-m. We also remark the filtered bundle 𝒫h𝕂U,r\mathcal{P}^{h}_{\ast}\mathbb{K}_{U^{\circ},r} is uniquely determined by the numbers 𝒃(h)=(bi(h)){\boldsymbol{b}}(h)=(b_{i}(h)) because of the decomposition (33).

Let hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) such that 𝒃(h1)=𝒃(h2){\boldsymbol{b}}(h_{1})={\boldsymbol{b}}(h_{2}). We obtain 𝒫h1𝕂U,r=𝒫h2𝕂U,r\mathcal{P}^{h_{1}}_{\ast}\mathbb{K}_{U^{\circ},r}=\mathcal{P}^{h_{2}}_{\ast}\mathbb{K}_{U^{\circ},r}. Hence, according to [39, Corollary 4.3], h1h_{1} and h2h_{2} are mutually bounded. ∎

Let us refine the estimate (32). For 𝒃𝒫r,m{\boldsymbol{b}}\in\mathcal{P}_{r,m}, we introduce a non-negative integer \ell and integers ν0\nu_{0}, ν1,,ν,ν+1\nu_{1},\ldots,\nu_{\ell},\nu_{\ell+1} as follows. If b1=brb_{1}=b_{r}, we set =0\ell=0, ν0=0\nu_{0}=0 and ν1=r\nu_{1}=r. If b1>brb_{1}>b_{r}, we obtain the numbers 1\ell\in{\mathbb{Z}}_{\geq 1} and 1ν1<<ν<r1\leq\nu_{1}<\cdots<\nu_{\ell}<r determined by the following condition.

  • b1=bν1b_{1}=b_{\nu_{1}}, bνj>bνj+1=bνj+1b_{\nu_{j}}>b_{\nu_{j}+1}=b_{\nu_{j+1}} (j=1,,1)(j=1,\ldots,\ell-1), and bν>bν+1=brb_{\nu_{\ell}}>b_{\nu_{\ell}+1}=b_{r}.

Moreover, we set ν0=0\nu_{0}=0 and ν+1=r\nu_{\ell+1}=r. For 1ir1\leq i\leq r, we obtain the number j(i)j(i) determined by νj(i)<iνj(i)+1\nu_{j(i)}<i\leq\nu_{j(i)+1}. If j(i)0,j(i)\neq 0,\ell, we set

ki(𝒃):=νj(i)+1νj(i)+12(iνj(i))=νj(i)+1+νj(i)+12i.k_{i}({\boldsymbol{b}}):=\nu_{j(i)+1}-\nu_{j(i)}+1-2(i-\nu_{j(i)})=\nu_{j(i)+1}+\nu_{j(i)}+1-2i. (34)

If b1mbrb_{1}-m\neq b_{r}, we define ki(𝒃)k_{i}({\boldsymbol{b}}) for ii such that j(i)=0,j(i)=0,\ell by the same formula (34). If b1m=brb_{1}-m=b_{r}, we define ki(𝒃)k_{i}({\boldsymbol{b}}) for ii such that j(i)=0j(i)=0 or j(i)=j(i)=\ell as follows:

ki(𝒃):={ν1+νr+12i(j(i)=0)ν1+ν+r+12i(j(i)=).k_{i}({\boldsymbol{b}}):=\left\{\begin{array}[]{ll}\nu_{1}+\nu_{\ell}-r+1-2i&(j(i)=0)\\ \nu_{1}+\nu_{\ell}+r+1-2i&(j(i)=\ell).\end{array}\right. (35)

Thus, we obtain 𝒌(𝒃)r{\boldsymbol{k}}({\boldsymbol{b}})\in{\mathbb{Z}}^{r}. The following is a consequence of the norm estimate of Simpson in [39].

Proposition 3.22.

For hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), we set 𝐤(h):=𝐤(𝐛(h)){\boldsymbol{k}}(h):={\boldsymbol{k}}({\boldsymbol{b}}(h)), where 𝐛(h)𝒫r,m{\boldsymbol{b}}(h)\in\mathcal{P}_{r,m} is given as in Proposition 3.21. Then, we obtain the following refined estimates for i=1,,ri=1,\ldots,r:

log|(dz)(r+12i)/2|h+(bi(h)+i)log|z|ki(h)2log(log|z|)=O(1).\log|(dz)^{(r+1-2i)/2}|_{h}+(b_{i}(h)+i)\log|z|-\frac{k_{i}(h)}{2}\log\bigl{(}-\log|z|\bigr{)}=O(1). (36)

Proof   We explain an outline of the proof (see also [30]). We use the notation in the proof of Proposition 3.21. We set 𝒫<ah(𝕂U,r)\mathcal{P}^{h}_{<a}(\mathbb{K}_{U^{\circ},r}). For 1<a0-1<a\leq 0, we obtain the following finite dimensional complex vector space

GraF(𝒫0h𝕂U,r|0):=𝒫ah(𝕂U,r)/𝒫<ah(𝕂U,r).\mathop{\rm Gr}\nolimits^{F}_{a}(\mathcal{P}^{h}_{0}\mathbb{K}_{U^{\circ},r|0}):=\mathcal{P}^{h}_{a}(\mathbb{K}_{U^{\circ},r})\big{/}\mathcal{P}^{h}_{<a}(\mathbb{K}_{U^{\circ},r}).

Let WW denote the filtration on GraF(𝒫0h𝕂U,r|0)\mathop{\rm Gr}\nolimits^{F}_{a}(\mathcal{P}^{h}_{0}\mathbb{K}_{U^{\circ},r|0}) obtained as the monodromy weight filtration of the nilpotent endomorphism Res(θ)\mathop{\rm Res}\nolimits(\theta).

Let ni(h)n_{i}(h) denote the integer determined by 1<bi(h)+ni(h)0-1<b_{i}(h)+n_{i}(h)\leq 0. We set ci(h):=bi(h)+ni(h)c_{i}(h):=b_{i}(h)+n_{i}(h). Then, v~i:=zni(h)vi\widetilde{v}_{i}:=z^{-n_{i}(h)}v_{i} is a section of 𝒫ci(h)h(𝕂U,r)\mathcal{P}^{h}_{c_{i}(h)}(\mathbb{K}_{U^{\circ},r}). Let [v~i][\widetilde{v}_{i}] denote the induced element of Grci(h)F(𝒫0h𝕂U,r|0)\mathop{\rm Gr}\nolimits^{F}_{c_{i}(h)}(\mathcal{P}^{h}_{0}\mathbb{K}_{U^{\circ},r|0}). Then, the tuple ([v~i]|ci(h)=a)([\widetilde{v}_{i}]\,|\,c_{i}(h)=a) is a base of GraF(𝒫0𝕂U,r|0)\mathop{\rm Gr}\nolimits^{F}_{a}(\mathcal{P}_{0}\mathbb{K}_{U^{\circ},r|0}).

We obtain the following for any i=1,,r1i=1,\ldots,r-1:

Res(θ)[v~i]={[v~i+1](bi(h)=bi+1(h))0(bi(h)>bi+1(h))\mathop{\rm Res}\nolimits(\theta)[\widetilde{v}_{i}]=\left\{\begin{array}[]{ll}[\widetilde{v}_{i+1}]&(b_{i}(h)=b_{i+1}(h))\\ 0&(b_{i}(h)>b_{i+1}(h))\end{array}\right. (37)

We also obtain the following:

Res(θ)[v~r]={α(0)[v~1](br(h)=b1(h)m)0(otherwise)\mathop{\rm Res}\nolimits(\theta)[\widetilde{v}_{r}]=\left\{\begin{array}[]{ll}\alpha(0)[\widetilde{v}_{1}]&(b_{r}(h)=b_{1}(h)-m)\\ 0&\mbox{\rm(otherwise)}\end{array}\right.

Hence, we obtain

WkGraF(𝒫0h𝕂U,r|0)=ci(h)=aki(h)k[v~i].W_{k}\mathop{\rm Gr}\nolimits^{F}_{a}(\mathcal{P}^{h}_{0}\mathbb{K}_{U^{\circ},r|0})=\bigoplus_{\begin{subarray}{c}c_{i}(h)=a\\ k_{i}(h)\leq k\end{subarray}}{\mathbb{C}}[\widetilde{v}_{i}]. (38)

According to the norm estimate for tame harmonic bundles [39, §7], there exists C>1C>1 such that

C1|z|ci(h)(log|z|)ki(h)/2|v~i|hC|z|ci(h)(log|z|)ki(h)/2.C^{-1}|z|^{-c_{i}(h)}(-\log|z|)^{k_{i}(h)/2}\leq|\widetilde{v}_{i}|_{h}\leq C|z|^{-c_{i}(h)}(-\log|z|)^{k_{i}(h)/2}.

Thus, we obtain (36). ∎

We also remark the following existence.

Proposition 3.23.

For any 𝐛𝒫r,m{\boldsymbol{b}}\in\mathcal{P}_{r,m}, there exist a neighbourhood U1U_{1} of 0 in UU and hHarm(q|U1U)h\in\mathop{\rm Harm}\nolimits(q_{|U_{1}\cap U^{\circ}}) such that 𝐛(h)=𝐛{\boldsymbol{b}}(h)={\boldsymbol{b}}.

Proof   We may assume that qq is nowhere vanishing on UU^{\circ}.

Lemma 3.24.

There exist a neighbourhood U1U_{1} of 0 and a holomorphic function ζ\zeta on U1U_{1} such that zζ(0)=1\partial_{z}\zeta(0)=1 and that q=α(0)ζm(dζ/ζ)rq=\alpha(0)\zeta^{m}(d\zeta/\zeta)^{r}. The germ of ζ\zeta at 0 is unique.

Proof   We set β:=α(0)1α\beta:=\alpha(0)^{-1}\alpha which is nowhere vanishing on UU. We have the Taylor expansion β=z+j=1βjzj\beta=z+\sum_{j=1}^{\infty}\beta_{j}z^{j}. First, let us prove that there exists a unique formal power series ζ(f)=1+j=2ζj(f)zj\zeta^{(f)}=1+\sum_{j=2}^{\infty}\zeta^{(f)}_{j}z^{j} satisfying β=(ζ(f)/z)m(zζ(f))r\beta=(\zeta^{(f)}/z)^{m}(\partial_{z}\zeta^{(f)})^{r}. The coefficient of zjz^{j} (j>1)(j>1) in (ζ(f)/z)m(zζ(f))r(\zeta^{(f)}/z)^{m}(\partial_{z}\zeta^{(f)})^{r} is described as the sum of (m+r(j+1))ζj+1(f)(m+r(j+1))\zeta^{(f)}_{j+1} and a polynomial of ζi(f)\zeta^{(f)}_{i} (i=2,,j)(i=2,\ldots,j). Hence, ζj+1(f)\zeta^{(f)}_{j+1} are uniquely determined by an easy induction. It particularly implies that the germ of ζ\zeta is uniquely determined. Moreover, if there exists n>0n\in{\mathbb{Z}}_{>0} such that βj=0\beta_{j}=0 unless jn0j\in n{\mathbb{Z}}_{\geq 0}, then we obtain ζj+1(f)=0\zeta^{(f)}_{j+1}=0 unless jn0j\in n{\mathbb{Z}}_{\geq 0}.

Let us prove the existence of a convergent solution in the case where m1:=m/r>0m_{1}:=m/r\in{\mathbb{Z}}_{>0}. Let β1/r\beta^{1/r} be the holomorphic function on UU determined by the conditions (β1/r)r=β(\beta^{1/r})^{r}=\beta and β1/r(0)=1\beta^{1/r}(0)=1. For any zUz\in U, we take a path γ(z)\gamma(z) connecting 0 and zz in UU, and we set

F(z)=γ(z)zm11β1/r𝑑z.F(z)=\int_{\gamma(z)}z^{m_{1}-1}\beta^{1/r}dz.

Then, F(z)F(z) is a holomorphic function on UU satisfying

dF(z)=zm11β1/rdz,F(0)=0.dF(z)=z^{m_{1}-1}\beta^{1/r}dz,\quad F(0)=0.

Note that F(z)/zm1F(z)/z^{m_{1}} is holomorphic at z=0z=0, and (F(z)/zm1)(0)=m11(F(z)/z^{m_{1}})(0)=m_{1}^{-1}. There exist a neighbourhood U1U_{1} of 0 in UU and a holomorphic function ζ\zeta on U1U_{1} such that F(z)=1m1ζm1F(z)=\frac{1}{m_{1}}\zeta^{m_{1}} and zζ(0)=1\partial_{z}\zeta(0)=1. Then, we obtain ζm11dζ=zm11β1/rdz\zeta^{m_{1}-1}d\zeta=z^{m_{1}-1}\beta^{1/r}\,dz, which implies ζmα(0)(dζ/ζ)r=zmα(dz/z)r\zeta^{m}\alpha(0)(d\zeta/\zeta)^{r}=z^{m}\alpha\cdot(dz/z)^{r}.

Let us study the existence in the general case. Let φr:\varphi_{r}:{\mathbb{C}}\longrightarrow{\mathbb{C}} be determined by φr(w)=wr\varphi_{r}(w)=w^{r}. We set U~:=φr1(U)\widetilde{U}:=\varphi_{r}^{-1}(U). There exists a neighbourhood U1U_{1} of 0 in UU and a holomorphic function ζ~\widetilde{\zeta} on φr1(U1)\varphi_{r}^{-1}(U_{1}) such that

φr(q)=rrα(0)ζ~mr(dζ~/ζ~)r=α(0)(ζ~r)m(dζ~r/ζ~r)r.\varphi_{r}^{\ast}(q)=r^{r}\alpha(0)\widetilde{\zeta}^{mr}(d\widetilde{\zeta}/\widetilde{\zeta})^{r}=\alpha(0)(\widetilde{\zeta}^{r})^{m}\cdot(d\widetilde{\zeta}^{r}/\widetilde{\zeta}^{r})^{r}.

By the consideration in the first paragraph of this proof, we obtain that ζ~r\widetilde{\zeta}^{r} is a convergent power series of wrw^{r}. Hence, there exists a holomorphic function ζ\zeta on U1U_{1} such that φ(ζ)=ζ~r\varphi^{\ast}(\zeta)=\widetilde{\zeta}^{r}, for which q=ζmα(0)(dζ/ζ)rq=\zeta^{m}\alpha(0)\cdot(d\zeta/\zeta)^{r} holds.

We obtain an embedding ζ:U11\zeta:U_{1}\longrightarrow{\mathbb{C}}\subset\mathbb{P}^{1}. Let ww be the standard coordinate on {\mathbb{C}}. We set q0=α(0)wm(dw/w)rq_{0}=\alpha(0)w^{m}(dw/w)^{r}. We obtain ζ(q0)=q\zeta^{\ast}(q_{0})=q. According to [30], there exists h1Harm(q0)h_{1}\in\mathop{\rm Harm}\nolimits(q_{0}) such that 𝒃(h1)=𝒃{\boldsymbol{b}}(h_{1})={\boldsymbol{b}}. Then, we obtain h=ζ(h1)Harm(q)h=\zeta^{\ast}(h_{1})\in\mathop{\rm Harm}\nolimits(q) which satisfies 𝒃(h)=𝒃{\boldsymbol{b}}(h)={\boldsymbol{b}}. ∎

3.6. Examples on {\mathbb{C}}

We introduce some examples as a preparation for the proof of Theorem 5.3. We shall see more examples later in §6.3.

3.6.1. Preliminary

Let 𝒫\mathcal{P} denote the set of 𝒂=(a1,,ar)r{\boldsymbol{a}}=(a_{1},\ldots,a_{r})\in{\mathbb{R}}^{r} satisfying

a1a2ara11,ai=0.a_{1}\geq a_{2}\geq\cdots\geq a_{r}\geq a_{1}-1,\quad\sum a_{i}=0.

For any 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}, we introduce a non-negative integer \ell and integers ν0,ν1,,ν,ν+1\nu_{0},\nu_{1},\ldots,\nu_{\ell},\nu_{\ell+1} as follows. If a1=ara_{1}=a_{r}, we set =0\ell=0, ν0=0\nu_{0}=0 and ν1=r\nu_{1}=r. If a1>ara_{1}>a_{r}, we obtain positive integers \ell and 1ν1<<ν<r1\leq\nu_{1}<\cdots<\nu_{\ell}<r by the following condition.

  • aν1=a1a_{\nu_{1}}=a_{1}, aνj>aνj+1a_{\nu_{j}}>a_{\nu_{j}+1} (j=1,,1)(j=1,\ldots,\ell-1) and aν>aν+1=ara_{\nu_{\ell}}>a_{\nu_{\ell}+1}=a_{r}.

Moreover, we set ν0=0\nu_{0}=0 and ν+1=r\nu_{\ell+1}=r. For 1ir1\leq i\leq r, we obtain the number j(i)j(i) determined by νj(i)<iνj(i)+1\nu_{j(i)}<i\leq\nu_{j(i)+1}. If j(i)0,j(i)\neq 0,\ell, we set

ki(𝒂):=νj(i)+1νj(i)+12(iνj(i))=νj(i)+1+νj(i)+12i.k_{i}({\boldsymbol{a}}):=\nu_{j(i)+1}-\nu_{j(i)}+1-2(i-\nu_{j(i)})=\nu_{j(i)+1}+\nu_{j(i)}+1-2i. (39)

If a11ara_{1}-1\neq a_{r}, we define ki(𝒂)k_{i}({\boldsymbol{a}}) for ii such that j(i)=0,j(i)=0,\ell by the same formula (39). If a11=ara_{1}-1=a_{r}, we define ki(𝒂)k_{i}({\boldsymbol{a}}) for ii such that j(i)=0j(i)=0 or j(i)=j(i)=\ell by

ki(𝒂):={ν1+νr+12i(j(i)=0)ν1+ν+r+12i(j(i)=).k_{i}({\boldsymbol{a}}):=\left\{\begin{array}[]{ll}\nu_{1}+\nu_{\ell}-r+1-2i&(j(i)=0)\\ \nu_{1}+\nu_{\ell}+r+1-2i&(j(i)=\ell).\end{array}\right. (40)

3.6.2. Examples

Let zz be the standard coordinate of {\mathbb{C}}. Let (x,y)(x,y) be the real coordinate system obtained as z=x+1yz=x+\sqrt{-1}y. We set q:=αe1z(dz)rq:=\alpha e^{\sqrt{-1}z}\,(dz)^{r}, where α\alpha is a non-zero complex number.

Proposition 3.25.

For any 𝐚𝒫{\boldsymbol{a}}\in\mathcal{P}, there exists h𝐚Harm(q)h_{{\boldsymbol{a}}}\in\mathop{\rm Harm}\nolimits(q) such that the following estimate holds on {y0}\{y\geq 0\}:

log|(dz)(r+12i)/2|h𝒂aiyki(𝒂)2log(y+2)=O(1).\log\bigl{|}(dz)^{(r+1-2i)/2}\bigr{|}_{h_{{\boldsymbol{a}}}}-a_{i}y-\frac{k_{i}({\boldsymbol{a}})}{2}\log(y+2)=O(1).

Proof   We apply the Kobayashi-Hitchin correspondence on {\mathbb{C}}^{\ast} to construct such hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) in [30]. Let φ:\varphi:{\mathbb{C}}\longrightarrow{\mathbb{C}}^{\ast} be determined by φ(z)=1re1z/r\varphi(z)=-\sqrt{-1}re^{\sqrt{-1}z/r}. We obtain φ((dw)r)=e1z(dz)r\varphi^{\ast}((dw)^{r})=e^{\sqrt{-1}z}(dz)^{r}.

We consider the rr-differential q1:=α(dw)rq_{1}:=\alpha(dw)^{r} on {\mathbb{C}}^{\ast}. For any 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}, we define 𝒃𝒫r,r{\boldsymbol{b}}\in\mathcal{P}_{r,r} by the following relation:

ai=1r(bi+r+12).a_{i}=\frac{1}{r}\left(b_{i}+\frac{r+1}{2}\right).

Note that 𝒌(𝒂)=𝒌(𝒃){\boldsymbol{k}}({\boldsymbol{a}})={\boldsymbol{k}}({\boldsymbol{b}}). According to [30], there exists h1Harm(q1)h_{1}\in\mathop{\rm Harm}\nolimits(q_{1}) such that

log|(dw)(r+12i)/2|h1+(bi+i)log|w|ki(𝒃)2log(log|w|)=O(1)\log|(dw)^{(r+1-2i)/2}|_{h_{1}}+(b_{i}+i)\log|w|-\frac{k_{i}({\boldsymbol{b}})}{2}\log\bigl{(}-\log|w|\bigr{)}=O(1)

around w=0w=0. We set h𝒂:=φ1(h1)Harm(q)h_{{\boldsymbol{a}}}:=\varphi^{-1}(h_{1})\in\mathop{\rm Harm}\nolimits(q). Note that

φ((dw)(r+12i)/2)=er+12i2r1z(dz)(r+12i)/2.\varphi^{\ast}((dw)^{(r+1-2i)/2})=e^{\frac{r+1-2i}{2r}\sqrt{-1}z}(dz)^{(r+1-2i)/2}.

Hence, we obtain the following estimates for i=0,,r1i=0,\ldots,r-1 as yy\to\infty:

log|(dz)(r+12i)/2|h𝒂(bi+r+12)yrki(𝒂)2log(y+2)=O(1).\log|(dz)^{(r+1-2i)/2}|_{h_{{\boldsymbol{a}}}}-\left(b_{i}+\frac{r+1}{2}\right)\frac{y}{r}-\frac{k_{i}({\boldsymbol{a}})}{2}\log(y+2)=O(1).

Because h𝒂h_{{\boldsymbol{a}}} is invariant under the translation by (x,y)(x+2π,y)(x,y)\longmapsto(x+2\pi,y), we obtain the desired estimate for h𝒂h_{{\boldsymbol{a}}}. ∎

Remark 3.26.

Later in §6.3, we shall see that the harmonic metric h𝐚h_{{\boldsymbol{a}}} is uniquely determined by the condition

log|(dz)(r+12i)/2|h𝒂aiy=O(log(2+y))\log\bigl{|}(dz)^{(r+1-2i)/2}\bigr{|}_{h_{{\boldsymbol{a}}}}-a_{i}y=O\bigl{(}\log(2+y)\bigr{)}

on {y0}\{y\geq 0\}. Moreover, for any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), there exists 𝐚Harm(q){\boldsymbol{a}}\in\mathop{\rm Harm}\nolimits(q) such that h=h𝐚h=h_{{\boldsymbol{a}}}.

Remark 3.27.

The examples in Proposition 3.25 will be used in the proof of Lemma 5.25.

4. Preliminary from the classical analysis

4.1. Subharmonic functions on sectors

4.1.1. Phragmén-Lindelöf theorem for subharmonic functions

For any 1/2<a1/2<a and 0R0\leq R, we set Da(R):={z||z|>R,|arg(z)|<π/2a}D_{a}(R):=\bigl{\{}z\in{\mathbb{C}}\,\big{|}\,|z|>R,\,|\arg(z)|<\pi/2a\bigr{\}}. Let D¯a(R)\overline{D}_{a}(R) denote the closure in {\mathbb{C}}.

Proposition 4.1 ([22, §7.3, Theorem 3]).

Let ff be a continuous function on D¯a(0)\overline{D}_{a}(0) which is subharmonic in Da(0)D_{a}(0). Assume that there exist C>0C>0, 0<ρ<a0<\rho<a, MM\in{\mathbb{R}} and an exhaustive family {Ki}\{K_{i}\} of Da(0)D_{a}(0) such that the following holds.

  • fC(|z|ρ+1)f\leq C(|z|^{\rho}+1) on Ki\partial K_{i}.

  • fMf\leq M on Da(0)\partial D_{a}(0).

Then, fMf\leq M on Da(0)D_{a}(0).

Proof   There exists C1>0C_{1}>0 such that fC1Re(zρ+1)f\leq C_{1}\mathop{\rm Re}\nolimits\bigl{(}z^{\rho}+1\bigr{)} on Ki\partial K_{i}. Because ff is assumed to be subharmonic, we obtain that fC1Re(zρ+1)f\leq C_{1}\mathop{\rm Re}\nolimits\bigl{(}z^{\rho}+1\bigr{)} on KiK_{i}, and hence on Da(0)D_{a}(0). There exists C2>0C_{2}>0 such that fC2(|z|ρ+1)f\leq C_{2}(|z|^{\rho}+1) on Da(0)D_{a}(0). Then, the claim follows from Phragmén-Lindelöf theorem for subharmonic functions (see [22, §7.3, Theorem 3]). ∎

Corollary 4.2.

Let ff be a subharmonic CC^{\infty}-function on a neighbourhood of D¯a(R)\overline{D}_{a}(R). Assume the following.

  • ff is bounded from above on D¯a(R)\partial\overline{D}_{a}(R).

  • There exist C>0C>0, 0<ρ<a0<\rho<a and an exhaustive family {Ki}\{K_{i}\} of Da(R)D_{a}(R) such that fC(|z|ρ+1)f\leq C(|z|^{\rho}+1) on Ki\partial K_{i}.

Then, ff is bounded from above on D¯a(R)\overline{D}_{a}(R).

Proof   There exists an {\mathbb{R}}-valued CC^{\infty}-function gg on a neighbourhood of D¯a(0)\overline{D}_{a}(0) such that g=fg=f on D¯a(R+1)\overline{D}_{a}(R+1). Let χ\chi be a CC^{\infty}-function on D¯a(0)\overline{D}_{a}(0) with compact support such that (i) 0χ10\leq\chi\leq 1, (ii) χ(z)=1\chi(z)=1 on {zD¯a(0)||z|R+1}\bigl{\{}z\in\overline{D}_{a}(0)\,\big{|}\,|z|\leq R+1\bigr{\}}. Note that (1χ)g0(1-\chi)\triangle g\geq 0 on D¯a(0)\overline{D}_{a}(0), where =4zz¯\triangle=4\partial_{z}\partial_{\overline{z}}. There exists a continuous bounded function GG on D¯a(0)\overline{D}_{a}(0) such that G=χg\triangle G=\chi\triangle g and that G|Da(0)=0G_{|\partial D_{a}(0)}=0. There exists C1>0C_{1}>0 such that gGC1g-G\leq C_{1} on D¯a(0)\partial\overline{D}_{a}(0). There exists C2>0C_{2}>0 such that gGC2Re(zρ+1)g-G\leq C_{2}\mathop{\rm Re}\nolimits\bigl{(}z^{\rho}+1\bigr{)} on Ki\partial K_{i}. Because gGg-G is subharmonic, we obtain that gG1C2Re(zρ+1)g-G_{1}\leq C_{2}\mathop{\rm Re}\nolimits\bigl{(}z^{\rho}+1\bigr{)} on KiK_{i}, and hence on Da(R)D_{a}(R). Because Da(0)Da(R)D_{a}(0)\setminus D_{a}(R) is relatively compact, there exists C3>0C_{3}>0 such that gG1C3(|z|ρ+1)g-G_{1}\leq C_{3}(|z|^{\rho}+1) on Da(0)D_{a}(0). By the Phragmén-Lindelöf theorem, we obtain the desired boundedness. ∎

Let us give a variant. Let b+b_{+}, bb_{-} and κ\kappa be non-negative numbers such that b:=max{b+,b}κb:=\max\{b_{+},b_{-}\}\leq\kappa.

Corollary 4.3.

Let ff be a subharmonic function on Da(R)D_{a}(R) which extends to a continuous function on D¯a(R)\overline{D}_{a}(R) such that the following holds.

  • For any δ>0\delta>0, there exists M,δ>0M_{-,\delta}>0 such that

    fM,δ(|z|b+1)f\leq M_{-,\delta}(|z|^{b_{-}}+1)

    on {zDa(R)|π/2a<arg(z)<δ}\bigl{\{}z\in D_{a}(R)\,\big{|}-\pi/2a<\arg(z)<-\delta\bigr{\}}.

  • For any δ>0\delta>0, there exists M+,δ>0M_{+,\delta}>0 such that

    fM+,δ(|z|b++1)f\leq M_{+,\delta}(|z|^{b_{+}}+1)

    on {zDa(R)|δ<arg(z)<π/2a}\bigl{\{}z\in D_{a}(R)\,\big{|}\,\delta<\arg(z)<\pi/2a\bigr{\}}.

  • There exist an exhaustive family {Ki}\{K_{i}\} of Da(R)D_{a}(R) and C>0C>0 such that

    fC(|z|κ+1)f\leq C(|z|^{\kappa}+1)

    on Ki\partial K_{i}.

Then, there exists C1>0C_{1}>0 such that fC1(|z|b+1)f\leq C_{1}(|z|^{b}+1) on Da(R)D_{a}(R).

Proof   We explain the case 0bb+0\leq b_{-}\leq b_{+}. The other case can be discussed similarly. For any a1>aa_{1}>a, we take a decreasing sequence cic_{i} in {a1<c<2a1}\{a_{1}<c<2a_{1}\} such that limci=a1\lim c_{i}=a_{1}. We set K1,i=KiDci(R)K_{1,i}=K_{i}\cap D_{c_{i}}(R). Then, {K1,i}\{K_{1,i}\} is an exhaustive family of Da1(R)D_{a_{1}}(R), and there exists C2>0C_{2}>0 such that fC2(|z|κ+1)f\leq C_{2}(|z|^{\kappa}+1) on K1,i\partial K_{1,i}. Hence, by making aa larger, we may assume that κ<a\kappa<a from the beginning.

There exists C3>0C_{3}>0 such that fC3Re(zb++1)0f-C_{3}\mathop{\rm Re}\nolimits\bigl{(}z^{b_{+}}+1\bigr{)}\leq 0 on {zDa(R)|π/4a<|arg(z)|<π/2a}\bigl{\{}z\in D_{a}(R)\,\big{|}\,\pi/4a<|\arg(z)|<\pi/2a\bigr{\}}. Then, we obtain the claim of the corollary from Proposition 4.1. ∎

4.1.2. Nevanlinna formula

We set :={z|Im(z)>0}\mathbb{H}:=\{z\in{\mathbb{C}}\,|\,\mathop{\rm Im}\nolimits(z)>0\}. For any R0R\geq 0, we set (R):={z||z|>R}\mathbb{H}(R):=\bigl{\{}z\in\mathbb{H}\,\big{|}\,|z|>R\bigr{\}}. Let ¯\overline{\mathbb{H}} and ¯(R)\overline{\mathbb{H}}(R) denote the closure of \mathbb{H} and ¯(R)\overline{\mathbb{H}}(R), respectively. Let (x,y)(x,y) be the real coordinate system on {\mathbb{C}} determined by z=x+1yz=x+\sqrt{-1}y.

Let ff be an {\mathbb{R}}-valued CC^{\infty}-function on ¯(R)\overline{\mathbb{H}}(R) for some R0R\geq 0. Suppose that there exist C>0C>0, 0<ρ<10<\rho<1 and δ0\delta\geq 0 such that the following conditions are satisfied.

  • ff is CC^{\infty} in (R)\mathbb{H}(R), and |(f)|C(1+y2)1δ|\triangle(f)|\leq C(1+y^{2})^{-1-\delta}. Here, =x2+y2\triangle=\partial_{x}^{2}+\partial_{y}^{2}.

  • There exists an exhaustive family {Ki}\{K_{i}\} of (R)\mathbb{H}(R) such that |f|y+C(|z|ρ+1)|f|\leq y+C(|z|^{\rho}+1) on Ki\partial K_{i} (i=1,2,)(i=1,2,\ldots).

Proposition 4.4.

There exist 1a(f)1-1\leq a(f)\leq 1 and a constant C1>0C_{1}>0 such that the following holds on (R)\mathbb{H}(R):

|fa(f)y|C1(|z|ρ+1).|f-a(f)y|\leq C_{1}(|z|^{\rho}+1).

If |f||f| is bounded on (R)\partial\mathbb{H}(R), we obtain the following stronger estimate on (R)\mathbb{H}(R) for a positive constant C2>0C_{2}>0:

|fa(f)y|C2log(2+y).|f-a(f)y|\leq C_{2}\log(2+y).

If |f||f| is bounded on (R)\partial\mathbb{H}(R), and if δ>0\delta>0, then |fa(f)y||f-a(f)y| is bounded on (R)\mathbb{H}(R).

Proof   There exists a CC^{\infty}-function f~\widetilde{f} on ¯\overline{\mathbb{H}} such that f~=f\widetilde{f}=f on ¯(R+1)\overline{\mathbb{H}}(R+1). There exists C~>0\widetilde{C}>0 such that the following holds.

  • |f~|C~(1+y2)1δ|\triangle\widetilde{f}|\leq\widetilde{C}(1+y^{2})^{-1-\delta}.

  • There exists an exhaustive family {K~i}\{\widetilde{K}_{i}\} such that |f~|y+C~(1+|z|ρ)|\widetilde{f}|\leq y+\widetilde{C}(1+|z|^{\rho}).

If |f||f| is bounded on (R)\partial\mathbb{H}(R), then |f~||\widetilde{f}| is bounded on \partial\mathbb{H}. It is enough to obtain the estimate for f~\widetilde{f}. Therefore, we may assume R=0R=0 from the beginning.

Let ζ\zeta be a complex variable, and let (ξ,η)(\xi,\eta) be the real variables determined by ζ=ξ+1η\zeta=\xi+\sqrt{-1}\eta. On {y>0}\{y>0\}, we consider the following function

F1(z):=14π𝑑ξ0𝑑η(f(ζ)log((xξ)2+(yη)2(xξ)2+(y+η)2)).F_{1}(z):=\frac{1}{4\pi}\int_{-\infty}^{\infty}d\xi\int_{0}^{\infty}\,d\eta\left(\triangle f(\zeta)\cdot\log\left(\frac{(x-\xi)^{2}+(y-\eta)^{2}}{(x-\xi)^{2}+(y+\eta)^{2}}\right)\right).

By Lemma 4.5 below, the integral is convergent, and we obtain the estimate

F1(z)={O(log(1+y))(δ=0)O(y(1+y)1)(δ>0)F_{1}(z)=\left\{\begin{array}[]{ll}O\bigl{(}\log(1+y)\bigr{)}&(\delta=0)\\ \\ O\bigl{(}y(1+y)^{-1}\bigr{)}&(\delta>0)\end{array}\right.

Moreover, we obtain F1=f\triangle F_{1}=\triangle f, and hence F2:=fF1F_{2}:=f-F_{1} is a harmonic function on \mathbb{H}, and continuous on ¯\overline{\mathbb{H}}.

We set ϕ:=Re(eπρ1/2(z+1)ρ)\phi:=\mathop{\rm Re}\nolimits\bigl{(}e^{-\pi\rho\sqrt{-1}/2}(z+\sqrt{-1})^{\rho}\bigr{)}, where we consider the branch of (z+1)ρ(z+\sqrt{-1})^{\rho} such that (b1+1)ρ=(b+1)ρeπρ1/2(b\sqrt{-1}+\sqrt{-1})^{\rho}=(b+1)^{\rho}e^{\pi\rho\sqrt{-1}/2} for b>0b>0. It is a harmonic function, and there exists ϵ>0\epsilon>0 such that ϵ(|z|ρ+1)ϕ\epsilon(|z|^{\rho}+1)\leq\phi on ¯\overline{\mathbb{H}}. By the assumption, there exist C10>0C_{10}>0 such that |F2|C10(y+ϕ)|F_{2}|\leq C_{10}\bigl{(}y+\phi\bigr{)} on Ki\partial K_{i} (i=1,2,)(i=1,2,\ldots). Because F2F_{2} is harmonic, we obtain |F2|C10(y+ϕ)|F_{2}|\leq C_{10}\bigl{(}y+\phi\bigr{)} on KiK_{i}, and hence on ¯\overline{\mathbb{H}}. In particular, we obtain |F2|C10ϕ|F_{2}|\leq C_{10}\phi on \partial\mathbb{H}, and F2+C10(y+ϕ)0F_{2}+C_{10}\bigl{(}y+\phi\bigr{)}\geq 0 on \mathbb{H}. Then, according to the Nevanlinna formula [22, §14, Theorem 1] (see also [22, Remark in page 101]), there exists a real number cc such that the following holds for y>0y>0:

F2+C10(y+ϕ)=cy+yπF2(t,0)+C10ϕ(t,0)dt(tx)2+y2.F_{2}+C_{10}\bigl{(}y+\phi\bigr{)}=cy+\frac{y}{\pi}\int_{-\infty}^{\infty}\frac{F_{2}(t,0)+C_{10}\phi(t,0)\,dt}{(t-x)^{2}+y^{2}}.

Note that

ϕ(x,y)=yπϕ(t,0)(tx)2+y2𝑑t.\phi(x,y)=\frac{y}{\pi}\int_{-\infty}^{\infty}\frac{\phi(t,0)}{(t-x)^{2}+y^{2}}\,dt.

We obtain

f=(cC10)y+F1+yπf(t,0)dt(tx)2+y2.f=(c-C_{10})y+F_{1}+\frac{y}{\pi}\int_{-\infty}^{\infty}\frac{f(t,0)\,dt}{(t-x)^{2}+y^{2}}.

Note that |f|=|F2|C10ϕ|f|=|F_{2}|\leq C_{10}\phi on \partial\mathbb{H}. We obtain that |yπf(t,0)dt(tx)2+y2|C10ϕ\Bigl{|}\frac{y}{\pi}\int_{-\infty}^{\infty}\frac{f(t,0)\,dt}{(t-x)^{2}+y^{2}}\Bigr{|}\leq C_{10}\phi on \mathbb{H}. Then, we can easily obtain the claims of the proposition. ∎

In the proof, we have used the following lemma.

Lemma 4.5.

There exists a constant C0>0C_{0}>0 such that the following holds for any zz\in\mathbb{H}:

0𝑑ξ0dη1+η2(log|ζz¯ζz|2)C0log(1+y).0\leq\int_{-\infty}^{\infty}\,d\xi\int_{0}^{\infty}\frac{d\eta}{1+\eta^{2}}\left(\log\left|\frac{\zeta-\overline{z}}{\zeta-z}\right|^{2}\right)\leq C_{0}\log(1+y). (41)

For any δ>0\delta>0, there exists a constant Cδ>0C_{\delta}>0 such that the following holds for any zz with y>0y>0:

0𝑑ξ0dη(1+η2)1+δ(log|ζz¯ζz|2)Cδy1+y.0\leq\int_{-\infty}^{\infty}\,d\xi\int_{0}^{\infty}\frac{d\eta}{(1+\eta^{2})^{1+\delta}}\left(\log\left|\frac{\zeta-\overline{z}}{\zeta-z}\right|^{2}\right)\leq C_{\delta}\frac{y}{1+y}. (42)

Proof   Let us study the case δ=0\delta=0. We set Z1:={(ξ,η)|η2y}Z_{1}:=\{(\xi,\eta)\,|\,\eta\geq 2y\}, Z2:={(ξ,η)|y2η4y}Z_{2}:=\{(\xi,\eta)\,|\,y\leq 2\eta\leq 4y\}, and Z3:={(ξ,η)| 0<2ηy}Z_{3}:=\{(\xi,\eta)\,|\,0<2\eta\leq y\}. We obtain

Z1dξdη1+η2log|ζz¯ζz|2=Z1dξdη1+η2log(ξ2+(η+y)2ξ2+(ηy)2)=Z1dξdη1+η2log(1+4yηξ2+(ηy)2)𝑑ξ2y21+y2η24ηdη(ξ2+(η1)2).\int_{Z_{1}}\frac{d\xi d\eta}{1+\eta^{2}}\log\left|\frac{\zeta-\overline{z}}{\zeta-z}\right|^{2}=\int_{Z_{1}}\frac{d\xi d\eta}{1+\eta^{2}}\log\left(\frac{\xi^{2}+(\eta+y)^{2}}{\xi^{2}+(\eta-y)^{2}}\right)\\ =\int_{Z_{1}}\frac{d\xi\,d\eta}{1+\eta^{2}}\log\left(1+\frac{4y\eta}{\xi^{2}+(\eta-y)^{2}}\right)\\ \leq\int_{-\infty}^{\infty}\,d\xi\int_{2}^{\infty}\frac{y^{2}}{1+y^{2}\eta^{2}}\frac{4\eta d\eta}{\bigl{(}\xi^{2}+(\eta-1)^{2}\bigr{)}}. (43)

For y1y\geq 1, (43) is dominated by

𝑑ξ24dηη(ξ2+(η1)2)<.\int_{-\infty}^{\infty}\,d\xi\int_{2}^{\infty}\frac{4d\eta}{\eta\bigl{(}\xi^{2}+(\eta-1)^{2}\bigr{)}}<\infty.

For 0<y10<y\leq 1, (43) is dominated by

C12y2dη1+y2η2C1y2dη1+y2η2C2y.C_{1}\int_{2}^{\infty}\frac{y^{2}\,d\eta}{1+y^{2}\eta^{2}}\leq C_{1}\int_{-\infty}^{\infty}\frac{y^{2}\,d\eta}{1+y^{2}\eta^{2}}\leq C_{2}y.

Similarly, we obtain

Z2dξdη1+η2log|ζz¯ζz|2=Z2dξdη1+η2log(1+4yηξ2+(yη)2)|ξ|1𝑑ξ122y21+y2η24ηdη(ξ2+(η1)2)+|ξ1𝑑ξ122y2dη1+y2η2log(1+ηξ2+(η1)2).\int_{Z_{2}}\frac{d\xi\,d\eta}{1+\eta^{2}}\log\left|\frac{\zeta-\overline{z}}{\zeta-z}\right|^{2}=\int_{Z_{2}}\frac{d\xi\,d\eta}{1+\eta^{2}}\log\left(1+\frac{4y\eta}{\xi^{2}+(y-\eta)^{2}}\right)\\ \leq\int_{|\xi|\geq 1}d\xi\int_{\frac{1}{2}}^{2}\frac{y^{2}}{1+y^{2}\eta^{2}}\frac{4\eta d\eta}{\bigl{(}\xi^{2}+(\eta-1)^{2}\bigr{)}}\\ +\int_{|\xi\leq 1}d\xi\int_{\frac{1}{2}}^{2}\frac{y^{2}\,d\eta}{1+y^{2}\eta^{2}}\log\left(1+\frac{\eta}{\xi^{2}+(\eta-1)^{2}}\right). (44)

If y1y\geq 1, (44) is dominated by

|ξ|1𝑑ξ1224dηη(ξ2+(η1)2)+|ξ1𝑑ξ122dηη2log(1+ηξ2+(η1)2)<\int_{|\xi|\geq 1}d\xi\int_{\frac{1}{2}}^{2}\frac{4d\eta}{\eta\bigl{(}\xi^{2}+(\eta-1)^{2}\bigr{)}}+\\ \int_{|\xi\leq 1}d\xi\int_{\frac{1}{2}}^{2}\frac{d\eta}{\eta^{2}}\log\left(1+\frac{\eta}{\xi^{2}+(\eta-1)^{2}}\right)<\infty (45)

If y1y\leq 1, (44) is dominated by

8y2|ξ|1dξξ2122𝑑η+y2|ξ|1𝑑ξ122𝑑ηlog(1+ηξ2+(η1)2)C3y2.8y^{2}\int_{|\xi|\geq 1}\frac{d\xi}{\xi^{2}}\int_{\frac{1}{2}}^{2}d\eta+y^{2}\int_{|\xi|\leq 1}d\xi\int_{\frac{1}{2}}^{2}d\eta\log\left(1+\frac{\eta}{\xi^{2}+(\eta-1)^{2}}\right)\leq C_{3}y^{2}.

As for the integral over Z3Z_{3}, we obtain

Z3dξdη1+η2log|ζz¯ζz|2Z3dξdη1+η24yηξ2+(ηy)2C100y/2dη1+η2yηyη2C100y/2ηdη1+η2C11log(1+y).\int_{Z_{3}}\frac{d\xi\,d\eta}{1+\eta^{2}}\log\left|\frac{\zeta-\overline{z}}{\zeta-z}\right|^{2}\leq\int_{Z_{3}}\frac{d\xi\,d\eta}{1+\eta^{2}}\frac{4y\eta}{\xi^{2}+(\eta-y)^{2}}\\ \leq C_{10}\int_{0}^{y/2}\frac{d\eta}{1+\eta^{2}}\frac{y\eta}{y-\eta}\leq 2C_{10}\int_{0}^{y/2}\frac{\eta d\eta}{1+\eta^{2}}\leq C_{11}\log(1+y). (46)

Thus, we obtain (41) in the case δ=0\delta=0.

As for the case δ>0\delta>0, the integrals over ZiZ_{i} (i=1,2)(i=1,2) are dominated as in the case of (41). The integral over Z3Z_{3} is dominated by 0y/2(1+η2)1δη𝑑η<C20(δ)y(1+y)1\int_{0}^{y/2}(1+\eta^{2})^{-1-\delta}\,\eta\,d\eta<C_{20}(\delta)y(1+y)^{-1}. Thus, we obtain (42). ∎

4.1.3. Holomorphic line bundles with a Hermitian metric on upper half plane

We state a consequence of the Nevanlinna formula (Proposition 4.4) on a holomorphic line bundle LL with a Hermitian metric hh on \mathbb{H}, which is fundamental in the classification of solutions of Toda equations in terms of parabolic structures. Let R(h)R(h) denote the curvature of the Chern connection associated with (L,h)(L,h). By using the standard Euclidean metric g=dzdz¯g=dz\,d\overline{z}, we assume the following condition.

  • |R(h)|h,g=O((1+y2)1)|R(h)|_{h,g}=O\bigl{(}(1+y^{2})^{-1}\bigr{)}.

Note that this condition is analogue to the acceptability condition in [38, 39].

Lemma 4.6.

Let vv be a global frame of LL. If there exist C>0C>0, 0<ρ<10<\rho<1, and an exhaustive family {Ki}\{K_{i}\} of \mathbb{H} such that |log|v|h|y+C(|z|ρ+1)\bigl{|}\log|v|_{h}\bigr{|}\leq y+C(|z|^{\rho}+1) on Ki\partial K_{i}, then there exists 1a(h,v)1-1\leq a(h,v)\leq 1 such that log|v|ha(h,v)y=O(|z|ρ+1)\log|v|_{h}-a(h,v)y=O(|z|^{\rho}+1) on \mathbb{H}. If log|v|h\log|v|_{h} is bounded on \partial\mathbb{H}, then we obtain log|v|ha(h,v)y=O(log(y+2))\log|v|_{h}-a(h,v)y=O\bigl{(}\log(y+2)\bigr{)} on \mathbb{H}.

Proof   Because R(h)=14(log|v|h2)dz¯dzR(h)=\frac{1}{4}\triangle(\log|v|^{2}_{h})\,d\overline{z}\,dz, we obtain (log|v|h2)=O((1+y2)1)\triangle(\log|v|_{h}^{2})=O\bigl{(}(1+y^{2})^{-1}\bigr{)}. Then, the claim follows from Proposition 4.4. ∎

Definition 4.7.

The number a(h,v)a(h,v) is called the parabolic order of vv with respect to hh.

4.2. Finite exponential sums and their perturbation

Let us recall some results in [41, 48], which are summarized in [21].

4.2.1. Finite exponential sums

Let c0<c1<c2<<cnc_{0}<c_{1}<c_{2}<\cdots<c_{n} be real numbers. Let aia_{i} (i=0,,n)(i=0,\ldots,n) be non-zero complex numbers. We consider the entire function

F(ζ)=i=0naie1ciζ.F(\zeta)=\sum_{i=0}^{n}a_{i}e^{\sqrt{-1}c_{i}\zeta}.

We set Z(F):={ζ|F(ζ)=0}Z(F):=\{\zeta\in{\mathbb{C}}\,|\,F(\zeta)=0\}. It is easy to see that there exists L>0L>0 such that any ζZ(F)\zeta\in Z(F) satisfies |Im(ζ)|<L|\mathop{\rm Im}\nolimits(\zeta)|<L. For any x1<x2x_{1}<x_{2}, we set Y(x1,x2):={ζ||Im(ζ)|L,x1Re(ζ)x2}Y(x_{1},x_{2}):=\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,|\mathop{\rm Im}\nolimits(\zeta)|\leq L,\,\,x_{1}\leq\mathop{\rm Re}\nolimits(\zeta)\leq x_{2}\bigr{\}}.

Proposition 4.8 ([21, 41, 48]).

If Z(F){Re(ζ)=xi}=Z(F)\cap\bigcup\{\mathop{\rm Re}\nolimits(\zeta)=x_{i}\}=\emptyset, we obtain

n+cnc02π(x2x1)#(Y(x1,x2)Z(F))n+cnc02π(x2x1).-n+\frac{c_{n}-c_{0}}{2\pi}(x_{2}-x_{1})\leq\#\Bigl{(}Y(x_{1},x_{2})\cap Z(F)\Bigr{)}\leq n+\frac{c_{n}-c_{0}}{2\pi}(x_{2}-x_{1}).

It particularly implies that the order of the zero of FF at any point is not larger than nn. ∎

Note that the proposition implies that Z(F){Re(ζ)=xi}nZ(F)\cap\{\mathop{\rm Re}\nolimits(\zeta)=x_{i}\}\leq n. Hence, we obtain the following for any x1<x2x_{1}<x_{2}:

3n+cnc02π(x2x1)#(Y(x1,x2)Z(F))3n+cnc02π(x2x1).-3n+\frac{c_{n}-c_{0}}{2\pi}(x_{2}-x_{1})\leq\#\Bigl{(}Y(x_{1},x_{2})\cap Z(F)\Bigr{)}\leq 3n+\frac{c_{n}-c_{0}}{2\pi}(x_{2}-x_{1}).

4.2.2. Perturbation

Let GG be a holomorphic function defined on W(C1,C2):={ζ|Re(ζ)>C1,|Im(ζ)|<C2}W(C_{1},C_{2}):=\{\zeta\in{\mathbb{C}}\,|\,\mathop{\rm Re}\nolimits(\zeta)>C_{1},\,\,|\mathop{\rm Im}\nolimits(\zeta)|<C_{2}\} such that

limRe(ζ)|F(ζ)G(ζ)|=0.\lim_{\mathop{\rm Re}\nolimits(\zeta)\to\infty}|F(\zeta)-G(\zeta)|=0.

We set Z(G):={ζW(C1,C2)|G(ζ)=0}Z(G):=\{\zeta\in W(C_{1},C_{2})\,|\,G(\zeta)=0\}. For any closed subset KK\subset{\mathbb{C}} and ζ\zeta\in{\mathbb{C}}, let d(ζ,K)d(\zeta,K) denote the Euclidean distance between ζ\zeta and AA, i.e., d(ζ,A)=min{|ζζ||ζA}d(\zeta,A)=\min\bigl{\{}|\zeta^{\prime}-\zeta|\,\big{|}\,\zeta^{\prime}\in A\bigr{\}}.

Proposition 4.9.

Take 0<C2<C20<C_{2}^{\prime}<C_{2}.

  • There exist C3>0C_{3}>0 such that the following holds for any C1+1<x1<x2C_{1}+1<x_{1}<x_{2}:

    #(Y(x1,x2)W(C1,C2)Z(G))C3(1+|x2x1|).\#\Bigl{(}Y(x_{1},x_{2})\cap W(C_{1},C_{2}^{\prime})\cap Z(G)\Bigr{)}\leq C_{3}(1+|x_{2}-x_{1}|).
  • There exist C1>C1C_{1}^{\prime}>C_{1} and C4>0C_{4}>0 such that |G(ζ)|C4δn|G(\zeta)|\geq C_{4}\delta^{n} holds for any 0<δ10<\delta\leq 1 and any ζW(C1,C2)\zeta\in W(C_{1}^{\prime},C_{2}^{\prime}) satisfying d(ζ,Z(G))δd(\zeta,Z(G))\geq\delta.

Proof   We apply the argument in [48]. Let us prove the first claim in the case |x2x1|=1|x_{2}-x_{1}|=1. Assume that there exists a sequence xjx_{j} such that

#(Y(xj,xj+1)W(C1,C2)Z(G)).\#\Bigl{(}Y(x_{j},x_{j}+1)\cap W(C_{1},C_{2}^{\prime})\cap Z(G)\Bigr{)}\to\infty.

Let

Ψj:{ζ|ϵ<Re(ζ)<1+ϵ,|Im(ζ)|<C2}{ζ|xjϵ<Re(ζ)<xj+1+ϵ,|Im(ζ)|<C2}\Psi_{j}:\{\zeta\in{\mathbb{C}}\,|\,-\epsilon<\mathop{\rm Re}\nolimits(\zeta)<1+\epsilon,\,|\mathop{\rm Im}\nolimits(\zeta)|<C_{2}\}\simeq\\ \{\zeta\in{\mathbb{C}}\,|\,x_{j}-\epsilon<\mathop{\rm Re}\nolimits(\zeta)<x_{j}+1+\epsilon,|\mathop{\rm Im}\nolimits(\zeta)|<C_{2}\} (47)

be the isomorphism defined by Ψj(ζ)=ζ+xj\Psi_{j}(\zeta)=\zeta+x_{j}. Going into a subsequence, we may assume that the sequence Ψj(F)\Psi_{j}^{\ast}(F) is convergent to bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta}, where bib_{i} are complex numbers such that |bi|=|ai||b_{i}|=|a_{i}|. It implies that the sequence Ψj(G)\Psi_{j}^{\ast}(G) is convergent to bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta}. Then, we obtain the contradiction by Proposition 4.8 and Lemma 4.12 below. Hence, there exists a constant C10>0C_{10}>0 such that

#(Y(x,x+1)W(C1,C2)Z(G))<C10\#\Bigl{(}Y(x,x+1)\cap W(C_{1},C_{2}^{\prime})\cap Z(G)\Bigr{)}<C_{10}

for any x>C1+1x>C_{1}+1.

Take any C1+1<x1<x2C_{1}+1<x_{1}<x_{2}. We set m0:=min{m|x2x1m}1m_{0}:=\min\{m\in{\mathbb{Z}}\,|\,x_{2}-x_{1}\leq m\}\geq 1. Then, we obtain

#(Y(x1,x2)W(C1,C2)Z(G))#(Y(x1,x1+m0)W(C1,C2)Z(G))m0C10(1+|x2x0|)C10.\#\Bigl{(}Y(x_{1},x_{2})\cap W(C_{1},C_{2}^{\prime})\cap Z(G)\Bigr{)}\leq\\ \#\Bigl{(}Y(x_{1},x_{1}+m_{0})\cap W(C_{1},C_{2}^{\prime})\cap Z(G)\Bigr{)}\leq m_{0}C_{10}\\ \leq(1+|x_{2}-x_{0}|)C_{10}. (48)

Thus, we obtain the first claim.

To prove the second claim, we prepare the following lemma.

Lemma 4.10.

There exists C1′′>C1C_{1}^{\prime\prime}>C_{1} such that the order of the zero of GG at any point of W(C1′′,C2)¯Z(G)\overline{W(C_{1}^{\prime\prime},C^{\prime}_{2})}\cap Z(G) is not larger than nn.

Proof   Suppose the contrary. Let ζjW(C1,C2)¯Z(G)\zeta_{j}\in\overline{W(C_{1},C_{2}^{\prime})}\cap Z(G) such that the order of zero of GG at ζj\zeta_{j} is strictly larger than nn. Let

Ψj:{ζ|1<Re(ζ)<1,|Im(ζ)|<C2}{ζ|1<Re(ζ)Re(ζj)<1,|Im(ζ)|<C2}\Psi_{j}:\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,-1<\mathop{\rm Re}\nolimits(\zeta)<1,\,|\mathop{\rm Im}\nolimits(\zeta)|<C_{2}\bigr{\}}\simeq\\ \bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,-1<\mathop{\rm Re}\nolimits(\zeta)-\mathop{\rm Re}\nolimits(\zeta_{j})<1,\,|\mathop{\rm Im}\nolimits(\zeta)|<C_{2}\bigr{\}} (49)

be the isomorphism determined by Ψj(ζ)=ζ+Re(ζj)\Psi_{j}(\zeta)=\zeta+\mathop{\rm Re}\nolimits(\zeta_{j}). We may assume that the sequence Ψj(F)\Psi_{j}^{\ast}(F) is convergent to a holomorphic function of the form bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta}, where bib_{i} are complex numbers such that |bi|=|ai||b_{i}|=|a_{i}|. It particularly implies that Ψj(G)\Psi_{j}^{\ast}(G) is convergent to bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta}. We may also assume that 1Im(ζj)\sqrt{-1}\mathop{\rm Im}\nolimits(\zeta_{j}) is convergent to 1β\sqrt{-1}\beta for some C2βC2-C_{2}^{\prime}\leq\beta\leq C_{2}^{\prime}. Because the order of zero of Ψj(G)\Psi_{j}^{\ast}(G) at 1Im(ζj)\sqrt{-1}\mathop{\rm Im}\nolimits(\zeta_{j}) are strictly larger than nn, we obtain that the order of bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta} is strictly larger than nn. But, it contradicts Proposition 4.8. ∎

Let us prove the second claim. Suppose the contrary. There exist sequences ϵj>0\epsilon_{j}>0, 0<δj10<\delta_{j}\leq 1 and ζjW(C1′′,C2)\zeta_{j}\in W(C_{1}^{\prime\prime},C_{2}^{\prime}) satisfying limϵj=0\lim\epsilon_{j}=0, d(ζj,Z(G))δjd(\zeta_{j},Z(G))\geq\delta_{j} and |G(ζ)|ϵjδjn|G(\zeta)|\leq\epsilon_{j}\delta_{j}^{n}. We shall deduce a contradiction.

Let us consider the case that the sequence ζj\zeta_{j} contains a bounded sequence. By going to a subsequence, the sequence ζj\zeta_{j} is convergent to ζW(C1′′,C2)¯\zeta_{\infty}\in\overline{W(C_{1}^{\prime\prime},C_{2}^{\prime})}. It is easy to observe that ζZ(G)\zeta_{\infty}\in Z(G), and the order of zero at ζ\zeta_{\infty} is strictly larger than nn. Hence, we obtain a contradiction by Lemma 4.10.

Let us consider the case that the sequence ζj\zeta_{j} does not contain a bounded sequence. In particular, Re(ζj)\mathop{\rm Re}\nolimits(\zeta_{j})\to\infty as jj\to\infty. We consider the maps Ψj\Psi_{j} as in (49). We may assume that the sequence Ψj(F)\Psi_{j}^{\ast}(F) is convergent to a holomorphic function of the form bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta}, where bib_{i} are complex numbers such that |bi|=|ai||b_{i}|=|a_{i}|. It particularly implies that Ψj(G)\Psi_{j}^{\ast}(G) is convergent to bie1ciζ\sum b_{i}e^{\sqrt{-1}c_{i}\zeta}. Then, we obtain a contradiction by Proposition 4.8 and Lemma 4.13 below. Thus, we obtain the second claim of the proposition. ∎

Corollary 4.11.

For any 0<δ10<\delta\leq 1, there exists C5>0C_{5}>0 such that we obtain |G(ζ)|C4δn|G(\zeta)|\geq C_{4}\delta^{n} for any ζW(C1,C2)\zeta\in W(C_{1},C_{2}^{\prime}) satisfying d(ζ,Z(G))δd(\zeta,Z(G))\geq\delta. ∎

4.2.3. Appendix

Let YY\subset{\mathbb{C}} be a connected open subset. Let Y1Y_{1} be a relatively compact open subset of YY. Let Fi:YF_{i}:Y\longrightarrow{\mathbb{C}} (i=1,2,)(i=1,2,\ldots) be holomorphic functions which uniformly converge to a holomorphic function F:YF_{\infty}:Y\longrightarrow{\mathbb{C}}. Note that any derivatives zjFi\partial_{z}^{j}F_{i} converges to zjF\partial_{z}^{j}F_{\infty}. Assume that FF_{\infty} is not constantly 0.

For i=1,2,,i=1,2,\ldots,\infty, we set Z(Fi):={ζY|Fi(ζ)=0}Z(F_{i}):=\{\zeta\in Y\,|\,F_{i}(\zeta)=0\}. Let 𝔪i(ζ)>0\mathfrak{m}_{i}(\zeta)\in{\mathbb{Z}}_{>0} denote the order of the zero of FiF_{i} at ζ\zeta.

Lemma 4.12.

There exists i0i_{0} such that the following holds for any ii0i\geq i_{0}:

ζZ(Fi)Y¯1𝔪i(ζ)ζZ(F)Y¯1𝔪(ζ).\sum_{\zeta\in Z(F_{i})\cap\overline{Y}_{1}}\mathfrak{m}_{i}(\zeta)\leq\sum_{\zeta\in Z(F_{\infty})\cap\overline{Y}_{1}}\mathfrak{m}_{\infty}(\zeta).

Proof   We may assume that Y1Y_{1} is simply connected. There exists a relatively compact open subset Y2YY_{2}\subset Y such that (i) Y2\partial Y_{2} is smooth, (ii) Y1Y2Y_{1}\subset Y_{2}, (iii) Z(F)Y¯1=Z(F)Y¯2Z(F_{\infty})\cap\overline{Y}_{1}=Z(F_{\infty})\cap\overline{Y}_{2}. The condition (iii) implies that F(ζ)0F_{\infty}(\zeta)\neq 0 for any ζY2\zeta\in\partial Y_{2}. There exists ϵ>0\epsilon>0 such that minζY2|F(ζ)|>2ϵ\min_{\zeta\in\partial Y_{2}}|F_{\infty}(\zeta)|>2\epsilon. There exists i0i_{0} such that |Fi(ζ)F(ζ)|ϵ/2|F_{i}(\zeta)-F_{\infty}(\zeta)|\leq\epsilon/2 for any ζY2\zeta\in\partial Y_{2} and ii0i\geq i_{0}. Then, the claim holds according to Rouché’s theorem. ∎

Let nn be a positive integer such that 𝔪(ζ)n\mathfrak{m}_{\infty}(\zeta)\leq n for any ζY¯1\zeta\in\overline{Y}_{1}. For any closed subset KK\subset{\mathbb{C}}, let d(ζ,K)d(\zeta,K) denote the Euclidean distance between ζ\zeta and KK. There exists δ0>0\delta_{0}>0 such that {ζ|d(ζ,Y¯1)2δ0}Y\{\zeta\in{\mathbb{C}}\,|\,d(\zeta,\overline{Y}_{1})\leq 2\delta_{0}\}\subset Y. For any 0<δ<δ00<\delta<\delta_{0}, we set Wi(δ):={ζY1|d(ζ,Z(Fi))δ}W_{i}(\delta):=\{\zeta\in Y_{1}\,|\,d(\zeta,Z(F_{i}))\geq\delta\}.

Lemma 4.13.

There exist C>0C>0 and i0i_{0} such that we obtain |Fi|Cδn|F_{i}|\geq C\delta^{n} on Wi(δ)W_{i}(\delta) for any ii0i\geq i_{0} and any 0<δ<δ00<\delta<\delta_{0}.

Proof   Assume the contrary. Then, there exist a sequence of positive numbers C1>C2>C_{1}>C_{2}>\cdots, a sequence i(j)i(j), positive numbers δj>0\delta_{j}>0 and ζjWi(j)(δj)\zeta_{j}\in W_{i(j)}(\delta_{j}) such that limCj=0\lim C_{j}=0 and that |Fi(j)(ζj)|<Cjδjn|F_{i(j)}(\zeta_{j})|<C_{j}\delta_{j}^{n}. Going to a subsequence, we may assume that i(j)=ji(j)=j. We may assume δj=d(ζj,Z(Fj))\delta_{j}=d(\zeta_{j},Z(F_{j})). Going to a subsequence, we may assume that the sequence ζj\zeta_{j} is convergent to ζY¯1\zeta_{\infty}\in\overline{Y}_{1}. Because F(ζ)=limFj(ζj)F_{\infty}(\zeta_{\infty})=\lim F_{j}(\zeta_{j}), we obtain F(ζ)=0F_{\infty}(\zeta_{\infty})=0, and hence ζZ(F)Y¯1\zeta_{\infty}\in Z(F_{\infty})\cap\overline{Y}_{1}. For any ϵ>0\epsilon>0, there exists j0j_{0} such that d(ζ,Z(Fj))<ϵd(\zeta_{\infty},Z(F_{j}))<\epsilon for any jj0j\geq j_{0}. Hence, we obtain limδj=0\lim\delta_{j}=0. Going to a subsequence, we may assume that the sequence δj\delta_{j} is decreasing.

Fix a relatively compact open neighbourhood UU of ζ\zeta_{\infty} in YY such that U¯Z(F)={ζ}\overline{U}\cap Z(F_{\infty})=\{\zeta_{\infty}\}. We set :=𝔪(ζ)n\ell:=\mathfrak{m}_{\infty}(\zeta_{\infty})\leq n. Going to a subsequence, we may assume that FjF_{j} is nowhere vanishing on U\partial U for any jj. We set Sj:=UZ(Fj)S_{j}:=U\cap Z(F_{j}). We obtain aSj𝔪j(a)=\sum_{a\in S_{j}}\mathfrak{m}_{j}(a)=\ell. We set

GU,j(ζ):=Fj|U(ζ)aSj(ζa)𝔪j(a),G_{U,j}(\zeta):=F_{j|U}(\zeta)\cdot\prod_{a\in S_{j}}(\zeta-a)^{-\mathfrak{m}_{j}(a)},

which is nowhere vanishing on UU for any jj. Because the sequence FjF_{j} is convergent to FF_{\infty}, the sequence aSj(ζa)𝔪j(a)\prod_{a\in S_{j}}(\zeta-a)^{\mathfrak{m}_{j}(a)} is convergent to (ζζ)(\zeta-\zeta_{\infty})^{\ell}. The sequence GU,jG_{U,j} is convergent to GU,(ζ):=F|U(ζ)(ζζ)G_{U,\infty}(\zeta):=F_{\infty|U}(\zeta)(\zeta-\zeta_{\infty})^{-\ell}, which is also nowhere vanishing on UU.

By the assumption, |ζja|δj|\zeta_{j}-a|\geq\delta_{j} for any aSja\in S_{j}. Hence, aSj(ζja)𝔪j(a)δj\prod_{a\in S_{j}}(\zeta_{j}-a)^{\mathfrak{m}_{j}(a)}\geq\delta_{j}^{\ell}. Then, there exists A>0A>0 such that the following holds for any jj:

|Fj(ζj)|Aδj.|F_{j}(\zeta_{j})|\geq A\cdot\delta_{j}^{\ell}.

Hence, we obtain Aδj|Fj(ζj)|CjδjnA\delta_{j}^{\ell}\leq|F_{j}(\zeta_{j})|\leq C_{j}\delta_{j}^{n}. It implies 0<ACjδjn00<A\leq C_{j}\delta_{j}^{n-\ell}\to 0, which is a contradiction. ∎

4.3. Holomorphic functions with multiple growth orders

4.3.1. Growth orders

Let ϖ:~\varpi:\widetilde{{\mathbb{C}}}\longrightarrow{\mathbb{C}} be the oriented real blowing up at 0. We set :={0}{\mathbb{C}}^{\ast}:={\mathbb{C}}\setminus\{0\}. Let ι:~\iota:{\mathbb{C}}^{\ast}\longrightarrow\widetilde{{\mathbb{C}}} denote the natural inclusion. Let zz denote the standard coordinate of {\mathbb{C}}. We identify ~\widetilde{{\mathbb{C}}} with 0×S1{\mathbb{R}}_{\geq 0}\times S^{1} by the polar decomposition of zz.

Notation 4.14.

Let Q=(0,e1θ0)ϖ1(0)Q=(0,e^{\sqrt{-1}\theta_{0}})\in\varpi^{-1}(0). Let ι(𝒪)Q\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} denote the stalk of the sheaf ι(𝒪)\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}}) at QQ. We fix a branch of logz\log z around QQ.

  • Let (Q)ι(𝒪)Q\mathfrak{I}(Q)\subset\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} be the {\mathbb{C}}-linear subspace generated by zaz^{-a} (a>0)(a>0), i.e., (Q)=a>0za\mathfrak{I}(Q)=\bigoplus_{a>0}{\mathbb{C}}\,z^{-a}.

  • Note that for any 𝔞(Q){0}\mathfrak{a}\in\mathfrak{I}(Q)\setminus\{0\}, |𝔞|1Re(𝔞)|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}) induces a continuous function on a neighbourhood of QQ. The continuous function is also denoted by |𝔞|1Re(𝔞)|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}).

  • For any 𝔞(Q)0\mathfrak{a}\in\mathfrak{I}(Q)\neq 0, we have the expression 𝔞=αzρ+0<c<ραczc\mathfrak{a}=\alpha z^{-\rho}+\sum_{0<c<\rho}\alpha_{c}z^{-c} where α0\alpha\neq 0. We set deg(𝔞):=ρ\deg(\mathfrak{a}):=\rho.

  • For 𝔞,𝔟(Q)\mathfrak{a},\mathfrak{b}\in\mathfrak{I}(Q), we say 𝔞Q𝔟\mathfrak{a}\prec_{Q}\mathfrak{b} if |𝔟𝔞|1Re(𝔟𝔞)(Q)>0|\mathfrak{b}-\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a})(Q)>0 or 𝔞=𝔟\mathfrak{a}=\mathfrak{b}. It defines a partial order Q\prec_{Q} on (Q)\mathfrak{I}(Q).

Remark 4.15.

Note that the orders Q\prec_{Q} are opposite to the order used in [29].

4.3.2. Regularly bounded holomorphic functions

Definition 4.16.

We say that fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} is regularly bounded if there exist a neighbourhood 𝒰\mathcal{U} of QQ in ~\widetilde{{\mathbb{C}}}, non-zero complex numbers α1,,αm\alpha_{1},\ldots,\alpha_{m}, and mutually distinct real numbers c1,c2,,cmc_{1},c_{2},\ldots,c_{m} such that the following holds.

  • ff induces a holomorphic function on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0).

  • |f(z)i=1mαiz1ci|0|f(z)-\sum_{i=1}^{m}\alpha_{i}z^{\sqrt{-1}c_{i}}|\to 0 as |z|0|z|\to 0 in 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0).

4.3.3. Holomorphic functions with single growth order

Definition 4.17.

We say that fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has a single growth order if there exist 𝔞(f,Q)(Q)\mathfrak{a}(f,Q)\in\mathfrak{I}(Q), a(f,Q)a(f,Q)\in{\mathbb{R}}, j(f,Q)0j(f,Q)\in{\mathbb{Z}}_{\geq 0} such that e𝔞(f,Q)za(f,Q)(logz)j(f,Q)fe^{-\mathfrak{a}(f,Q)}z^{-a(f,Q)}(\log z)^{-j(f,Q)}f is regularly bounded.

Definition 4.18.

Suppose that fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has single growth order at QQ.

  • We say that ff is simply positive (resp. negative) at QQ if

    |𝔞(f,Q)|1Re(𝔞(f,Q))(Q)>0(resp. |𝔞(f,Q)|1Re(𝔞(f,Q))(Q)<0).|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q))(Q)>0\quad\mbox{(resp. $|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q))(Q)<0$)}.
  • If 𝔞(f,Q)=0\mathfrak{a}(f,Q)=0, then we say that ff is neutral at QQ.

  • If 𝔞(f,Q)0\mathfrak{a}(f,Q)\neq 0 but |𝔞(f,Q)|1Re(𝔞(f,Q))(Q)=0|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q))(Q)=0, we say that ff is turning at QQ.

The following lemma is easy to see.

Lemma 4.19.

If fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has a single growth order, there exists a neighbourhood 𝒰\mathcal{U} of QQ such that (i) 𝒰ϖ1(0)\mathcal{U}\cap\varpi^{-1}(0) is connected, (ii) ff induces a holomorphic function on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0), (iii) ff has a single growth order at any point Q𝒰ϖ1(0)Q^{\prime}\in\mathcal{U}^{\prime}\cap\varpi^{-1}(0), (iv) 𝔞(f,Q)=𝔞(f,Q)\mathfrak{a}(f,Q^{\prime})=\mathfrak{a}(f,Q). ∎

Note that (iv) implies that the conditions “simply positive”, “simply negative” and “neutral” are preserved if 𝒰\mathcal{U} is sufficiently small. If ff is turning at QQ, |𝔞(f,Q)|1Re(𝔞(f,Q))|\mathfrak{a}(f,Q)|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q)) is positive on a connected component of (𝒰ϖ1(0)){Q}(\mathcal{U}\cap\varpi^{-1}(0))\setminus\{Q\} and negative on the other connected component of (𝒰ϖ1(0)){Q}(\mathcal{U}\cap\varpi^{-1}(0))\setminus\{Q\}.

Lemma 4.20.

Let ff be a section of ι𝒪\iota_{\ast}\mathcal{O}_{{\mathbb{C}}^{\ast}} on an open subset V~V\subset\widetilde{{\mathbb{C}}}. Let IVϖ1(0)I\subset V\cap\varpi^{-1}(0) be an interval. If ff has single growth order at each QIQ\in I, then we obtain 𝔞(f,Q1)=𝔞(f,Q2)\mathfrak{a}(f,Q_{1})=\mathfrak{a}(f,Q_{2}) for any Q1,Q2IQ_{1},Q_{2}\in I.

Proof   It follows from Lemma 4.19. ∎

4.3.4. Holomorphic functions with multiple growth orders

Definition 4.21.

We say that fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has multiple growth orders if ff is expressed as a finite sum i=1Nfi\sum_{i=1}^{N}f_{i} such that (i) each fif_{i} has a single growth order, (ii) 𝔞(fi,Q)𝔞(fj,Q)\mathfrak{a}(f_{i},Q)\neq\mathfrak{a}(f_{j},Q) (ij)(i\neq j).

Lemma 4.22.

Suppose that fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has multiple growth orders.

  • There exists a finite subset (Q)\mathcal{I}\subset\mathfrak{I}(Q) and an expression f=𝔞f𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}}f_{\mathfrak{a}} such that (i) for two distinct elements 𝔞,𝔟\mathfrak{a},\mathfrak{b}\in\mathcal{I}, neither 𝔞Q𝔟\mathfrak{a}\prec_{Q}\mathfrak{b} nor 𝔟Q𝔞\mathfrak{b}\prec_{Q}\mathfrak{a} holds, (ii) each f𝔞f_{\mathfrak{a}} has single growth order with 𝔞(f𝔞,Q)=𝔞\mathfrak{a}(f_{\mathfrak{a}},Q)=\mathfrak{a}. Such an expression is called reduced.

  • There exists a neighbourhood 𝒰\mathcal{U} of QQ such that (i) 𝒰ϖ1(0)\mathcal{U}\cap\varpi^{-1}(0) is connected, (ii) ff induces a holomorphic function on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0), (iii) ff has single growth order at any point Q(𝒰ϖ1(0)){Q}Q^{\prime}\in\bigl{(}\mathcal{U}\cap\varpi^{-1}(0)\bigr{)}\setminus\{Q\}, and 𝔞(f,Q)\mathfrak{a}(f,Q^{\prime}) is the unique maximal element of the partially ordered set (,Q)(\mathcal{I},\prec_{Q^{\prime}}).

Proof   There exists an expression f=i=1mfif=\sum_{i=1}^{m}f_{i} as in Definition 4.21. Let \mathcal{I} denote the set of the maximal elements in the partially ordered set {𝔞(f1,Q),,𝔞(fm,Q)}\{\mathfrak{a}(f_{1},Q),\ldots,\mathfrak{a}(f_{m},Q)\} with Q\prec_{Q}. There exists a decomposition

{𝔞(f1,Q),,𝔞(fm,Q)}=𝔞𝒥𝔞\{\mathfrak{a}(f_{1},Q),\ldots,\mathfrak{a}(f_{m},Q)\}=\coprod_{\mathfrak{a}\in\mathcal{I}}\mathcal{J}_{\mathfrak{a}}

such that any 𝔟𝒥𝔞\mathfrak{b}\in\mathcal{J}_{\mathfrak{a}} satisfy 𝔟Q𝔞\mathfrak{b}\prec_{Q}\mathfrak{a}. We set f𝔞:=𝔞(fi,Q)𝒥𝔞fif_{\mathfrak{a}}:=\sum_{\mathfrak{a}(f_{i},Q)\in\mathcal{J}_{\mathfrak{a}}}f_{i}. We obtain the expression f=𝔞f𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}}f_{\mathfrak{a}}. It is easy to see that f𝔞f_{\mathfrak{a}} has a single growth order with 𝔞(f𝔞,Q)=𝔞\mathfrak{a}(f_{\mathfrak{a}},Q)=\mathfrak{a}, and the first claim is proved. The second claim is clear. ∎

Note that a reduced expression f=𝔞f𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}}f_{\mathfrak{a}} is not uniquely determined.

Notation 4.23.

Let VV be an open subset of ~\widetilde{{\mathbb{C}}}. Let ff be a holomorphic function on Vϖ1(0)V\setminus\varpi^{-1}(0) which has multiple growth orders at any point of Vϖ1(0)V\cap\varpi^{-1}(0). Then, let 𝒵(f)\mathcal{Z}(f) denote the set of the points QVϖ1(0)Q\in V\cap\varpi^{-1}(0) such that one of the following condition is satisfied.

  • ff does not have single growth order at QQ.

  • ff has single growth order and turning at QQ.

Note that 𝒵(f)\mathcal{Z}(f) is discrete in Vϖ1(0)V\cap\varpi^{-1}(0).

Remark 4.24.

Let UU be a neighbourhood of 0 in {\mathbb{C}}. Suppose that a non-zero holomorphic function ff on U{0}U\setminus\{0\} satisfies a linear differential equation znf+j=0n1aj(z)zjf=0\partial_{z}^{n}f+\sum_{j=0}^{n-1}a_{j}(z)\partial_{z}^{j}f=0, where aja_{j} are meromorphic functions on (U,0)(U,0). Then, ff has multiple growth orders at any point of ϖ1(0)\varpi^{-1}(0), which is a consequence of the classical asymptotic analysis. (For example, see [28, §II.1].)

4.3.5. Coordinate change

Let UU be an open neighbourhood of 0 in {\mathbb{C}}. Let γ:U\gamma:U\longrightarrow{\mathbb{C}} be a holomorphic function such that γ(0)=0\gamma(0)=0 and zγ(0)=1\partial_{z}\gamma(0)=1. It induces an automorphism γ\gamma^{\ast} of ι(𝒪)Q\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q}.

For any zaz^{-a} (a>0)(a>0), we have the expansion

γ(za)=za(γ(z)/z)a=za+b>aαa,bzb.\gamma^{\ast}(z^{-a})=z^{-a}(\gamma^{\ast}(z)/z)^{-a}=z^{-a}+\sum_{b>-a}\alpha_{a,b}z^{b}.

We define

γ(za):=za+a<b<0αa,bzb.\gamma^{\ast}_{-}(z^{-a}):=z^{-a}+\sum_{-a<b<0}\alpha_{a,b}z^{b}.

It induces an injection γ:(Q)(Q)\gamma_{-}^{\ast}:\mathfrak{I}(Q)\longrightarrow\mathfrak{I}(Q). The following lemma is easy to see.

Lemma 4.25.

fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has a single growth order if and only if γ(f)\gamma^{\ast}(f) has a single growth order. In that case, we have 𝔞(γ(f),Q)=γ(𝔞(f,Q))\mathfrak{a}(\gamma^{\ast}(f),Q)=\gamma_{-}^{\ast}(\mathfrak{a}(f,Q)). As a result fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} has multiple growth orders if and only if γ(f)\gamma^{\ast}(f) has multiple growth orders. ∎

Let 𝔅~,Q\mathfrak{B}_{\widetilde{{\mathbb{C}}},Q} denote the set of fι(𝒪)Qf\in\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{Q} with multiple growth orders. It is independent of the choice of a coordinate zz by Lemma 4.25.

4.3.6. Sheaf of holomorphic functions with multiple growth orders

Let XX be any Riemann surface with a discrete subset DD. Let ϖX,D:X~DX\varpi_{X,D}:\widetilde{X}_{D}\longrightarrow X denote the oriented real blowing up. Let ιXD:XDX~D\iota_{X\setminus D}:X\setminus D\longrightarrow\widetilde{X}_{D} denote the inclusion.

Notation 4.26.

Let 𝔅X~DιXD,(𝒪XD)\mathfrak{B}_{\widetilde{X}_{D}}\subset\iota_{X\setminus D,\ast}(\mathcal{O}_{X\setminus D}) denote the subsheaf determined as follows.

  • 𝔅X~D|XD=𝒪XD\mathfrak{B}_{\widetilde{X}_{D}|X\setminus D}=\mathcal{O}_{X\setminus D}.

  • For any PDP\in D, we take a holomorphic embedding zP:XPz_{P}:X_{P}\longrightarrow{\mathbb{C}} of a neighbourhood XPX_{P} of PP in XX such that zP(P)=0z_{P}(P)=0. Note that it induces the isomorphisms z~P:ϖX,D1(P)ϖ1(0)\widetilde{z}_{P}:\varpi_{X,D}^{-1}(P)\simeq\varpi^{-1}(0) and z~P:ι(𝒪)z~P(Q)ιXD,(𝒪XD)Q\widetilde{z}_{P}^{\ast}:\iota_{\ast}(\mathcal{O}_{{\mathbb{C}}^{\ast}})_{\widetilde{z}_{P}(Q)}\simeq\iota_{X\setminus D,\ast}(\mathcal{O}_{X\setminus D})_{Q} for any QϖX,D1(P)Q\in\varpi_{X,D}^{-1}(P). Then, 𝔅X~D,Q=z~P1(𝔅~,z~P(Q))\mathfrak{B}_{\widetilde{X}_{D},Q}=\widetilde{z}_{P}^{-1}\bigl{(}\mathfrak{B}_{\widetilde{{\mathbb{C}}},\widetilde{z}_{P}(Q)}\bigr{)} holds.

If ff is a section of 𝔅X~D\mathfrak{B}_{\widetilde{X}_{D}} on an open subset VX~DV\subset\widetilde{X}_{D}, αf\alpha f (α)(\alpha\in{\mathbb{C}}^{\ast}) are also sections of 𝔅X~D\mathfrak{B}_{\widetilde{X}_{D}} on VV. But, even if fif_{i} (i=1,2)(i=1,2) are sections of 𝔅X~D\mathfrak{B}_{\widetilde{X}_{D}} on VV, f1+f2f_{1}+f_{2} is not necessarily a section of 𝔅X~D\mathfrak{B}_{\widetilde{X}_{D}} on VV. For example, we set f1=ez1f_{1}=e^{-z^{-1}} and f2=ez1+eez1f_{2}=-e^{-z^{-1}}+e^{-e^{z^{-1}}} around arg(z)=0\arg(z)=0. Then, fif_{i} are sections of 𝔙X~D\mathfrak{V}_{\widetilde{X}_{D}}, but f1+f2=eez1f_{1}+f_{2}=e^{-e^{z^{-1}}} is not.

4.3.7. Intervals with some properties

Let ff be a section of 𝔅~\mathfrak{B}_{\widetilde{{\mathbb{C}}}} on an open subset V~V\subset\widetilde{{\mathbb{C}}}.

Definition 4.27.

An open interval Iϖ1(0)VI\subset\varpi^{-1}(0)\cap V is called positive (resp. negative) with respect to ff if ff is simply positive (resp. negative) at each point of I𝒵(f)I\setminus\mathcal{Z}(f). It is called maximal if moreover there does not exist an interval I1ϖ1(0)VI_{1}\subset\varpi^{-1}(0)\cap V such that (i) I1I_{1} is positive (resp. negative) with respect to ff, (ii) II1I\subsetneq I_{1}.

Definition 4.28.

An open interval Iϖ1(0)VI\subset\varpi^{-1}(0)\cap V is called neutral with respect to ff if ff is neutral at each point of I𝒵(f)I\setminus\mathcal{Z}(f). It is called maximal if moreover there does not exist an interval I1ϖ1(0)VI_{1}\subset\varpi^{-1}(0)\cap V such that (i) I1I_{1} is neutral with respect to ff, (ii) II1I\subsetneq I_{1}.

Definition 4.29.

An open interval Iϖ1(0)VI\subset\varpi^{-1}(0)\cap V is called non-positive with respect to ff if ff is neutral or simply negative at each point of I𝒵(f)I\setminus\mathcal{Z}(f). It is called maximal if moreover there does not exist an interval I1ϖ1(0)VI_{1}\subset\varpi^{-1}(0)\cap V such that (i) I1I_{1} is non-positive with respect to ff, (ii) II1I\subsetneq I_{1}.

In Definitions 4.274.29, if I¯\overline{I} is contained in VV and if II is maximal, then I𝒵(f)\partial I\subset\mathcal{Z}(f).

Lemma 4.30.

If an interval II is non-positive with respect to ff, then II is either negative or neutral with respect to ff. Moreover, if II is neutral with respect to ff, then I𝒵(f)=I\cap\mathcal{Z}(f)=\emptyset.

Proof   Let II be a non-positive interval. Let I𝒵(f)=IiI\setminus\mathcal{Z}(f)=\coprod I_{i} be the decomposition into the connected components. The numbering IiI_{i} is given in the counter-clockwise way. If ff is negative (neutral) at a point QIiQ\in I_{i}, then ff is negative at any point of IiI_{i}.

Let QI𝒵(f)Q\in I\cap\mathcal{Z}(f) such that Q=I¯iI¯i+1Q=\overline{I}_{i}\cap\overline{I}_{i+1}. There exists a reduced expression f=𝔞f𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}}f_{\mathfrak{a}} at QQ as in Lemma 4.22. Take QiIiQ_{i}\in I_{i} and Qi+1Ii+1Q_{i+1}\in I_{i+1}. Suppose that ff is negative at QiQ_{i} and neutral at Qi+1Q_{i+1}, and we shall deduce a contradiction. Note that 𝔞(f,Qi)=max(,Qi)=:𝔟\mathfrak{a}(f,Q_{i})=\max(\mathcal{I},\prec_{Q_{i}})=:\mathfrak{b} and 𝔞(f,Qi+1)=max(,Qi+1)=0\mathfrak{a}(f,Q_{i+1})=\max(\mathcal{I},\prec_{Q_{i+1}})=0. We obtain |𝔟|1Re(𝔟)(Q)=0|\mathfrak{b}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{b})(Q)=0. If Qi+1Q_{i+1} is sufficiently close to QQ, we obtain |𝔟|1Re(𝔟)(Qi+1)>0|\mathfrak{b}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{b})(Q_{i+1})>0, i.e., 0Qi+1𝔟0\prec_{Q_{i+1}}\mathfrak{b}. It contradicts 𝔟Qi+10\mathfrak{b}\prec_{Q_{i+1}}0. Therefore, if ff is negative at QiQ_{i}, then ff is negative at Qi+1Q_{i+1}. If ff is neutral at QiQ_{i}, then ff is neutral at Qi+1Q_{i+1}, and moreover we obtain ={0}\mathcal{I}=\{0\}. Then, we easily obtain the claim of the lemma. ∎

Lemma 4.31.

Suppose that II is negative with respect to ff such that I𝒵(f)I\cap\mathcal{Z}(f) consists of one point QQ. Choose Q1Q_{1} and Q2Q_{2} from the two connected components of I{Q}I\setminus\{Q\}. Let f=𝔞f𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}}f_{\mathfrak{a}} be a reduced expression at QQ. Then, the following holds.

  • deg𝔞(f,Q1)=deg𝔞(f,Q2)\deg\mathfrak{a}(f,Q_{1})=\deg\mathfrak{a}(f,Q_{2}).

  • For any 𝔞\mathfrak{a}\in\mathcal{I}, we obtain deg(𝔞)=deg(𝔞(f,Qi))\deg(\mathfrak{a})=\deg(\mathfrak{a}(f,Q_{i})). Moreover, |𝔞|1Re(𝔞)(Q)<0|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a})(Q)<0.

Proof   Note that 𝔞(f,Qi)\mathfrak{a}(f,Q_{i})\in\mathcal{I}. Let us prove

deg𝔞(f,Q1)deg𝔞(f,Q2).\deg\mathfrak{a}(f,Q_{1})\geq\deg\mathfrak{a}(f,Q_{2}). (50)

Suppose deg𝔞(f,Q1)<deg𝔞(f,Q2)\deg\mathfrak{a}(f,Q_{1})<\deg\mathfrak{a}(f,Q_{2}), and we shall deduce a contradiction. If |𝔞(f,Q2)|1Re(𝔞(f,Q2))(Q)<0|\mathfrak{a}(f,Q_{2})|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q_{2}))(Q)<0, we obtain 𝔞(f,Q2)Q𝔞(f,Q1)\mathfrak{a}(f,Q_{2})\prec_{Q}\mathfrak{a}(f,Q_{1}), and hence 𝔞(f,Q2)Q2𝔞(f,Q1)\mathfrak{a}(f,Q_{2})\prec_{Q_{2}}\mathfrak{a}(f,Q_{1}) holds, which contradicts 𝔞(f,Q2)=max(,Q2)\mathfrak{a}(f,Q_{2})=\max(\mathcal{I},\prec_{Q_{2}}). If |𝔞(f,Q2)|1Re(𝔞(f,Q2))(Q)0|\mathfrak{a}(f,Q_{2})|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q_{2}))(Q)\geq 0, because 𝔞(f,Q1)Q2𝔞(f,Q2)\mathfrak{a}(f,Q_{1})\prec_{Q_{2}}\mathfrak{a}(f,Q_{2}), we obtain |𝔞(f,Q2)|1Re(𝔞(f,Q2))(Q2)>0|\mathfrak{a}(f,Q_{2})|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q_{2}))(Q_{2})>0, which contradicts that ff is negative at Q2Q_{2}. Thus, we obtain (50). Similarly, we obtain deg𝔞(f,Q1)deg𝔞(f,Q2)\deg\mathfrak{a}(f,Q_{1})\leq\deg\mathfrak{a}(f,Q_{2}), and hence deg𝔞(f,Q1)=deg𝔞(f,Q2)\deg\mathfrak{a}(f,Q_{1})=\deg\mathfrak{a}(f,Q_{2}).

Take 𝔞\mathfrak{a}\in\mathcal{I}. Let us prove deg(𝔞)=deg(𝔞(f,Qi))\deg(\mathfrak{a})=\deg(\mathfrak{a}(f,Q_{i})). Suppose deg𝔞>deg𝔞(f,Qi)\deg\mathfrak{a}>\deg\mathfrak{a}(f,Q_{i}). If |𝔞|1Re(𝔞)(Q)<0|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a})(Q)<0, we obtain 𝔞Q𝔞(f,Qi)\mathfrak{a}\prec_{Q}\mathfrak{a}(f,Q_{i}), which contradicts that the expression f=𝔞f𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}}f_{\mathfrak{a}} is reduced. If |𝔞|1Re(𝔞)(Q)0|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a})(Q)\geq 0, then either |𝔞|1Re(𝔞)(Q1)>0|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a})(Q_{1})>0 or |𝔞|1Re(𝔞)(Q2)>0|\mathfrak{a}|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a})(Q_{2})>0 holds, which contradicts that ff is negative at QiQ_{i}. Suppose deg𝔞<deg𝔞(f,Qi)\deg\mathfrak{a}<\deg\mathfrak{a}(f,Q_{i}). Because |𝔞(f,Qi)|1Re(𝔞(f,Qi))(Qi)<0|\mathfrak{a}(f,Q_{i})|^{-1}\mathop{\rm Re}\nolimits(\mathfrak{a}(f,Q_{i}))(Q_{i})<0, we obtain 𝔞(f,Qi)Qi𝔞\mathfrak{a}(f,Q_{i})\prec_{Q_{i}}\mathfrak{a}, which contradicts 𝔞(f,Qi)=max(,Qi)\mathfrak{a}(f,Q_{i})=\max(\mathcal{I},\prec_{Q_{i}}). In all, we obtain deg(𝔞)=deg(𝔞(f,Qi))\deg(\mathfrak{a})=\deg(\mathfrak{a}(f,Q_{i})). ∎

Recall that we identify ~0×S1\widetilde{{\mathbb{C}}}\simeq{\mathbb{R}}_{\geq 0}\times S^{1} by the polar coordinate system. For any a<ba<b, we set ]a,b[={θ|a<θ<b}]a,b[=\{\theta\in{\mathbb{R}}\,|\,a<\theta<b\}. If ba<2πb-a<2\pi, we may naturally regard it as an interval in S1=/2πS^{1}={\mathbb{R}}/2\pi{\mathbb{Z}}.

Definition 4.32.

An open interval Iϖ1(0)VI\subset\varpi^{-1}(0)\cap V is called special with respect to ff if the following holds.

  • The length of II is π/ρ\pi/\rho for ρ>1/2\rho>1/2. Namely, I={0}×]θ1,θ1+π/ρ[I=\{0\}\times]\theta_{1},\theta_{1}+\pi/\rho[ for some θ1\theta_{1}.

  • There exists α{0}\alpha\in{\mathbb{C}}\setminus\{0\} such that (i) deg(𝔞(f,Q)αzρ)<deg(𝔞(f,Q))=ρ\deg(\mathfrak{a}(f,Q)-\alpha z^{-\rho})<\deg(\mathfrak{a}(f,Q))=\rho for any QI𝒵(f)Q\in I\setminus\mathcal{Z}(f), (ii) Re(αeρ1θ)<0\mathop{\rm Re}\nolimits(\alpha e^{-\rho\sqrt{-1}\theta})<0 on II, which particularly implies that

    Re(αeρ1θ1)=Re(αeρ1(θ1+π/ρ))=0.\mathop{\rm Re}\nolimits(\alpha e^{-\rho\sqrt{-1}\theta_{1}})=\mathop{\rm Re}\nolimits(\alpha e^{-\rho\sqrt{-1}(\theta_{1}+\pi/\rho)})=0.
Lemma 4.33.

Let II be an interval which is negative with respect to ff. Note that ρ:=deg(𝔞(f,Q))\rho:=\deg(\mathfrak{a}(f,Q)) (QI𝒵(f))(Q\in I\setminus\mathcal{Z}(f)) is well defined. Then, either one of the following holds.

  • II is special with respect to ff.

  • The length of II is strictly smaller than π/ρ\pi/\rho.

Proof   There exists an interval I~={ψ0θψ1}\widetilde{I}=\{\psi_{0}\leq\theta\leq\psi_{1}\}\subset{\mathbb{R}} such that the map θe1θ\theta\longmapsto e^{\sqrt{-1}\theta} induces a diffeomorphism I~I\widetilde{I}\simeq I. Let ψ0<θ1<<θ<ψ1\psi_{0}<\theta_{1}<\cdots<\theta_{\ell}<\psi_{1} be determined by {(0,e1θi)}=I𝒵(f)\{(0,e^{\sqrt{-1}\theta_{i}})\}=I\cap\mathcal{Z}(f). We set θ0:=ψ0\theta_{0}:=\psi_{0} and θ+1:=ψ1\theta_{\ell+1}:=\psi_{1}. We set I~i={0}×]θi1,θi[I~\widetilde{I}_{i}=\{0\}\times]\theta_{i-1},\theta_{i}[\subset\widetilde{I} (i=1,,+1)(i=1,\ldots,\ell+1). We obtain the corresponding subsets IiII_{i}\subset I. Choose QiIiQ_{i}\in I_{i}. There exist αi\alpha_{i}\in{\mathbb{C}}^{\ast} such that deg(𝔞(f,Qi)αizρ)<deg(𝔞(f,Qi))\deg(\mathfrak{a}(f,Q_{i})-\alpha_{i}z^{-\rho})<\deg(\mathfrak{a}(f,Q_{i})). There exist intervals Ji=]φi,φi+π/ρ[J_{i}=]\varphi_{i},\varphi_{i}+\pi/\rho[ such that (i) φiθi1<θiφi+π/ρ\varphi_{i}\leq\theta_{i-1}<\theta_{i}\leq\varphi_{i}+\pi/\rho, (ii) Re(αieρ1θ)<0\mathop{\rm Re}\nolimits(\alpha_{i}e^{-\rho\sqrt{-1}\theta})<0 on JiJ_{i}.

For 1i1\leq i\leq\ell, we obtain Re(αieρ1θi)=Re(αi+1eρ1θi)<0\mathop{\rm Re}\nolimits(\alpha_{i}e^{-\rho\sqrt{-1}\theta_{i}})=\mathop{\rm Re}\nolimits(\alpha_{i+1}e^{-\rho\sqrt{-1}\theta_{i}})<0. We also obtain

Re(αieρ1(θi+ϵ))<Re(αi+1eρ1(θi+ϵ)),\mathop{\rm Re}\nolimits(\alpha_{i}e^{-\rho\sqrt{-1}(\theta_{i}+\epsilon)})<\mathop{\rm Re}\nolimits(\alpha_{i+1}e^{-\rho\sqrt{-1}(\theta_{i}+\epsilon)}),
Re(αieρ1(θiϵ))>Re(αi+1eρ1(θiϵ))\mathop{\rm Re}\nolimits(\alpha_{i}e^{-\rho\sqrt{-1}(\theta_{i}-\epsilon)})>\mathop{\rm Re}\nolimits(\alpha_{i+1}e^{-\rho\sqrt{-1}(\theta_{i}-\epsilon)})

for any sufficiently small positive number ϵ\epsilon. Therefore, we obtain φi+1φi\varphi_{i+1}\leq\varphi_{i}. Moreover, if φi+1=φi\varphi_{i+1}=\varphi_{i}, there exists a>0a>0 such that αi+1=aαi\alpha_{i+1}=a\alpha_{i}. Because Re(αieρ1θi)=Re(αi+1eρ1θi)<0\mathop{\rm Re}\nolimits(\alpha_{i}e^{-\rho\sqrt{-1}\theta_{i}})=\mathop{\rm Re}\nolimits(\alpha_{i+1}e^{-\rho\sqrt{-1}\theta_{i}})<0, we obtain αi+1=αi\alpha_{i+1}=\alpha_{i}. Therefore, we obtain

θ+1φ+1+π/ρφ1+π/ρθ0+π/ρ.\theta_{\ell+1}\leq\varphi_{\ell+1}+\pi/\rho\leq\cdots\leq\varphi_{1}+\pi/\rho\leq\theta_{0}+\pi/\rho.

Namely, we obtain θ+1θ0π/ρ\theta_{\ell+1}-\theta_{0}\leq\pi/\rho. If θ+1=θ0+π/ρ\theta_{\ell+1}=\theta_{0}+\pi/\rho, we obtain α1=α2==α+1\alpha_{1}=\alpha_{2}=\cdots=\alpha_{\ell+1}. ∎

We also obtain the following lemma.

Lemma 4.34.

Suppose that VV is a neighbourhood of ϖ1(0)\varpi^{-1}(0), and that ff is non-positive at each point of ϖ1(0)𝒵(f)\varpi^{-1}(0)\setminus\mathcal{Z}(f). Then, ff is meromorphic at 0.

Proof   If ff is neutral at a point of ϖ1(0)𝒵(f)\varpi^{-1}(0)\setminus\mathcal{Z}(f), we obtain that ff is neutral at any point of ϖ1(0)𝒵(f)\varpi^{-1}(0)\setminus\mathcal{Z}(f) according to Lemma 4.30, and hence we obtain that ff is meromorphic. Suppose that ff is simply negative at any point of ϖ1(0)𝒵(f)\varpi^{-1}(0)\setminus\mathcal{Z}(f). By Lemma 4.31, there exist ϵ>0\epsilon>0 and ρ>0\rho>0 such that |f|=O(eϵ|z|ρ)|f|=O\bigl{(}e^{-\epsilon|z|^{-\rho}}\bigr{)} on Vϖ1(0)V\setminus\varpi^{-1}(0), which implies f=0f=0. ∎

4.3.8. Estimates in a single case

For any R>0R>0 and 0<L<π0<L<\pi, we set W(R,L):={w||w|>R,|arg(w)|<L}W(R,L):=\bigl{\{}w\in{\mathbb{C}}\,\big{|}\,|w|>R,\,|\arg(w)|<L\bigr{\}}. Let ~11\widetilde{\mathbb{P}}^{1}_{\infty}\longrightarrow\mathbb{P}^{1} be the oriented real blowing up at \infty. We regard W(R,L)W(R,L) as an open subset of ~1\widetilde{\mathbb{P}}^{1}_{\infty}. Let W¯(R,L)\overline{W}(R,L) denote the closure of W(R,L)W(R,L) in ~1\widetilde{\mathbb{P}}^{1}_{\infty}. Let Q0W¯(R,L)Q_{0}\in\overline{W}(R,L) denote the point corresponding to ++\infty, i.e., Q0Q_{0} is the limit of any sequence of positive numbers tit_{i} in ~1\widetilde{\mathbb{P}}^{1}_{\infty} with tit_{i}\to\infty. Let Φβ,b:W(R,L)\Phi_{\beta,b}:W(R,L)\longrightarrow{\mathbb{C}}^{\ast} be given by Φβ,b(w)=βwb\Phi_{\beta,b}(w)=\beta w^{-b} for a positive number bb and a non-zero complex number β\beta. It induces Φ~β,b:W¯(R,L)~\widetilde{\Phi}_{\beta,b}:\overline{W}(R,L)\longrightarrow\widetilde{{\mathbb{C}}}.

Let ff be a section of 𝔅~\mathfrak{B}_{\widetilde{{\mathbb{C}}}} on an open subset V~V\subset\widetilde{{\mathbb{C}}}. Suppose that ff has a single growth order at QVϖ1(0)Q\in V\cap\varpi^{-1}(0). Let 0<L<L~<π0<L<\widetilde{L}<\pi. By choosing β\beta\in{\mathbb{C}}^{\ast}, b>0b>0 and R>0R>0 appropriately, we obtain Φ~β,b:W(R,L~)~\widetilde{\Phi}_{\beta,b}:W(R,\widetilde{L})\to\widetilde{{\mathbb{C}}} such that Φ~β,b(Q0)=Q\widetilde{\Phi}_{\beta,b}(Q_{0})=Q and Φ~β,b(W¯(R,L~))V\widetilde{\Phi}_{\beta,b}(\overline{W}(R,\widetilde{L}))\subset V. We obtain the holomorphic function Φb,β(f)\Phi_{b,\beta}^{\ast}(f) on W(R,L~)W(R,\widetilde{L}) which induces a section of 𝔅~1\mathfrak{B}_{\widetilde{\mathbb{P}}^{1}_{\infty}} on W¯(R,L~)\overline{W}(R,\widetilde{L}). If bb is sufficiently small and RR is sufficiently large,

e𝔞(f,Q)za(f,Q)(logz)j(f,Q)fe^{-\mathfrak{a}(f,Q)}z^{-a(f,Q)}(\log z)^{-j(f,Q)}f

is regularly bounded on Φ~β,b(W¯(R,L~))ϖ1(0)\widetilde{\Phi}_{\beta,b}(\overline{W}(R,\widetilde{L}))\setminus\varpi^{-1}(0).

Lemma 4.35.

For any 0<L2<L1<L0<L_{2}<L_{1}<L, there exist R2>R1>RR_{2}>R_{1}>R and subsets 𝒞1𝒞2W(R,L)\mathcal{C}_{1}\subset\mathcal{C}_{2}\subset W(R,L) such that the following holds.

  • There exists C>0C>0 such that

    |Φβ,b(f)|CeReΦβ,b𝔞(f,Q)|w|ba(f,Q)(log|w|)j(f,Q)\bigl{|}\Phi_{\beta,b}^{\ast}(f)\bigr{|}\geq Ce^{\mathop{\rm Re}\nolimits\Phi_{\beta,b}^{\ast}\mathfrak{a}(f,Q)}|w|^{ba(f,Q)}(\log|w|)^{j(f,Q)}

    on W(R,L)𝒞1W(R,L)\setminus\mathcal{C}_{1}.

  • Let 𝒟\mathcal{D} be any connected component of 𝒞2\mathcal{C}_{2} such that

    𝒟W(R2,L2).\mathcal{D}\cap W(R_{2},L_{2})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in W(R1,L1)W(R_{1},L_{1}).

  • For any w0W(R1,L1)𝒞2w_{0}\in W(R_{1},L_{1})\setminus\mathcal{C}_{2}, we obtain {|ww0|<1}W(R,L)𝒞1\{|w-w_{0}|<1\}\subset W(R,L)\setminus\mathcal{C}_{1}.

Proof   It is enough to prove the case where 𝔞(f,Q)=0\mathfrak{a}(f,Q)=0, a(f,Q)=0a(f,Q)=0 and j(f,Q)=0j(f,Q)=0, i.e., ff is regularly bounded at QQ.

For any B1B_{1}\in{\mathbb{R}} and B2>0B_{2}>0, we set W~(B1,B2):={ζ|Re(ζ)>B1,|Im(ζ)|<B2}\widetilde{W}(B_{1},B_{2}):=\{\zeta\in{\mathbb{C}}\,|\,\mathop{\rm Re}\nolimits(\zeta)>B_{1},\,|\mathop{\rm Im}\nolimits(\zeta)|<B_{2}\}. Let Φ1:\Phi_{1}:{\mathbb{C}}\longrightarrow{\mathbb{C}} be given by Φ1(ζ)=eζ\Phi_{1}(\zeta)=e^{\zeta}. It induces W~(B1,B2)W(eB1,B2)\widetilde{W}(B_{1},B_{2})\simeq W(e^{B_{1}},B_{2}) for any B1,B2>0B_{1},B_{2}>0. We set f~:=Φ1Φβ,b(f)\widetilde{f}:=\Phi_{1}^{\ast}\Phi_{\beta,b}^{\ast}(f) on W(logR,L~)W(\log R,\widetilde{L}). Let Z(f~)Z(\widetilde{f}) denote the zero set of f~\widetilde{f}. By Proposition 4.9, there exists N>0N>0 such that the following holds for any x1>logRx_{1}>\log R:

#(Z(f~){ζ|x1<Re(ζ)<x1+L,|Im(ζ)|<L})<N.\#\Bigl{(}Z(\widetilde{f})\cap\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,x_{1}<\mathop{\rm Re}\nolimits(\zeta)<x_{1}+L,\,\,|\mathop{\rm Im}\nolimits(\zeta)|<L\bigr{\}}\Bigr{)}<N.

Take

0<δ<1103Nmin{|LL1|,|L1L2|,|L2|}.0<\delta<\frac{1}{10^{3}N}\min\bigl{\{}|L-L_{1}|,|L_{1}-L_{2}|,|L_{2}|\bigr{\}}.

We set

1:=ζ1Z(f~){|ζζ1|<δ}W~(logR,L),\mathcal{B}_{1}:=\bigcup_{\zeta_{1}\in Z(\widetilde{f})}\{|\zeta-\zeta_{1}|<\delta\}\cap\widetilde{W}(\log R,L),
2:=ζ1Z(f~){|ζζ1|<3δ}W~(logR,L).\mathcal{B}_{2}:=\bigcup_{\zeta_{1}\in Z(\widetilde{f})}\{|\zeta-\zeta_{1}|<3\delta\}\cap\widetilde{W}(\log R,L).

There exists C>0C>0 such that |f~|>C|\widetilde{f}|>C on W~(logR,L)1\widetilde{W}(\log R,L)\setminus\mathcal{B}_{1} by Corollary 4.11.

We set 𝒞i=Φ1(i)W(R,L)\mathcal{C}_{i}=\Phi_{1}(\mathcal{B}_{i})\subset W(R,L). Then, the first condition is satisfied. Take R1>max{ReL,10δ1}R_{1}>\max\{Re^{L},10\delta^{-1}\}. For any ζ1,ζ2W~(logR)\zeta_{1},\zeta_{2}\in\widetilde{W}(\log R), we have

|ζ1ζ2||eζ1eζ2|max{eRe(ζ1),eRe(ζ2)}.|\zeta_{1}-\zeta_{2}|\leq|e^{\zeta_{1}}-e^{\zeta_{2}}|\max\{e^{-\mathop{\rm Re}\nolimits(\zeta_{1})},e^{-\mathop{\rm Re}\nolimits(\zeta_{2})}\}.

Then, the third condition is satisfied. Take R2>R1eLR_{2}>R_{1}e^{L}. Let 𝒟~\widetilde{\mathcal{D}} be a connected component of 2\mathcal{B}_{2} such that 𝒟~W~(logR2,L2)\widetilde{\mathcal{D}}\cap\widetilde{W}(\log R_{2},L_{2})\neq\emptyset. Then, there exists a subset Z𝒟~(f~)Z(f~)Z_{\widetilde{\mathcal{D}}}(\widetilde{f})\subset Z(\widetilde{f}) such that 𝒟~\widetilde{\mathcal{D}} is the union of {|ζζj|<3δ}\{|\zeta-\zeta_{j}|<3\delta\} (ζjZ𝒟~(f~))(\zeta_{j}\in Z_{\widetilde{\mathcal{D}}}(\widetilde{f})). Let us observe that #Z𝒟~(f~)N\#Z_{\widetilde{\mathcal{D}}}(\widetilde{f})\leq N. Take ζ0Z𝒟~(f~)\zeta_{0}\in Z_{\widetilde{\mathcal{D}}}(\widetilde{f}). If #Z𝒟~(f~)>N\#Z_{\widetilde{\mathcal{D}}}(\widetilde{f})>N, there exists ηZ𝒟~(f~)\eta\in Z_{\widetilde{\mathcal{D}}}(\widetilde{f}) such that |Re(ηζ0)|>L/2|\mathop{\rm Re}\nolimits(\eta-\zeta_{0})|>L/2. There exists a sequence ζ1,,ζm,ζm+1=ηZ𝒟~(f~)\zeta_{1},\ldots,\zeta_{m},\zeta_{m+1}=\eta\in Z_{\widetilde{\mathcal{D}}}(\widetilde{f}) such that d(ζi,ζi+1)<3δd(\zeta_{i},\zeta_{i+1})<3\delta for i=0,,mi=0,\ldots,m. There exists i0i_{0} such that |Re(ζ0ζi)|L/2|\mathop{\rm Re}\nolimits(\zeta_{0}-\zeta_{i})|\leq L/2 for i<i0i<i_{0} and |Re(ζ0ζi0)|>L/2|\mathop{\rm Re}\nolimits(\zeta_{0}-\zeta_{i_{0}})|>L/2. Note that i0Ni_{0}\leq N. Because of the choice of δ\delta, we obtain d(ζ0,ζi0)<L/2d(\zeta_{0},\zeta_{i_{0}})<L/2. Thus, we arrive at a contradiction, and hence we obtain #Z𝒟~(f~)N\#Z_{\widetilde{\mathcal{D}}}(\widetilde{f})\leq N. Then, we obtain that 𝒟~\widetilde{\mathcal{D}} is relatively compact in W~(logR1,L1)\widetilde{W}(\log R_{1},L_{1}), and hence the second condition is satisfied. ∎

We set g=w+αwcg=w+\alpha w^{c} on W(R,L~)W(R,\widetilde{L}) for α{0}\alpha\in{\mathbb{C}}\setminus\{0\} and 0<c<10<c<1. For any B(1)B^{(1)}\in{\mathbb{R}} and 0<L(1)<L0<L^{(1)}<L, there exists R(1)RR^{(1)}\geq R such that

S(g,R(1),B(1),L(1))={wW(R,L~)||g(w)|>R(1),Im(g(w))B(1),arg(g(w))<L(1)}W(R,L).S(g,R^{(1)},B^{(1)},L^{(1)})=\\ \bigl{\{}w\in W(R,\widetilde{L})\,\big{|}\,|g(w)|>R^{(1)},\,\,\mathop{\rm Im}\nolimits(g(w))\geq B^{(1)},\,\,\arg(g(w))<L^{(1)}\bigr{\}}\\ \subset W(R,L). (51)

Let Z(Φb,β(f))Z(\Phi_{b,\beta}^{\ast}(f)) denote the zero set of Φb,β(f)\Phi_{b,\beta}^{\ast}(f).

Lemma 4.36.

There exists N>0N>0 such that the following holds for any M>R(1)eLM>R^{(1)}e^{L}:

#(Z(Φb,β(f)){wS(g,R(1),B(1),L(1))|M<|g(w)|<MeL})<N.\#\Bigl{(}Z(\Phi_{b,\beta}^{\ast}(f))\cap\bigl{\{}w\in S(g,R^{(1)},B^{(1)},L^{(1)})\,\big{|}\,M<|g(w)|<Me^{L}\bigr{\}}\Bigr{)}<N.

Proof   There exist M0>RM_{0}>R and ϵ>0\epsilon>0 such that the following holds for any M>M0M>M_{0}:

{wS(g,R(1),B(1),L(1))|M<|g(w)|<MeL}{wW(R,L)|Meϵ<|w|<MeL+ϵ}.\bigl{\{}w\in S(g,R^{(1)},B^{(1)},L^{(1)})\,\big{|}\,M<|g(w)|<Me^{L}\bigr{\}}\subset\\ \bigl{\{}w\in W(R,L)\,\big{|}\,Me^{-\epsilon}<|w|<Me^{L+\epsilon}\bigr{\}}. (52)

Then, the claim follows from Proposition 4.9. ∎

Lemma 4.37.

For any ρ>0\rho>0 and κ>0\kappa>0, there exist R0(1)>0R_{0}^{(1)}>0, C1>0C_{1}>0 and κ1>0\kappa_{1}>0 such that

eReΦβ,b𝔞(f,Q)|w|ba(f,Q)(log|w|)j(f,Q)|Φβ,b(f)(w)|C1eκ1|w|ρe^{-\mathop{\rm Re}\nolimits\Phi_{\beta,b}^{\ast}\mathfrak{a}(f,Q)}|w|^{-ba(f,Q)}(\log|w|)^{-j(f,Q)}\bigl{|}\Phi_{\beta,b}^{\ast}(f)(w)\bigr{|}\geq C_{1}e^{-\kappa_{1}|w|^{\rho}}

for any wS(g,R0(1),B(1),L(1))w\in S(g,R_{0}^{(1)},B^{(1)},L^{(1)}) such that d(w,Z(Φβ,b(f)))12eκ|w|ρd(w,Z(\Phi_{\beta,b}^{\ast}(f)))\geq\frac{1}{2}e^{-\kappa|w|^{\rho}}.

Proof   We have only to study the case where 𝔞(f,Q)=0\mathfrak{a}(f,Q)=0, a(f,Q)=0a(f,Q)=0 and j(f,Q)=0j(f,Q)=0. We use the notation in the proof of Lemma 4.35.

For ζ1,ζ2\zeta_{1},\zeta_{2}\in{\mathbb{C}} with |ζ1ζ2|1|\zeta_{1}-\zeta_{2}|\leq 1, we have |eζ1eζ2|eRe(ζ1)+1|ζ1ζ2||e^{\zeta_{1}}-e^{\zeta_{2}}|\leq e^{\mathop{\rm Re}\nolimits(\zeta_{1})+1}|\zeta_{1}-\zeta_{2}|. Hence, there exists C2>0C_{2}>0 such that the following holds for any wS(g,R(1),B(1),L(1))w\in S(g,R^{(1)},B^{(1)},L^{(1)}) satisfying d(w,Z(Φβ,b(f)))12eκ|w|ρd(w,Z(\Phi_{\beta,b}^{\ast}(f)))\geq\frac{1}{2}e^{-\kappa|w|^{\rho}}.

  • Take ζW~(logR,L)\zeta\in\widetilde{W}(\log R,L) such that eζ=we^{\zeta}=w. Then, we obtain d(ζ,Z(f~))>C2|w|1eκ|w|ρd(\zeta,Z(\widetilde{f}))>C_{2}|w|^{-1}e^{-\kappa|w|^{\rho}}.

By Corollary 4.11, there exist constants Ci>0C_{i}>0 (i=3,4,5,6)(i=3,4,5,6) such that the following holds for any wS(g,R(1),B(1),L(1))w\in S(g,R^{(1)},B^{(1)},L^{(1)}) satisfying d(w,Z(Φβ,b(f)))12eκ|w|ρd(w,Z(\Phi_{\beta,b}^{\ast}(f)))\geq\frac{1}{2}e^{-\kappa|w|^{\rho}}:

|Φβ,b(f)(w)|C3(|w|1eκ|w|ρ)C4C5eC6κ|w|ρ.\bigl{|}\Phi_{\beta,b}^{\ast}(f)(w)\bigr{|}\geq C_{3}(|w|^{-1}e^{-\kappa|w|^{\rho}})^{C_{4}}\geq C_{5}e^{-C_{6}\kappa|w|^{\rho}}.

Thus, we obtain the claim of the lemma. ∎

Lemma 4.38.

For any B1(1)>B(1)B^{(1)}_{1}>B^{(1)}, 0<L1(1)<L(1)0<L^{(1)}_{1}<L^{(1)} and ρ>0\rho>0, there exist R1(1)>R(1)R^{(1)}_{1}>R^{(1)}, and a subset 𝒞S(g,R(1),B(1),L(1))\mathcal{C}\subset S(g,R^{(1)},B^{(1)},L^{(1)}) such that the following holds:

  • There exist Ci>0C_{i}>0 (i=1,2)(i=1,2) such that

    eRe(Φβ,b𝔞(f,Q))|w|ba(f,Q)(log|w|)j(f,Q)|Φβ,b(f)(w)|C1exp(C2|w|ρ)e^{-\mathop{\rm Re}\nolimits(\Phi_{\beta,b}^{\ast}\mathfrak{a}(f,Q))}|w|^{-ba(f,Q)}(\log|w|)^{-j(f,Q)}\bigl{|}\Phi_{\beta,b}^{\ast}(f)(w)\bigr{|}\geq\\ C_{1}\exp(-C_{2}|w|^{\rho}) (53)

    on S(g,R(1),B(1),L(1))𝒞S(g,R^{(1)},B^{(1)},L^{(1)})\setminus\mathcal{C}.

  • Let 𝒟\mathcal{D} be a connected component of 𝒞\mathcal{C} such that

    𝒟S(g,R1(1),B1(1),L1(1)).\mathcal{D}\cap S(g,R_{1}^{(1)},B^{(1)}_{1},L^{(1)}_{1})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in S(g,R(1),B(1),L(1))S(g,R^{(1)},B^{(1)},L^{(1)}).

  • There exist an increasing sequence of positive numbers TiT_{i} such that (i) limTi=\lim T_{i}=\infty, (ii) 𝒞{|g|=Ti}S(g,R(1),B(1),L(1))=\mathcal{C}\cap\{|g|=T_{i}\}\cap S(g,R^{(1)},B^{(1)},L^{(1)})=\emptyset.

Proof   We have only to study the case where 𝔞(f,Q)=0\mathfrak{a}(f,Q)=0, a(f,Q)=0a(f,Q)=0 and j(f,Q)=0j(f,Q)=0. Let NN be as in Lemma 4.36. Take κ>0\kappa>0 satisfying

103Neκ(R(1))ρ<min{|B1(1)B(1)|,eR(1)sin((L(1)L1(1))/2),eR(1)sin(L1(1)/2),R(1)(eL1)}.10^{3}Ne^{-\kappa(R^{(1)})^{\rho}}<\\ \min\Bigl{\{}|B^{(1)}_{1}-B^{(1)}|,e^{R^{(1)}}\!\!\!\sin\bigl{(}(L^{(1)}-L_{1}^{(1)})/2\bigr{)},e^{R^{(1)}}\!\!\!\sin(L_{1}^{(1)}/2),R^{(1)}(e^{L}-1)\Bigr{\}}. (54)

We set

𝒞:=w1Z(Φβ,b(f)){|ww1|<eκ|w1|ρ}S(g,R(1),B(1),L(1)).\mathcal{C}:=\bigcup_{w_{1}\in Z(\Phi_{\beta,b}^{\ast}(f))}\bigl{\{}|w-w_{1}|<e^{-\kappa|w_{1}|^{\rho}}\bigr{\}}\cap S(g,R^{(1)},B^{(1)},L^{(1)}).

For 0<ν<10<\nu<1 and for w,aw,a\in{\mathbb{C}} with |w|>1|w|>1 and |a|<ν|a|<\nu, we have

||w+a|ρ|w|ρ|ρν(|w|+ν)ρ1+ρν(|w|ν)ρ1.\bigl{|}|w+a|^{\rho}-|w|^{\rho}\bigr{|}\leq\rho\nu\left(|w|+\nu\right)^{\rho-1}+\rho\nu\left(|w|-\nu\right)^{\rho-1}.

Let R0(1)R(1)R_{0}^{(1)}\geq R^{(1)} be as in Lemma 4.37. There exists R2(1)>R0(1)eL+1R_{2}^{(1)}>R_{0}^{(1)}e^{L}+1 such that the following holds for any ww\in{\mathbb{C}} with |w|>R1(1)|w|>R_{1}^{(1)}:

12exp(κρeκ|w|ρ((|w|+1)ρ1+(|w|1)ρ1+2|w|ρ1))3/4.\frac{1}{2}\exp\Bigl{(}\kappa\rho e^{-\kappa|w|^{\rho}}\bigl{(}(|w|+1)^{\rho-1}+(|w|-1)^{\rho-1}+2|w|^{\rho-1}\bigr{)}\Bigr{)}\leq 3/4.

For w2,w3w_{2},w_{3}\in{\mathbb{C}} satisfying |wi|>R2(1)|w_{i}|>R_{2}^{(1)} satisfying |w2w3|12eκ|w2|ρ|w_{2}-w_{3}|\leq\frac{1}{2}e^{-\kappa|w_{2}|^{\rho}}, we obtain |w2w3|(3/4)eκ|w3|ρ|w_{2}-w_{3}|\leq(3/4)e^{-\kappa|w_{3}|^{\rho}}. Hence, by Lemma 4.37, there exist Ci>0C^{\prime}_{i}>0 (i=1,2)(i=1,2) such that

eRe(Φβ,b𝔞(f,Q))|w|ba(f,Q)(log|w|)j(f,Q)|Φβ,b(f)(w)|C1exp(C2|w|ρ)e^{-\mathop{\rm Re}\nolimits(\Phi_{\beta,b}^{\ast}\mathfrak{a}(f,Q))}|w|^{-ba(f,Q)}(\log|w|)^{-j(f,Q)}\bigl{|}\Phi_{\beta,b}^{\ast}(f)(w)\bigr{|}\geq\\ C^{\prime}_{1}\exp(-C_{2}^{\prime}|w|^{\rho}) (55)

on S(g,R2(1),B(1),L(1))𝒞S(g,R_{2}^{(1)},B^{(1)},L^{(1)})\setminus\mathcal{C}. Because

S(g,R(1),B(1),L(1))S(g,R1(1),B(1),L(1))S(g,R^{(1)},B^{(1)},L^{(1)})\setminus S(g,R^{(1)}_{1},B^{(1)},L^{(1)})

is relatively compact, we obtain the first claim. Take R1(1)>R(1)eLR^{(1)}_{1}>R^{(1)}e^{L}. We can check the second claim by using the argument in the proof of Lemma 4.35. By Lemma 4.36, there exists an infinite sequence TiT_{i} as desired. ∎

4.3.9. Estimates in a multiple case

Let ff be a section of 𝔅~\mathfrak{B}_{\widetilde{{\mathbb{C}}}} on an open subset V~V\subset\widetilde{{\mathbb{C}}}. Let QVϖ1(0)Q\in V\cap\varpi^{-1}(0).

Lemma 4.39.

There exist 𝔞(Q)=b>0zb\mathfrak{a}\in\mathfrak{I}(Q)=\bigoplus_{b>0}{\mathbb{C}}z^{-b}, positive constants CiC_{i} (i=1,2)(i=1,2) and a neighbourhood 𝒰\mathcal{U} of QQ in VV such that

|f|C1|z|C2eRe𝔞|f|\leq C_{1}|z|^{-C_{2}}e^{\mathop{\rm Re}\nolimits\mathfrak{a}}

on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0). If ff has an expression f=fjf=\sum f_{j} such that each fjf_{j} is simply negative, then we may assume 𝔞Q0\mathfrak{a}\prec_{Q}0.

Proof   Let f=fjf=\sum f_{j} be an expression such that each fjf_{j} has a single growth order at QQ. There exists 𝔞(Q)\mathfrak{a}\in\mathfrak{I}(Q) such that 𝔞(fj,Q)Q𝔞\mathfrak{a}(f_{j},Q)\prec_{Q}\mathfrak{a} for any jj. There exists C2>0C_{2}>0 such that C2>max{𝔞(fj,Q)}C_{2}>\max\{\mathfrak{a}(f_{j},Q)\} for any jj. (See Definition 4.17 for 𝔞(fj,Q)\mathfrak{a}(f_{j},Q).) Then, there exist C3>0C_{3}>0 and a neighbourhood 𝒰\mathcal{U} of QQ in VV such that |fj|C3|z|C2eRe(𝔞)|f_{j}|\leq C_{3}|z|^{-C_{2}}e^{\mathop{\rm Re}\nolimits(\mathfrak{a})} on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0) for any jj. Then, we obtain the claim of the lemma. ∎

Let us state a lower bound of ff around QQ outside an exceptional subset. We identify ~=0×S1\widetilde{{\mathbb{C}}}={\mathbb{R}}_{\geq 0}\times S^{1} by the polar decomposition z=re1θz=re^{\sqrt{-1}\theta}. We choose θ0\theta_{0} such that QQ is expressed as (0,e1θ0)(0,e^{\sqrt{-1}\theta_{0}}). Suppose that ff is expressed as a sum i=1mfi\sum_{i=1}^{m}f_{i} on

V0:={(r,e1θ)||θθ0|<π2κ0,   0<r<r0}V,V_{0}:=\Bigl{\{}(r,e^{\sqrt{-1}\theta})\,\Big{|}\,|\theta-\theta_{0}|<\frac{\pi}{2}\kappa_{0},\,\,\,0<r<r_{0}\Bigr{\}}\subset V,

where fif_{i} are holomorphic functions with a single growth order, and κ0\kappa_{0} and r0r_{0} are positive constants.

Lemma 4.40.

Take 0<κ<κ00<\kappa<\kappa_{0} such that max{κdeg𝔞(fi,Q)}<1\max\{\kappa\deg\mathfrak{a}(f_{i},Q)\}<1. There exist a neighbourhood 𝒰\mathcal{U} of QQ in ~\widetilde{{\mathbb{C}}} and a subset Z𝒰ϖ1(0)Z\subset\mathcal{U}\setminus\varpi^{-1}(0) such that the following holds.

  • There exists a subset Z1>0Z_{1}\subset{\mathbb{R}}_{>0} with Z1𝑑t/t<\int_{Z_{1}}dt/t<\infty such that ZZ1×S1Z\subset Z_{1}\times S^{1}.

  • For any δ>0\delta>0, there exists C>0C>0 such that the following holds on 𝒰(ϖ1(0)Z)\mathcal{U}\setminus(\varpi^{-1}(0)\cup Z):

    |f|Cexp(δ|z|1/κ).|f|\geq C\exp\Bigl{(}-\delta|z|^{-1/\kappa}\Bigr{)}.

Proof   We may assume that θ0=0\theta_{0}=0. We set

W2:={w|Re(w)0}.W_{2}:=\{w\in{\mathbb{C}}\,|\,\mathop{\rm Re}\nolimits(w)\geq 0\}.

We define the map Ψ:W2\Psi:W_{2}\longrightarrow{\mathbb{C}} by Ψ(w)=(w+C1)κ\Psi(w)=(w+C_{1})^{-\kappa} for some sufficiently large C1>0C_{1}>0 such that the image of Ψ\Psi is contained in V0V_{0}. We obtain the expression

Ψ(f)=i=1mΨ(fi)\Psi^{\ast}(f)=\sum_{i=1}^{m}\Psi^{\ast}(f_{i})

on W2W_{2}. There exist C2>0C_{2}>0 such that

|Ψ(fi)|C2eRe(Ψ𝔞(fi,Q))|w+C1|κa(fi,Q)(log|w+C1|)j(fi,Q)\bigl{|}\Psi^{\ast}(f_{i})\bigr{|}\leq C_{2}e^{\mathop{\rm Re}\nolimits(\Psi^{\ast}\mathfrak{a}(f_{i},Q))}|w+C_{1}|^{\kappa\cdot a(f_{i},Q)}(\log|w+C_{1}|)^{j(f_{i},Q)}

on W2W_{2}. Take bb such that max{κdeg(𝔞(fi,Q))}<b<1\max\{\kappa\deg(\mathfrak{a}(f_{i},Q))\}<b<1. There exists A>0A>0 such that the following holds on W2W_{2}:

Re(Ψ𝔞(fi,Q))+log(|w+C1|κa(fi,Q)(log|w+C1|)j(fi,Q))ARe((w+C1)b).\mathop{\rm Re}\nolimits(\Psi^{\ast}\mathfrak{a}(f_{i},Q))+\log\bigl{(}|w+C_{1}|^{\kappa\cdot a(f_{i},Q)}(\log|w+C_{1}|)^{j(f_{i},Q)}\bigr{)}\\ \leq A\mathop{\rm Re}\nolimits\bigl{(}(w+C_{1})^{b}\bigr{)}. (56)

Hence, there exists B>0B>0 such that

log|Ψ(f)|ARe((w+C1)b)+B.\log|\Psi^{\ast}(f)|\leq A\mathop{\rm Re}\nolimits\Bigl{(}(w+C_{1})^{b}\Bigr{)}+B. (57)

Let a1,a2,a_{1},a_{2},\ldots be the zeroes of Ψ(f)\Psi^{\ast}(f) in W2W_{2}. According to [22, §14.2, Theorem 2], we obtain

j=1Re(aj)1+|aj|2<.\sum_{j=1}^{\infty}\frac{\mathop{\rm Re}\nolimits(a_{j})}{1+|a_{j}|^{2}}<\infty.

According to [22, §14.2, Theorem 3], we obtain the following description of log|Ψ(f)|\log|\Psi^{\ast}(f)|:

log|Ψ(f)(w)|=j=1log|wajw+a¯j|+Re(w)πdν1(t)|t1w|2Re(w)πdν2(t)|t1w|2+σRe(w).\log|\Psi^{\ast}(f)(w)|=\sum_{j=1}^{\infty}\log\left|\frac{w-a_{j}}{w+\overline{a}_{j}}\right|\\ +\frac{\mathop{\rm Re}\nolimits(w)}{\pi}\int_{-\infty}^{\infty}\frac{d\nu_{1}(t)}{|t-\sqrt{-1}w|^{2}}-\frac{\mathop{\rm Re}\nolimits(w)}{\pi}\int_{-\infty}^{\infty}\frac{d\nu_{2}(t)}{|t-\sqrt{-1}w|^{2}}+\sigma\mathop{\rm Re}\nolimits(w). (58)

Here, σ\sigma is a real number, and νi\nu_{i} are non-negative measures on {\mathbb{R}} such that

dνi(t)1+t2<.\int_{-\infty}^{\infty}\frac{d\nu_{i}(t)}{1+t^{2}}<\infty.

According to the Hayman theorem [22, §15, Theorem 1], there exist a sequence w1,w2,w_{1},w_{2},\ldots in {w|Re(w)>0}\{w\in{\mathbb{C}}\,|\,\mathop{\rm Re}\nolimits(w)>0\} and a sequence of positive numbers ρ1,ρ2,\rho_{1},\rho_{2},\ldots such that ρj/|wj|<\sum\rho_{j}/|w_{j}|<\infty and that

|j=1log|wajw+a¯j||+|Re(w)πdν1(t)|t1w|2|+|Re(w)πdν2(t)|t1w|2|=o(|w|)\left|\sum_{j=1}^{\infty}\log\left|\frac{w-a_{j}}{w+\overline{a}_{j}}\right|\right|+\left|\frac{\mathop{\rm Re}\nolimits(w)}{\pi}\int_{-\infty}^{\infty}\frac{d\nu_{1}(t)}{|t-\sqrt{-1}w|^{2}}\right|\\ +\left|\frac{\mathop{\rm Re}\nolimits(w)}{\pi}\int_{-\infty}^{\infty}\frac{d\nu_{2}(t)}{|t-\sqrt{-1}w|^{2}}\right|=o(|w|) (59)

outside of {w||wwj|<ρj}\bigcup\bigl{\{}w\,\big{|}\,|w-w_{j}|<\rho_{j}\bigr{\}}. By (57), we obtain that σ=0\sigma=0, and hence log|Ψ(f)|=o(|w|)\log\bigl{|}\Psi^{\ast}(f)\bigr{|}=o\bigl{(}|w|\bigr{)} outside of j=1{w||wwj|<ρj}\bigcup_{j=1}^{\infty}\bigl{\{}w\,\big{|}\,|w-w_{j}|<\rho_{j}\bigr{\}}. We note that {|wwj|<ρj}{|wj|ρj<|w|<|wj|+ρj}\{|w-w_{j}|<\rho_{j}\}\subset\{|w_{j}|-\rho_{j}<|w|<|w_{j}|+\rho_{j}\}. Because ρj/|wj|<\sum\rho_{j}/|w_{j}|<\infty, we have ρj/|wj|0\rho_{j}/|w_{j}|\to 0 as jj\to\infty, and there exists A>0A>0 such that

|wj|ρj|wj|+ρj𝑑t/t=log(|wj|+ρj|wj|ρj)Aρj|wj|.\int_{|w_{j}|-\rho_{j}}^{|w_{j}|+\rho_{j}}dt/t=\log\left(\frac{|w_{j}|+\rho_{j}}{|w_{j}|-\rho_{j}}\right)\leq A\frac{\rho_{j}}{|w_{j}|}.

Then, the claim of the lemma follows. ∎

5. Cyclic Higgs bundles with multiple growth orders on sectors

5.1. Statements

Let UU be a neighbourhood of 0 in {\mathbb{C}}. Let ϖ:U~U\varpi:\widetilde{U}\longrightarrow U be the oriented real blowing up. Let VV be an open subset of U~\widetilde{U}. Let ff be a section of 𝔅U~\mathfrak{B}_{\widetilde{U}} on VV. We set q=f(dz)rq=f\cdot(dz)^{r}.

5.1.1. Positive intervals

Let Iϖ1(0)VI\subset\varpi^{-1}(0)\cap V be a positive interval with respect to ff.

Proposition 5.1.

Let h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q). For any relatively compact subset KIK\subset I, there exists a neighbourhood 𝒰K\mathcal{U}_{K} of KK in VV such that h1h_{1} and h2h_{2} are mutually bounded on 𝒰Kϖ1(0)\mathcal{U}_{K}\setminus\varpi^{-1}(0).

5.1.2. Maximal non-positive but non-special intervals

Suppose that there exists an interval II in ϖ1(0)V\varpi^{-1}(0)\cap V satisfying the following conditions.

  • II is maximally non-positive but not special with respect to ff.

  • I¯V\overline{I}\subset V.

Theorem 5.2.

Let h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q). There exists a neighbourhood 𝒰\mathcal{U} of I¯\overline{I} in VV such that h1h_{1} and h2h_{2} are mutually bounded on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0).

5.1.3. Special intervals

Suppose that there exists an interval II in ϖ1(0)V\varpi^{-1}(0)\cap V satisfying the following conditions.

  • I¯V\overline{I}\subset V.

  • II is special with respect to ff.

There exist a non-zero complex number α\alpha and ρ>0\rho>0 such that deg(𝔞(f,Q)αzρ)<ρ\deg(\mathfrak{a}(f,Q)-\alpha z^{-\rho})<\rho for any QI𝒵(f)Q\in I\setminus\mathcal{Z}(f). Let 𝒫\mathcal{P} be the set of the tuples 𝒂=(a1,,ar)r{\boldsymbol{a}}=(a_{1},\ldots,a_{r})\in{\mathbb{R}}^{r} such that

a1a2ara11,ai=0.a_{1}\geq a_{2}\geq\cdots\geq a_{r}\geq a_{1}-1,\quad\quad\sum a_{i}=0.
Theorem 5.3.

  • For any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), there exist 𝒂I(h)=(aI,i(h))𝒫{\boldsymbol{a}}_{I}(h)=(a_{I,i}(h))\in\mathcal{P} and ϵ>0\epsilon>0 such that the following holds on {|arg(αzρ)π|<(1δ)π/2}\bigl{\{}|\arg(\alpha z^{-\rho})-\pi|<(1-\delta)\pi/2\bigr{\}} for any δ>0\delta>0:

    log|(dz)(r+12i)/2|h+aI,i(h)Re(αzρ)=O(|z|ρ+ϵ).\log\bigl{|}(dz)^{(r+1-2i)/2}\bigr{|}_{h}+a_{I,i}(h)\mathop{\rm Re}\nolimits\bigl{(}\alpha z^{-\rho}\bigr{)}=O\bigl{(}|z|^{-\rho+\epsilon}\bigr{)}.

    The tuple 𝒂I(h){\boldsymbol{a}}_{I}(h) is uniquely determined by the condition.

  • For h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q) such that 𝒂I(h1)=𝒂I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}), there exists a neighbourhood 𝒰\mathcal{U} of I¯\overline{I} in VV such that h1h_{1} and h2h_{2} are mutually bounded on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0).

  • For any 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}, there exists hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that 𝒂I(h)=𝒂{\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}.

We shall also prove the following auxiliary statement.

Proposition 5.4.

There exist relatively compact open neighbourhoods 𝒰I,i\mathcal{U}_{I,i} (i=1,2)(i=1,2) of II in VV, an 0{\mathbb{R}}_{\geq 0}-valued harmonic function ϕI\phi_{I} on 𝒰I,1ϖ1(0)\mathcal{U}_{I,1}\setminus\varpi^{-1}(0), and a smooth exhaustive family {Ki}\{K_{i}\} of Vϖ1(0)V\setminus\varpi^{-1}(0) such that the following holds.

  • 𝒰I,2𝒰I,1\mathcal{U}_{I,2}\subset\mathcal{U}_{I,1}.

  • 𝒰I,iϖ1(0)=I\mathcal{U}_{I,i}\cap\varpi^{-1}(0)=I holds, and (𝒰I,iϖ1(0))\partial(\mathcal{U}_{I,i}\setminus\varpi^{-1}(0)) is smooth in Vϖ1(0)V\setminus\varpi^{-1}(0).

  • There exists ϵ>0\epsilon>0 such that ϕI=O(|z|ρ+ϵ)\phi_{I}=O\bigl{(}|z|^{-\rho+\epsilon}\bigr{)} on 𝒰I,1ϖ1(0)\mathcal{U}_{I,1}\setminus\varpi^{-1}(0).

  • Take hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) and hiHarm(q|Ki)h_{i}\in\mathop{\rm Harm}\nolimits(q_{|K_{i}}) such that hi|𝒰I,1Ki=h|𝒰I,1Kih_{i|\mathcal{U}_{I,1}\cap\partial K_{i}}=h_{|\mathcal{U}_{I,1}\cap\partial K_{i}}. Let sis_{i} be the automorphism of 𝕂𝒰I,1Ki,r\mathbb{K}_{\mathcal{U}_{I,1}\cap K_{i},r} determined by hi|𝒰I,1Ki=h|𝒰I,1Kisih_{i|\mathcal{U}_{I,1}\cap K_{i}}=h_{|\mathcal{U}_{I,1}\cap K_{i}}\cdot s_{i}. Then, we obtain

    logTr(si)ϕI\log\mathop{\rm Tr}\nolimits(s_{i})\leq\phi_{I}

    on 𝒰I,2Ki\mathcal{U}_{I,2}\cap K_{i}.

5.2. Some estimates on sectors around infinity

5.2.1. Positive and negative cases

For any R>0R>0 and 0<L<π0<L<\pi, we set W(R,L):={w||w|>R,|arg(w)|<L}W(R,L):=\bigl{\{}w\in{\mathbb{C}}\,\big{|}\,|w|>R,\,\,|\arg(w)|<L\bigr{\}}. Let ff be a holomorphic function on a sector W(R,L)W(R,L). We set q=f(dw)rq=f\,(dw)^{r}. Let Z(f)Z(f) denote the set of the zeroes of ff. Let ρ\rho be a positive number such that ρ<min{1,π/2L}\rho<\min\{1,\pi/2L\}. Let α\alpha be a non-zero complex number. Let 0<κ1<<κm<ρ0<\kappa_{1}<\cdots<\kappa_{m}<\rho and αi\alpha_{i}\in{\mathbb{C}}^{\ast} (i=1,,m)(i=1,\ldots,m). We set

𝔞:=αwρ+i=1mαiwκi.\mathfrak{a}:=\alpha w^{\rho}+\sum_{i=1}^{m}\alpha_{i}w^{\kappa_{i}}.

Let aa\in{\mathbb{R}} and n0n\in{\mathbb{Z}}_{\geq 0}. We impose the following condition on ff in this subsection.

Condition 5.5.

There exists C0>0C_{0}>0 such that the following holds on W(R,L)W(R,L):

eRe𝔞(w)|w|a|logw|n|f|C0.e^{-\mathop{\rm Re}\nolimits\mathfrak{a}(w)}|w|^{-a}|\log w|^{-n}|f|\leq C_{0}. (60)

Moreover, for any L2<L1<LL_{2}<L_{1}<L, there exist R2>R1>RR_{2}>R_{1}>R and subsets 𝒞1𝒞2W(R,L)\mathcal{C}_{1}\subset\mathcal{C}_{2}\subset W(R,L) such that the following holds.

  • Let 𝒟\mathcal{D} be any connected component of 𝒞2\mathcal{C}_{2} such that

    𝒟W(R2,L2).\mathcal{D}\cap W(R_{2},L_{2})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in W(R1,L1)W(R_{1},L_{1}).

  • For any w0W(R1,L1)𝒞2w_{0}\in W(R_{1},L_{1})\setminus\mathcal{C}_{2}, {|ww0|<1}\{|w-w_{0}|<1\} is contained in W(R,L)𝒞1W(R,L)\setminus\mathcal{C}_{1}.

  • There exists C1>0C_{1}>0 such that the following holds on W(R,L)𝒞1W(R,L)\setminus\mathcal{C}_{1}:

    eRe𝔞(w)|w|a|logw|n|f|C1.e^{-\mathop{\rm Re}\nolimits\mathfrak{a}(w)}|w|^{-a}|\log w|^{-n}|f|\geq C_{1}. (61)
Lemma 5.6.

Suppose that Re(αe1ρθ)>0\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\rho\theta})>0 for |θ|L|\theta|\leq L. Then, any h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q) are mutually bounded on W(R,L)W(R^{\prime},L^{\prime}) for any R>RR^{\prime}>R and 0<L<L0<L^{\prime}<L.

Proof   Let ss be the automorphism of 𝕂W(R,L),r\mathbb{K}_{W(R,L),r} determined by h2=h1sh_{2}=h_{1}\cdot s. We set L2=LL_{2}=L^{\prime}. We take L1L_{1} such that L2<L1<LL_{2}<L_{1}<L. We take R2>R1>RR_{2}>R_{1}>R and 𝒞1𝒞2\mathcal{C}_{1}\subset\mathcal{C}_{2} as in Condition 5.5. Let w0W(R1,L1)𝒞2w_{0}\in W(R_{1},L_{1})\setminus\mathcal{C}_{2}. On {|ww0|<1}\{|w-w_{0}|<1\}, the estimates (60) and (61) hold, and we may apply Corollary 3.10. Hence, we obtain the boundedness of Tr(s)\mathop{\rm Tr}\nolimits(s) on W(R1,L1)𝒞2W(R_{1},L_{1})\setminus\mathcal{C}_{2}. Let 𝒟\mathcal{D} be a connected component of 𝒞2\mathcal{C}_{2} with 𝒟W(R2,L2)\mathcal{D}\cap W(R_{2},L_{2})\neq\emptyset. Then, it is relatively compact in W(R1,L1)W(R_{1},L_{1}). Recall that Tr(s)\mathop{\rm Tr}\nolimits(s) is subharmonic (see §2.4). By the maximum principle, we obtain that sup𝒟Tr(s)supW(R1,L1)𝒞2Tr(s)\sup_{\mathcal{D}}\mathop{\rm Tr}\nolimits(s)\leq\sup_{W(R_{1},L_{1})\setminus\mathcal{C}_{2}}\mathop{\rm Tr}\nolimits(s). Hence, we obtain the boundedness of Tr(s)\mathop{\rm Tr}\nolimits(s) on W(R2,L2)W(R_{2},L_{2}). For any 0<R<R20<R^{\prime}<R_{2}, W(R2,L2)W(R,L2)W(R_{2},L_{2})\setminus W(R^{\prime},L_{2}) is relatively compact in W(R,L)W(R,L). Hence, we obtain the boundedness of Tr(s)\mathop{\rm Tr}\nolimits(s) on W(R,L2)W(R^{\prime},L_{2}). ∎

Lemma 5.7.

Suppose that Re(αe1ρθ)<0\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\rho\theta})<0 for |θ|L|\theta|\leq L. Then, for any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), and for any R>RR^{\prime}>R and 0<L<L0<L^{\prime}<L, there exists C>0C>0 such that the following holds on W(R,L)W(R^{\prime},L^{\prime}):

|log|(dw)(r+12i)/2|h|C|Re(αwρ)|=CRe(αwρ).\Bigl{|}\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}\Bigr{|}\leq C\bigl{|}\mathop{\rm Re}\nolimits(\alpha w^{\rho})\bigr{|}=-C\mathop{\rm Re}\nolimits(\alpha w^{\rho}).

Proof   Let θ(q)\theta(q) denote the Higgs field of 𝕂W(R,L),r\mathbb{K}_{W(R,L),r} associated with qq. Take L<L′′<LL^{\prime}<L^{\prime\prime}<L. There exist ϵ>0\epsilon>0, R10>0R_{10}>0 and the map Ψ:W(R10,(1+ϵ)π/2)W(R,L)\Psi:W(R_{10},(1+\epsilon)\pi/2)\longrightarrow W(R,L) defined by Ψ(ζ)=ζ2L′′/π\Psi(\zeta)=\zeta^{2L^{\prime\prime}/\pi}.

For ζ1W(2R10,π/2)\zeta_{1}\in W(2R_{10},\pi/2), we may assume that there exists {|ζζ1|<1}Ψ1(W(R,L))\{|\zeta-\zeta_{1}|<1\}\subset\Psi^{-1}(W(R,L)). By Proposition 3.7, there exists C10>0C_{10}>0, which is independent of ζ1\zeta_{1}, such that |Ψ1(θ(q))|Ψ1(h)C10|\Psi^{-1}(\theta(q))|_{\Psi^{-1}(h)}\leq C_{10} on {|ζζ1|<1/2}\{|\zeta-\zeta_{1}|<1/2\}. Hence, we obtain |Ψ1(θ(q))|Ψ1(h)C10|\Psi^{-1}(\theta(q))|_{\Psi^{-1}(h)}\leq C_{10} on W(2R10,π/2)W(2R_{10},\pi/2).

For ζ1W(R10,π/2)\zeta_{1}\in W(R_{10},\pi/2) such that Re(ζ1)>2R10\mathop{\rm Re}\nolimits(\zeta_{1})>2R_{10}, {|ζζ1|<Re(ζ1)R10}\bigl{\{}|\zeta-\zeta_{1}|<\mathop{\rm Re}\nolimits(\zeta_{1})-R_{10}\bigr{\}} is contained in W(R10,π/2)W(R_{10},\pi/2). There exists the isomorphism Φζ1:{|η|<1}{|ζζ1|<Re(ζ1)R10}\Phi_{\zeta_{1}}:\{|\eta|<1\}\simeq\bigl{\{}|\zeta-\zeta_{1}|<\mathop{\rm Re}\nolimits(\zeta_{1})-R_{10}\bigr{\}} given by Φζ1(η)=ζ1+(Reζ1R10)η\Phi_{\zeta_{1}}(\eta)=\zeta_{1}+(\mathop{\rm Re}\nolimits\zeta_{1}-R_{10})\eta. By applying Proposition 3.7 to Φζ1Ψ(𝕂X(R,L),r,θ(q),h)\Phi_{\zeta_{1}}^{\ast}\Psi^{\ast}(\mathbb{K}_{X(R,L),r},\theta(q),h), we obtain that there exists C11>0C_{11}>0, which is independent of ζ1\zeta_{1}, such that |Φζ1Ψ(θ(q))|Φζ1Ψ(h)C11\bigl{|}\Phi_{\zeta_{1}}^{\ast}\Psi^{\ast}(\theta(q))\bigr{|}_{\Phi_{\zeta_{1}}^{\ast}\Psi^{\ast}(h)}\leq C_{11} on {|η|<1/2}\{|\eta|<1/2\}. Because Φζ1(dζ)=(Reζ1R10)dη\Phi_{\zeta_{1}}^{\ast}(d\zeta)=(\mathop{\rm Re}\nolimits\zeta_{1}-R_{10})d\eta, we obtain

Ψ(θ(q))=O((|Re(ζ)|+1)1)dζ\Psi^{\ast}(\theta(q))=O\bigl{(}(|\mathop{\rm Re}\nolimits(\zeta)|+1)^{-1}\bigr{)}\,d\zeta

with respect to Ψ1(h)\Psi^{-1}(h) on W(R10,π/2)W(R_{10},\pi/2). Recall that R(Ψ(h))R(\Psi^{\ast}(h)) denotes the curvature of the Chern connection of Ψ(𝕂W(R,L),r,h)\Psi^{\ast}(\mathbb{K}_{W(R,L),r},h). The Hitchin equation implies R(Ψ(h))=O((|Re(ζ)|+1)2)dζdζ¯R(\Psi^{\ast}(h))=O\bigl{(}(|\mathop{\rm Re}\nolimits(\zeta)|+1)^{-2}\bigr{)}\,d\zeta\,d\overline{\zeta} on W(R10,π/2)W(R_{10},\pi/2). It implies the following estimates with respect to Ψ1(h)\Psi^{-1}(h) on W(R10,(1κ)π/2)W(R_{10},(1-\kappa)\pi/2) for any κ>0\kappa>0:

Ψ(θ(q))=O(|ζ|1)dζ,R(Ψ(h))=O(|ζ|2)dζdζ¯.\Psi^{\ast}(\theta(q))=O(|\zeta|^{-1})\,d\zeta,\quad R(\Psi^{\ast}(h))=O(|\zeta|^{-2})\,d\zeta\,d\overline{\zeta}.

Take L<L′′′<L′′L^{\prime}<L^{\prime\prime\prime}<L^{\prime\prime}. Recall that R(h)R(h) denotes the curvature of the Chern connection of (𝕂W(R,L),r,h)(\mathbb{K}_{W(R,L),r},h). By the previous consideration, and by w=ζ2L′′/πw=\zeta^{2L^{\prime\prime}/\pi}, we obtain the estimates θ(q)=O(|w|1)dw\theta(q)=O(|w|^{-1})\,dw and R(h)=O(|w|2)dwdw¯R(h)=O(|w|^{-2})\,dw\,d\overline{w} with respect to hh on W(R,L′′′)W(R,L^{\prime\prime\prime}).

We set L2:=LL_{2}:=L^{\prime} and take L2<L1<L′′′L_{2}<L_{1}<L^{\prime\prime\prime}. Take R2>R1>RR_{2}>R_{1}>R and 𝒞1𝒞2\mathcal{C}_{1}\subset\mathcal{C}_{2} as in Condition 5.5. Note that θ(q)=O(|w|1)dw\theta(q)=O(|w|^{-1})\,dw with respect to Ψ1(h)\Psi^{-1}(h) on W(R1,L1)W(R_{1},L_{1}). At any point of W(R1,L1)𝒞1W(R_{1},L_{1})\setminus\mathcal{C}_{1}, the estimate (61) holds. By Corollary 3.5, there exists C12>0C_{12}>0 such that the following holds for i=1,,ri=1,\ldots,r on W(R1,L1)𝒞1W(R_{1},L_{1})\setminus\mathcal{C}_{1}:

|log|(dw)(r+12i)/2|h|C12|Re(αwρ)|.\Bigl{|}\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}\Bigr{|}\leq C_{12}\bigl{|}\mathop{\rm Re}\nolimits(\alpha w^{\rho})\bigr{|}.

Because R(h)=O(|w|2)dwdw¯R(h)=O(|w|^{-2})dw\,d\overline{w} on W(R1,L′′′)W(R_{1},L^{\prime\prime\prime}), we obtain the following estimate on W(R,L′′′)W(R,L^{\prime\prime\prime}):

ww¯log|(dw)(r+12i)/2|h=O(|w|2).\partial_{w}\partial_{\overline{w}}\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}=O(|w|^{-2}).

Hence, by Lemma 4.5, we can prove that there exist functions βi\beta_{i} on W(R1,L1)W(R_{1},L_{1}) such that (i) the functions log|(dw)(r+12i)/2|hβi\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}-\beta_{i} are harmonic, (ii) |βi|=O(log(|w|+1))|\beta_{i}|=O\bigl{(}\log(|w|+1)\bigr{)}. There exists C13>0C_{13}>0 such that the following holds on W(R1,L1)𝒞1W(R_{1},L_{1})\setminus\mathcal{C}_{1}:

C13|Re(αwρ)|log|(dw)(r+12i)/2|hβiC13|Re(αwρ)|.-C_{13}\bigl{|}\mathop{\rm Re}\nolimits(\alpha w^{\rho})\bigr{|}\leq\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}-\beta_{i}\leq C_{13}\bigl{|}\mathop{\rm Re}\nolimits(\alpha w^{\rho})\bigr{|}. (62)

Let 𝒟\mathcal{D} be any connected component of 𝒞2\mathcal{C}_{2} such that 𝒟W(R2,L2)\mathcal{D}\cap W(R_{2},L_{2})\neq\emptyset. Because it is relatively compact in W(R1,L1)W(R_{1},L_{1}), the inequalities (62) hold on 𝒟\mathcal{D}. Hence, (62) holds on W(R2,L2)W(R_{2},L_{2}). Then, we can deduce the claim of the lemma easily. ∎

5.2.2. Neutral case

We set

W~(R,L):={ζ|Re(ζ)>logR,|Im(ζ)|<L}.\widetilde{W}(R,L):=\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,\mathop{\rm Re}\nolimits(\zeta)>\log R,\,\,|\mathop{\rm Im}\nolimits(\zeta)|<L\bigr{\}}.

Let Φ1:\Phi_{1}:{\mathbb{C}}\longrightarrow{\mathbb{C}}^{\ast} be given by Φ1(ζ)=eζ\Phi_{1}(\zeta)=e^{\zeta}. It induces W~(R,L)W(R,L)\widetilde{W}(R,L)\simeq W(R,L). Let aa\in{\mathbb{R}} and n0n\in{\mathbb{Z}}_{\geq 0}. We impose the following condition to ff.

Condition 5.8.

For any L2<L1<LL_{2}<L_{1}<L, there exist R2>R1>RR_{2}>R_{1}>R and subsets 12W~(R,L)\mathcal{B}_{1}\subset\mathcal{B}_{2}\subset\widetilde{W}(R,L) such that the following holds.

  • Let 𝒟\mathcal{D} be any connected component of 2\mathcal{B}_{2} such that

    𝒟W~(R2,L2).\mathcal{D}\cap\widetilde{W}(R_{2},L_{2})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in W~(R1,L1)\widetilde{W}(R_{1},L_{1}).

  • For any ζ0W~(R1,L1)2\zeta_{0}\in\widetilde{W}(R_{1},L_{1})\setminus\mathcal{B}_{2}, we obtain {|ζζ0|<1}W~(R,L)1\{|\zeta-\zeta_{0}|<1\}\subset\widetilde{W}(R,L)\setminus\mathcal{B}_{1}.

  • There exists C>1C>1 such that the following holds on W~(R,L)1\widetilde{W}(R,L)\setminus\mathcal{B}_{1}:

    C1eaRe(ζ)|ζ|n|Φ1(f)(ζ)|C.C^{-1}\leq e^{-a\mathop{\rm Re}\nolimits(\zeta)}|\zeta|^{-n}|\Phi_{1}^{\ast}(f)(\zeta)|\leq C. (63)
Lemma 5.9.

If Condition 5.8 is satisfied, the following holds.

  • If ara\geq-r, any h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q) are mutually bounded on W(R,L)W(R^{\prime},L^{\prime}) for any R>RR^{\prime}>R and 0<L<L0<L^{\prime}<L.

  • Suppose a<ra<-r. Then, for any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), and for any R>RR^{\prime}>R and 0<L<L0<L^{\prime}<L, there exists C>0C>0 such that the following holds for i=1,,ri=1,\ldots,r on W(R,L)W(R^{\prime},L^{\prime}):

    |log|(dw)(r+12i)/2|h|Clog|w|.\Bigl{|}\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}\Bigr{|}\leq C\log|w|. (64)

Proof   Let us study the first claim. We note that Φ1(f(dw)r)=Φ1(f)erζ(dζ)r\Phi_{1}^{\ast}\bigl{(}f\cdot(dw)^{r}\bigr{)}=\Phi_{1}^{\ast}(f)\cdot e^{r\zeta}(d\zeta)^{r}. The inequalities (63) is rewritten as follows:

C1e(a+r)Re(ζ)|ζ|n|Φ1(f)(ζ)erζ|C.C^{-1}\leq e^{-(a+r)\mathop{\rm Re}\nolimits(\zeta)}|\zeta|^{-n}\bigl{|}\Phi_{1}^{\ast}(f)(\zeta)\cdot e^{r\zeta}\bigr{|}\leq C.

For h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q), let ss be the automorphism of 𝕂W~(R,L),r\mathbb{K}_{\widetilde{W}(R,L),r} determined by h2=h1sh_{2}=h_{1}\cdot s. We set L2:=LL_{2}:=L^{\prime}, and take L2<L1<LL_{2}<L_{1}<L. Let R2>R1>RR_{2}>R_{1}>R and 12\mathcal{B}_{1}\subset\mathcal{B}_{2} as in Condition 5.8.

  • If (a+r,n)(0×0){(0,0)}(a+r,n)\in({\mathbb{R}}_{\geq 0}\times{\mathbb{Z}}_{\geq 0})\setminus\{(0,0)\}, we obtain that Tr(s)\mathop{\rm Tr}\nolimits(s) is bounded on W~(R1,L1)2\widetilde{W}(R_{1},L_{1})\setminus\mathcal{B}_{2} by Corollary 3.10.

  • If (a+r,n)=(0,0)(a+r,n)=(0,0), we obtain the boundedness of Tr(s)\mathop{\rm Tr}\nolimits(s) on W~(R1,L1)2\widetilde{W}(R_{1},L_{1})\setminus\mathcal{B}_{2} from Corollary 3.8.

By applying the argument in the proof of Lemma 5.6, we obtain that Tr(s)\mathop{\rm Tr}\nolimits(s) is bounded on W~(R2,L2)\widetilde{W}(R_{2},L_{2}). Then, we obtain the first claim immediately.

Let us study the second claim. Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). By using the argument in the proof of Lemma 5.7, we can prove that there exists C1>0C_{1}>0 such that the following holds on W~(R,L)\widetilde{W}(R^{\prime},L^{\prime}):

|log|(dζ)(r+12i)/2|Φ1(h)|C1(|Re(ζ)|+1)\Bigl{|}\log\bigl{|}(d\zeta)^{(r+1-2i)/2}\bigr{|}_{\Phi_{1}^{\ast}(h)}\Bigr{|}\leq C_{1}\bigl{(}|\mathop{\rm Re}\nolimits(\zeta)|+1\bigr{)} (65)

Because Φ1(dw/w)=dζ\Phi_{1}^{\ast}(dw/w)=d\zeta, we obtain (64) from (65). ∎

5.2.3. Estimates on sectors around infinity in the multiple case

Let us continue to use the notation in §5.2.1.

Lemma 5.10.

Suppose that ff satisfies the following conditions.

  • There exist C0>0C_{0}>0 and κ0>0\kappa_{0}>0 such that |f|C0exp(κ0|w|ρ)|f|\leq C_{0}\exp(\kappa_{0}|w|^{\rho}) on W(R,L)W(R,L).

  • There exist C1>0C_{1}>0, κ1>0\kappa_{1}>0, and a subset Z1>0Z_{1}\subset{\mathbb{R}}_{>0} with Z1𝑑t/t<\int_{Z_{1}}dt/t<\infty, such that |f|C1exp(κ1|w|ρ)|f|\geq C_{1}\exp(-\kappa_{1}|w|^{\rho}) on {wW(R,L)||w|Z1}\bigl{\{}w\in W(R,L)\,\big{|}\,|w|\not\in Z_{1}\bigr{\}}.

Let h1,h2Harm(q)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits(q). Let ss be the automorphism of 𝕂W(R,L),r\mathbb{K}_{W(R,L),r} determined by h2=h1sh_{2}=h_{1}s. Then, for any R>RR^{\prime}>R and 0<L<L0<L^{\prime}<L, there exist C>0C>0 such that |logTr(s)|C|w|ρ\bigl{|}\log\mathop{\rm Tr}\nolimits(s)\bigr{|}\leq C|w|^{\rho} on {wW(R,L)||w|Z1}\bigl{\{}w\in W(R^{\prime},L^{\prime})\,\big{|}\,|w|\not\in Z_{1}\bigr{\}}.

Proof   By Proposition 3.7, there exists a constant C10>0C_{10}>0 and κ10>0\kappa_{10}>0 such that the following holds on W(R,L)W(R^{\prime},L^{\prime}):

|θ(q)|hjC10exp(κ10|w|ρ).|\theta(q)|_{h_{j}}\leq C_{10}\exp(\kappa_{10}|w|^{\rho}).

By Corollary 3.6, there exist C11>0C_{11}>0 and κ11>0\kappa_{11}>0 such that the following holds on {wW(R,L)||w|Z1}\bigl{\{}w\in W(R^{\prime},L^{\prime})\,\big{|}\,|w|\not\in Z_{1}\bigr{\}}:

|s|h1C11exp(κ11|w|ρ).|s|_{h_{1}}\leq C_{11}\exp(\kappa_{11}|w|^{\rho}).

Thus, we obtain the claim of the lemma. ∎

5.3. Proof of Proposition 5.1 and Theorem 5.2

5.3.1. Proof of Proposition 5.1

It is enough to prove the following claim for any QIQ\in I.

  • There exists a relatively compact neighbourhood 𝒰\mathcal{U} of QQ in VV such that h1h_{1} and h2h_{2} are mutually bounded on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0).

If QI𝒵(f)Q\in I\setminus\mathcal{Z}(f), the claim follows from Lemma 5.6. Let us study the case where QI𝒵(f)Q\in I\cap\mathcal{Z}(f). We may assume that I𝒵(f)={Q}I\cap\mathcal{Z}(f)=\{Q\}. Let ss be determined by h2=h1sh_{2}=h_{1}\cdot s. We have already proved that for any QI{Q}Q^{\prime}\in I\setminus\{Q\} there exists a neighbourhood 𝒰Q\mathcal{U}_{Q^{\prime}} in VV such that ss is bounded on 𝒰Qϖ1(0)\mathcal{U}_{Q^{\prime}}\setminus\varpi^{-1}(0). By Lemma 4.39, Lemma 4.40 and Lemma 5.10, there exist N>0N>0, a neighbourhood of 𝒰\mathcal{U} of QQ in VV, and a subset Z1>0Z_{1}\subset{\mathbb{R}}_{>0} with Z1𝑑t/t<\int_{Z_{1}}dt/t<\infty such that logTr(s)=O(|z|N)\log\mathop{\rm Tr}\nolimits(s)=O(|z|^{-N}) on 𝒰(ϖ1(0)(Z1×S1))\mathcal{U}\setminus\bigl{(}\varpi^{-1}(0)\cup(Z_{1}\times S^{1})\bigr{)}. Then, we obtain the boundedness of logTr(s)\log\mathop{\rm Tr}\nolimits(s) on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0) by Corollary 4.3. ∎

5.3.2. Proof of Theorem 5.2

Let Q1,Q2Q_{1},Q_{2} denote the points of I\partial I. Let IiI_{i} be a neighbourhood of QiQ_{i} in ϖ1(0)V\varpi^{-1}(0)\cap V such that Ii𝒵(f)={Qi}I_{i}\cap\mathcal{Z}(f)=\{Q_{i}\}. Then, ff is simply positive at each point of IiI¯I_{i}\setminus\overline{I}.

Let us study the case where II is negative with respect to ff. Then, there exists ρ>0\rho>0 such that deg(𝔞(f,Q))=ρ\deg(\mathfrak{a}(f,Q))=\rho for any QI𝒵(f)Q\in I\setminus\mathcal{Z}(f). Because II is not special with respect to ff, the length of II is strictly less than π/ρ\pi/\rho.

Let ss be the automorphism determined by h2=h1sh_{2}=h_{1}\cdot s. According to Proposition 5.1, Tr(s)\mathop{\rm Tr}\nolimits(s) is bounded around any point of (I1I2)I¯(I_{1}\cup I_{2})\setminus\overline{I}. According to Lemma 5.7, we obtain logTr(s)=O(|z|ρ)\log\mathop{\rm Tr}\nolimits(s)=O\Bigl{(}|z|^{-\rho}\Bigr{)} around any point of I𝒵(f)I\setminus\mathcal{Z}(f). Moreover, according to Lemma 5.10, there exist N>0N>0, a neighbourhood 𝒰0\mathcal{U}_{0} of 𝒵(f)I¯\mathcal{Z}(f)\cap\overline{I} in VV, and a subset Z1>0Z_{1}\subset{\mathbb{R}}_{>0} with Z1𝑑t/t<\int_{Z_{1}}dt/t<\infty, such that logTr(s)=O(|z|N)\log\mathop{\rm Tr}\nolimits(s)=O\Bigl{(}|z|^{-N}\Bigr{)} on 𝒰(ϖ1(0)(Z1×S1))\mathcal{U}\setminus\bigl{(}\varpi^{-1}(0)\cup(Z_{1}\times S^{1})\bigr{)}. By Corollary 4.3, we obtain that logTr(s)=O(|z|ρ)\log\mathop{\rm Tr}\nolimits(s)=O(|z|^{-\rho}) on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0). Because the length of II is strictly smaller than π/ρ\pi/\rho, the Phragmén-Lindelöf theorem (see Corollary 4.2) implies that logTr(s)\log\mathop{\rm Tr}\nolimits(s) is bounded. Thus, we obtain the claim of the proposition in the case where II is negative with respect to ff.

We can prove the claim by a similar and easier argument in the case where II is neutral with respect to ff by using Lemma 5.9 instead of Lemma 5.7. ∎

5.4. Outline of the proof of Theorem 5.3 and Proposition 5.4

5.4.1. Setting in a simple case

Let ϖ:~11\varpi_{\infty}:\widetilde{\mathbb{P}}^{1}_{\infty}\longrightarrow\mathbb{P}^{1} denote the oriented real blow up at \infty. We identify ϖ1()S1\varpi_{\infty}^{-1}(\infty)\simeq S^{1} by the polar decomposition w=|w|e1θw=|w|e^{\sqrt{-1}\theta}.

For R>0R>0 and δ>0\delta>0, we set

X(R,δ):={w||w|>R,|arg(w)π/2|<π/2+δ}.X(R,\delta):=\bigl{\{}w\in{\mathbb{C}}\,\big{|}\,|w|>R,\,|\arg(w)-\pi/2|<\pi/2+\delta\bigr{\}}.

We regard X(R,δ)X(R,\delta) as an open subset of ~1\widetilde{\mathbb{P}}^{1}_{\infty}. Let X¯(R,δ)\overline{X}(R,\delta) denote the closure of X(R,δ)X(R,\delta) in ~1\widetilde{\mathbb{P}}^{1}_{\infty}.

For the proof of Theorem 5.3 and Proposition 5.4, we introduce the coordinate ww determined by αzρ=1w\alpha z^{-\rho}=\sqrt{-1}w, and we study harmonic metrics of the Higgs bundle associated with q=f(dw)rq=f(dw)^{r} on X(R,δ)X(R,\delta) such that deg(𝔞(f,θ)1w)<deg(𝔞(f,θ))\deg(\mathfrak{a}(f,\theta)-\sqrt{-1}w)<\deg(\mathfrak{a}(f,\theta)) for any 0<θ<π0<\theta<\pi.

In the following of this subsection, CiC_{i} will denote positive constants.

As an outline of the proof of Theorem 5.3 and Proposition 5.4, we explain rather detail of our arguments by assuming that ff satisfies the following stronger condition.

  • There exist Ci>0C_{i}>0 (i=1,2,3)(i=1,2,3) such that the following inequalities are satisfied on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

    |f|C1eIm(w).|f|\leq C_{1}e^{-\mathop{\rm Im}\nolimits(w)}. (66)
    |f|C2eIm(w)C3|w|1ϵ.|f|\geq C_{2}e^{-\mathop{\rm Im}\nolimits(w)-C_{3}|w|^{1-\epsilon}}. (67)
  • There exist Ci>0C_{i}>0 (i=4,5,6)(i=4,5,6) such that

    C4eC4|w|C4|f|C5eC5|w|C5C_{4}e^{-C_{4}|w|^{C_{4}}}\leq|f|\leq C_{5}e^{C_{5}|w|^{C_{5}}} (68)

    on {|arg(w)|<C6}{|arg(w)π|<C6}\bigl{\{}|\arg(w)|<C_{6}\bigr{\}}\cup\bigl{\{}|\arg(w)-\pi|<C_{6}\bigr{\}}.

Note that the interval {}×]0,π[\{\infty\}\times]0,\pi[ is special with respect to ff, and hence ff is simply positive at any points of {}×(]δ,0[]π,π+δ[)\{\infty\}\times\bigl{(}]-\delta,0[\cup]\pi,\pi+\delta[\bigr{)}.

5.4.2. Parabolic structure

Let us explain an outline of the proof of the first claim of Theorem 5.3 under the simplified setting in §5.4.1. (We shall study a general case in §5.6.1.)

Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). For any w1w_{1} with Im(w1)>R\mathop{\rm Im}\nolimits(w_{1})>R, there exists an isomorphism

Φw1:{|η|<1}{|ww1|<Im(w1)R}\Phi_{w_{1}}:\{|\eta|<1\}\simeq\{|w-w_{1}|<\mathop{\rm Im}\nolimits(w_{1})-R\}

defined by Φw1(η)=w1+(Im(w1)R)η\Phi_{w_{1}}(\eta)=w_{1}+(\mathop{\rm Im}\nolimits(w_{1})-R)\eta. Note that Φw1(f(dw)r)=Φw1(f)(Im(w1)R)r(dη)r\Phi_{w_{1}}^{\ast}(f(dw)^{r})=\Phi_{w_{1}}^{\ast}(f)\cdot(\mathop{\rm Im}\nolimits(w_{1})-R)^{r}\,(d\eta)^{r}. By (66), there exists C1(1)>0C^{(1)}_{1}>0, which is independent of w1w_{1} such that the following holds on {|η|<1}\{|\eta|<1\}:

|Φw1(f)|(Im(w1)R)rC1(1).|\Phi_{w_{1}}^{\ast}(f)|(\mathop{\rm Im}\nolimits(w_{1})-R)^{r}\leq C^{(1)}_{1}.

Hence, by Proposition 3.7, there exists C2(1)>0C^{(1)}_{2}>0, which is independent of w1w_{1}, such that the following holds on {|η|<1/2}\{|\eta|<1/2\}:

|Φw1(θ(q))|Φw1(h)C2(1).|\Phi_{w_{1}}^{\ast}(\theta(q))|_{\Phi_{w_{1}}^{\ast}(h)}\leq C^{(1)}_{2}.

Because Φw1(dw)=(Im(w1)R)dη\Phi_{w_{1}}^{\ast}(dw)=(\mathop{\rm Im}\nolimits(w_{1})-R)\,d\eta, there exists C3(1)>0C^{(1)}_{3}>0 such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|θ(q)|hC3(1)(ImwR)1.|\theta(q)|_{h}\leq C^{(1)}_{3}(\mathop{\rm Im}\nolimits w-R)^{-1}. (69)

By the Hitchin equation, there exists C4(1)>0C^{(1)}_{4}>0 such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|R(h)|hC4(1)(ImwR)2.|R(h)|_{h}\leq C^{(1)}_{4}(\mathop{\rm Im}\nolimits w-R)^{-2}. (70)

By (70), there exists C5(1)>0C^{(1)}_{5}>0 such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|zz¯log|(dw)(r+12i)/2|h|C5(1)(ImwR)2.\Bigl{|}\partial_{z}\partial_{\overline{z}}\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}\Bigr{|}\leq C^{(1)}_{5}(\mathop{\rm Im}\nolimits w-R)^{-2}. (71)

By (67), (69) and Corollary 3.5, there exists C6(1)C^{(1)}_{6} such that

|log|(dw)(r+12i)/2|h|C6(1)(Im(w)+|w|1ϵ).\Bigl{|}\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}\Bigr{|}\leq C^{(1)}_{6}(\mathop{\rm Im}\nolimits(w)+|w|^{1-\epsilon}). (72)

By (71), (72) and the Nevanlinna formula (Proposition 4.4), there exists ai(h)a_{i}(h) such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

log|(dw)(r+12i)/2|hai(h)Im(w)=O(|w|1ϵ).\log|(dw)^{(r+1-2i)/2}|_{h}-a_{i}(h)\mathop{\rm Im}\nolimits(w)=O(|w|^{1-\epsilon}). (73)

Because of (69) and θ(q)(dw)(r+12i)/2=(dw)(r+12(i+1))/2\theta(q)(dw)^{(r+1-2i)/2}=(dw)^{(r+1-2(i+1))/2}, we obtain ai(h)ai+1(h)a_{i}(h)\geq a_{i+1}(h). By (67), (69), and the relation θ(q)(dw)(r+1)/2=f(dw)(r1)/2\theta(q)(dw)^{(-r+1)/2}=f(dw)^{(r-1)/2}, we also obtain ar(h)a1(h)1a_{r}(h)\geq a_{1}(h)-1. In this way, we obtain the parabolic structure of hh.

5.4.3. Mutually boundedness

Let us explain an outline of the proof of the second claim of Theorem 5.3 under the simplified setting in §5.4.1. (We shall study a general case in §5.6.2.)

Suppose that hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) satisfy 𝒂(h1)=𝒂(h2){\boldsymbol{a}}(h_{1})={\boldsymbol{a}}(h_{2}). Let ss be the automorphism of 𝕂X(R,δ),r\mathbb{K}_{X(R,\delta),r} determined by h2=h1sh_{2}=h_{1}\cdot s. By (21), logTr(s)\log\mathop{\rm Tr}\nolimits(s) is subharmonic.

Note that ff is simply positive at any points of {}×(]δ,0[]π,π+δ[)\{\infty\}\times\bigl{(}]-\delta,0[\cup]\pi,\pi+\delta[\bigr{)}. Hence, by Proposition 5.1, h1h_{1} and h2h_{2} are mutually bounded on X(R1,δ1)X(R1,δ2)X(R_{1},\delta_{1})\setminus X(R_{1},\delta_{2}) for any R1>RR_{1}>R and 0<δ2<δ1<δ0<\delta_{2}<\delta_{1}<\delta.

By the assumption 𝒂(h1)=𝒂(h2){\boldsymbol{a}}(h_{1})={\boldsymbol{a}}(h_{2}), and by (73), we obtain

logTr(s)=O(|w|1ϵ)\log\mathop{\rm Tr}\nolimits(s)=O(|w|^{1-\epsilon})

on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}. By (68) and Proposition 3.7, there exists C1(2)C^{(2)}_{1} such that the following holds on {|arg(w)|<C6/2}{|arg(w)π|<C6/2}\{|\arg(w)|<C_{6}/2\}\cup\{|\arg(w)-\pi|<C_{6}/2\}:

|θ(q)|hjC1(2)eC1(2)|w|C1(2)(j=1,2).|\theta(q)|_{h_{j}}\leq C^{(2)}_{1}e^{C^{(2)}_{1}|w|^{C_{1}^{(2)}}}\quad(j=1,2). (74)

By (68), (74) and Corollary 3.6, there exists C2(2)C^{(2)}_{2} such that the following holds on {|arg(w)|<C6/2}{|arg(w)π|<C6/2}\{|\arg(w)|<C_{6}/2\}\cup\{|\arg(w)-\pi|<C_{6}/2\}:

logTr(s)C2(2)(1+|w|C2(2)).\log\mathop{\rm Tr}\nolimits(s)\leq C^{(2)}_{2}\bigl{(}1+|w|^{C_{2}^{(2)}}\bigr{)}. (75)

By using Phragmén-Lindelöf theorem (Corollary 4.3) on small sectors around arg(w)=0\arg(w)=0 and arg(w)=π\arg(w)=\pi, we obtain that there exists C3(2)C_{3}^{(2)} such that the following holds on X(R1,δ1)X(R_{1},\delta_{1}):

logTr(s)C3(2)(1+|w|1ϵ).\log\mathop{\rm Tr}\nolimits(s)\leq C_{3}^{(2)}(1+|w|^{1-\epsilon}).

By using Phragmén-Lindelöf theorem (Corollary 4.2) again, we obtain that logTr(s)\log\mathop{\rm Tr}\nolimits(s) is bounded on X(R1,δ3)X(R_{1},\delta_{3}) if δ3>0\delta_{3}>0 is sufficiently small. Then, we obtain that logTr(s)\log\mathop{\rm Tr}\nolimits(s) is bounded on X(R1,δ4)X(R_{1},\delta_{4}) for any 0<δ4<δ0<\delta_{4}<\delta, i.e. h1h_{1} and h2h_{2} are mutually bounded on X(R1,δ4)X(R_{1},\delta_{4}).

5.4.4. Auxiliary metrics

We introduce auxiliary metrics under the simplified setting in §5.4.1. (We shall study a general case in §5.6.3.)

We set q0=e1w(dw)rq_{0}=e^{\sqrt{-1}w}(dw)^{r} on {\mathbb{C}}. For any 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}, there exists h0Harm(q0)h_{0}\in\mathop{\rm Harm}\nolimits(q_{0}) such that 𝒂(h0)=𝒂{\boldsymbol{a}}(h_{0})={\boldsymbol{a}}. (See Proposition 3.25.) There exists C1(3)C^{(3)}_{1} such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|θ(q0)|h0C1(3)Im(w)1.|\theta(q_{0})|_{h_{0}}\leq C^{(3)}_{1}\mathop{\rm Im}\nolimits(w)^{-1}. (76)

Note that

θ(q)(dw)(r+12i)/2=(dw)(r+12(i+1))/2dw=θ(q0)(dw)(r+12i)/2\theta(q)(dw)^{(r+1-2i)/2}=(dw)^{(r+1-2(i+1))/2}\,dw=\theta(q_{0})(dw)^{(r+1-2i)/2}

for i=1,,r1i=1,\ldots,r-1. We also note that

θ(q)(dw)(r+1)/2=f(dw)(r1)/2dw,\theta(q)(dw)^{(-r+1)/2}=f(dw)^{(r-1)/2}\,dw,
θ(q0)(dw)(r+1)/2=e1w(dw)(r1)/2dw.\theta(q_{0})(dw)^{(-r+1)/2}=e^{\sqrt{-1}w}(dw)^{(r-1)/2}dw.

We obtain

|θ(q)|h0=O(|θ(q0)|h0).|\theta(q)|_{h_{0}}=O(|\theta(q_{0})|_{h_{0}}).

Hence, there exists C2(3)>0C^{(3)}_{2}>0 such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|θ(q)|h0C2(3)Im(w)1.|\theta(q)|_{h_{0}}\leq C^{(3)}_{2}\mathop{\rm Im}\nolimits(w)^{-1}. (77)

By the Hitchin equation for (𝕂,r,θ(q0),h0)(\mathbb{K}_{{\mathbb{C}},r},\theta(q_{0}),h_{0}), there exists C3(3)>0C^{(3)}_{3}>0 such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|R(h0)|h0C3(3)Im(w)2.|R(h_{0})|_{h_{0}}\leq C^{(3)}_{3}\mathop{\rm Im}\nolimits(w)^{-2}.

We set F(h0,θ(q)):=R(h0)+[θ(q),θ(q)h0]F(h_{0},\theta(q)):=R(h_{0})+[\theta(q),\theta(q)^{\dagger}_{h_{0}}]. There exists C4(3)C^{(3)}_{4} such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

|F(h0,θ(q))|h0C4(3)Im(w)2.|F(h_{0},\theta(q))|_{h_{0}}\leq C_{4}^{(3)}\mathop{\rm Im}\nolimits(w)^{-2}. (78)

5.4.5. Comparison with auxiliary metric

We explain how to obtain an estimate for the difference between a harmonic metric and an auxiliary metric. (We shall study a general case in §5.6.4.)

Let YY be a relatively compact open subset of X(R,δ)X(R,\delta) such that Y\partial Y is smooth, and that

Y{2R1<Imw<4R+1}={2R1<Imw<4R+1,|Re(w)|<aY}Y\cap\{2R-1<\mathop{\rm Im}\nolimits w<4R+1\}=\\ \{2R-1<\mathop{\rm Im}\nolimits w<4R+1,\,|\mathop{\rm Re}\nolimits(w)|<a_{Y}\} (79)

for some aY>2a_{Y}>2. According to Proposition 2.1, there exists hYHarm(q|Y)h_{Y}\in\mathop{\rm Harm}\nolimits(q_{|Y}) such that hY|Y=h0|Yh_{Y|\partial Y}=h_{0|\partial Y}. Let sYs_{Y} be the automorphism of 𝕂Y,r\mathbb{K}_{Y,r} determined by hY=h0|YsYh_{Y}=h_{0|Y}s_{Y}. In the following Ci(4)C^{(4)}_{i} are positive constants which are independent of YY.

By Proposition 3.7 and (66), there exists C1(4)C^{(4)}_{1} such that the following holds on {2R<Imw<4R,|Re(w)|<aY1}\{2R<\mathop{\rm Im}\nolimits w<4R,\,|\mathop{\rm Re}\nolimits(w)|<a_{Y}-1\}:

|θ(q)|hYC1(4).|\theta(q)|_{h_{Y}}\leq C_{1}^{(4)}. (80)

Note that |θ(q)|h0=|θ(q)|hY|\theta(q)|_{h_{0}}=|\theta(q)|_{h_{Y}} on Y\partial Y. By Proposition 3.11, (66), and (77), there exists C2(4)C^{(4)}_{2}, such that the following holds on {2R<Imw<4R,aY1<|Re(w)|<aY}\{2R<\mathop{\rm Im}\nolimits w<4R,\,a_{Y}-1<|\mathop{\rm Re}\nolimits(w)|<a_{Y}\}:

|θ(q)|hYC2(4).|\theta(q)|_{h_{Y}}\leq C_{2}^{(4)}. (81)

By (67), (76), (80), (81) and Corollary 3.6, there exists C3(4)>0C^{(4)}_{3}>0 such that the following holds on {Im(w)=3R}Y\{\mathop{\rm Im}\nolimits(w)=3R\}\cap Y:

logTr(sY)C3(4)|w|1ϵ.\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C_{3}^{(4)}|w|^{1-\epsilon}. (82)

There exists a complex number γ\gamma such that Re(γw1ϵ)>0\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})>0 on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}. There exists C4(4)>0C^{(4)}_{4}>0 such that the following holds on {Im(w)=3R}Y\{\mathop{\rm Im}\nolimits(w)=3R\}\cap Y:

logTr(sY)C4(4)Re(γw1ϵ).\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C_{4}^{(4)}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon}). (83)

By (21) and (78), there exists C5(4)>0C^{(4)}_{5}>0 such that the following holds on {Im(w)>R}\{\mathop{\rm Im}\nolimits(w)>R\}:

zz¯logTr(sY)C5(4)(Imw)2.-\partial_{z}\partial_{\overline{z}}\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C^{(4)}_{5}(\mathop{\rm Im}\nolimits w)^{-2}.

Hence, there exists C6(4)>0C^{(4)}_{6}>0 such that

zz¯(logTr(sY)C3(4)Re(γw1ϵ)C6(4)log(Im(w))logr)0.-\partial_{z}\partial_{\overline{z}}\Bigl{(}\log\mathop{\rm Tr}\nolimits(s_{Y})-C^{(4)}_{3}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})-C^{(4)}_{6}\log(\mathop{\rm Im}\nolimits(w))-\log r\Bigr{)}\leq 0. (84)

On Y{Im(w)=3R}Y\cap\{\mathop{\rm Im}\nolimits(w)=3R\}, we obtain

logTr(sY)C3(4)Re(γw1ϵ)C6(4)log(Im(w))logrlogTr(sY)C3(4)Re(γw1ϵ)0.\log\mathop{\rm Tr}\nolimits(s_{Y})-C^{(4)}_{3}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})-C^{(4)}_{6}\log(\mathop{\rm Im}\nolimits(w))-\log r\\ \leq\log\mathop{\rm Tr}\nolimits(s_{Y})-C^{(4)}_{3}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})\leq 0. (85)

On (Y){Im(w)3R}\partial(Y)\cap\{\mathop{\rm Im}\nolimits(w)\geq 3R\}, we obtain

logTr(sY)C3(4)Re(γw1ϵ)C6(4)log(Im(w))logrlogTr(sY)logr0.\log\mathop{\rm Tr}\nolimits(s_{Y})-C^{(4)}_{3}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})-C^{(4)}_{6}\log(\mathop{\rm Im}\nolimits(w))-\log r\leq\\ \log\mathop{\rm Tr}\nolimits(s_{Y})-\log r\leq 0. (86)

Hence, we obtain logTr(sY)C3(4)Re(γw1ϵ)C6(4)log(Im(w))logr0\log\mathop{\rm Tr}\nolimits(s_{Y})-C^{(4)}_{3}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})-C^{(4)}_{6}\log(\mathop{\rm Im}\nolimits(w))-\log r\leq 0 on Y{Im(w)3R}Y\cap\{\mathop{\rm Im}\nolimits(w)\geq 3R\}. There exists C7(4)C^{(4)}_{7} such that the following holds on Y{Im(w)3R}Y\cap\{\mathop{\rm Im}\nolimits(w)\geq 3R\}:

logTr(sY)C7(4)Re(γw1ϵ)=:ϕ.\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C^{(4)}_{7}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})=:\phi. (87)

5.4.6. Construction of harmonic metrics

Let us explain an outline of the proof of the third claim of Theorem 5.3 under the simplified setting in §5.4.1. (We shall study a general case in §5.6.5.)

Suppose that R>10R>10. Let 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}. We take h0Harm(q0)h_{0}\in\mathop{\rm Harm}\nolimits(q_{0}) such that 𝒂(h0)=𝒂{\boldsymbol{a}}(h_{0})={\boldsymbol{a}}. Let YiY_{i} be a smooth exhaustive family of X(R,δ)X(R,\delta) such that Yi{2R1<Imw<4R+1}={2R1<Imw<4R+1,|Rew|<aYi}Y_{i}\cap\{2R-1<\mathop{\rm Im}\nolimits w<4R+1\}=\{2R-1<\mathop{\rm Im}\nolimits w<4R+1,\,\,|\mathop{\rm Re}\nolimits w|<a_{Y_{i}}\} for aYi>2a_{Y_{i}}>2. We obtain hYiHarm(q|Yi)h_{Y_{i}}\in\mathop{\rm Harm}\nolimits(q_{|Y_{i}}) such that hYi|Yi=h0|Yih_{Y_{i}|\partial Y_{i}}=h_{0|\partial Y_{i}}. According to Proposition 3.15, we may assume that the sequence {hYi}\{h_{Y_{i}}\} is convergent to hHarm(q)h_{\infty}\in\mathop{\rm Harm}\nolimits(q). By (87), we obtain logTr(hYih01)ϕ\log\mathop{\rm Tr}\nolimits(h_{Y_{i}}h_{0}^{-1})\leq\phi on Yi{Im(w)3R}Y_{i}\cap\{\mathop{\rm Im}\nolimits(w)\geq 3R\}, independently of ii. Hence, we obtain logTr(hYh01)ϕ\log\mathop{\rm Tr}\nolimits(h_{Y_{\infty}}h_{0}^{-1})\leq\phi, which implies 𝒂(h)=𝒂(h0)=𝒂{\boldsymbol{a}}(h_{\infty})={\boldsymbol{a}}(h_{0})={\boldsymbol{a}}.

5.4.7. Comparison with harmonic metrics

Let us explain an outline of the proof of Proposition 5.4 under the simplified setting in §5.4.1. (We shall study a general case in §5.6.6.)

Suppose that R>10R>10. Let YY be a relatively compact open subset of X(R,δ)X(R,\delta) such that Y\partial Y is smooth and that it satisfies (79). Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). Let hYHarm(q|Y)h_{Y}\in\mathop{\rm Harm}\nolimits(q_{|Y}) such that hY|Y{Im(w)>R}=h|Y{Im(w)>R}h_{Y|\partial Y\cap\{\mathop{\rm Im}\nolimits(w)>R\}}=h_{|\partial Y\cap\{\mathop{\rm Im}\nolimits(w)>R\}}. We obtain the automorphism sYs_{Y} determined by hY=h|YsYh_{Y}=h_{|Y}\cdot s_{Y}. In the following, constants Ci(5)C^{(5)}_{i} are independent of hh and YY.

By (66), there exists C1(5)>0C^{(5)}_{1}>0 such that the following holds on {Im(w)>2R}\{\mathop{\rm Im}\nolimits(w)>2R\}:

|θ(q)|hC1(5)Im(w)1.|\theta(q)|_{h}\leq C^{(5)}_{1}\mathop{\rm Im}\nolimits(w)^{-1}. (88)

Because hY|Y{Im(w)>R}=h|Y{Im(w)>R}h_{Y|\partial Y\cap\{\mathop{\rm Im}\nolimits(w)>R\}}=h_{|\partial Y\cap\{\mathop{\rm Im}\nolimits(w)>R\}}, the following holds on Y{Im(w)>2R}\partial Y\cap\{\mathop{\rm Im}\nolimits(w)>2R\}:

|θ(q)|hYC1(5)Im(w)1.|\theta(q)|_{h_{Y}}\leq C^{(5)}_{1}\mathop{\rm Im}\nolimits(w)^{-1}. (89)

By Proposition 3.7 and (66), there exists C2(5)C^{(5)}_{2} such that the following holds on {2R<Imw<4R,|Re(w)|aY1}\{2R<\mathop{\rm Im}\nolimits w<4R,\,|\mathop{\rm Re}\nolimits(w)|\leq a_{Y}-1\}:

|θ(q)|hYC2(5).|\theta(q)|_{h_{Y}}\leq C_{2}^{(5)}. (90)

By Proposition 3.11, (66) and (89), there exists C3(5)C^{(5)}_{3} such that the following holds on {2R<Imw<4R,aY1|Re(w)|aY}\{2R<\mathop{\rm Im}\nolimits w<4R,\,a_{Y}-1\leq|\mathop{\rm Re}\nolimits(w)|\leq a_{Y}\}:

|θ(q)|hYC3(5).|\theta(q)|_{h_{Y}}\leq C_{3}^{(5)}. (91)

By (67), (88), (90), (91), and Corollary 3.6, there exists C4(5)>0C^{(5)}_{4}>0 such that the following holds on {Im(w)=3R}Y\{\mathop{\rm Im}\nolimits(w)=3R\}\cap Y:

logTr(sY)C4(5)Re(γw1ϵ).\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C_{4}^{(5)}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon}). (92)

Because logTr(sY)\log\mathop{\rm Tr}\nolimits(s_{Y}) is subharmonic by (21), we obtain the following on Y{Im(w)3R}Y\cap\{\mathop{\rm Im}\nolimits(w)\geq 3R\}:

logTr(sY)C5(5)Re(γw1ϵ)+logr.\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C_{5}^{(5)}\mathop{\rm Re}\nolimits(\gamma w^{1-\epsilon})+\log r.

5.5. Preliminary for the proof of Theorem 5.3 and Proposition 5.4

We prepare some notation and lemmas for the proof of Theorem 5.3 and Proposition 5.4 in §5.6.

5.5.1. Setting

Take R0>10R_{0}>10 and 0<δ0<π/20<\delta_{0}<\pi/2. Let ff be a holomorphic function on X(R0,δ0)X(R_{0},\delta_{0}), which induces a section of 𝔅~1\mathfrak{B}_{\widetilde{\mathbb{P}}^{1}_{\infty}} on X¯(R0,δ0)\overline{X}(R_{0},\delta_{0}). We also assume the following.

  • 𝒵(f)(]δ0,0[]π,π+δ0[)=\mathcal{Z}(f)\cap\bigl{(}]-\delta_{0},0[\cup]\pi,\pi+\delta_{0}[\bigr{)}=\emptyset.

  • For θ]0,π[𝒵(f)\theta\in]0,\pi[\setminus\mathcal{Z}(f), 𝔞(f,θ)\mathfrak{a}(f,\theta) is of the form:

    𝔞(f,θ)=1w+0<κ<1𝔞(f,θ)κwκ.\mathfrak{a}(f,\theta)=\sqrt{-1}w+\sum_{0<\kappa<1}\mathfrak{a}(f,\theta)_{\kappa}w^{\kappa}.

    Note that 𝔞(f,θ)κ\mathfrak{a}(f,\theta)_{\kappa} are constant on each connected component of ]0,π[𝒵(f)]0,\pi[\setminus\mathcal{Z}(f).

For each θ]δ0,π+δ0[\theta\in]-\delta_{0},\pi+\delta_{0}[, there exist a finite subset (f,θ)κ>0wκ\mathcal{I}(f,\theta)\subset\bigoplus_{\kappa>0}{\mathbb{C}}w^{\kappa} and a sector

W(f,θ)={|w|>R(θ),|arg(w)θ|<L(θ)}W(f,\theta)=\bigl{\{}|w|>R(\theta),|\arg(w)-\theta|<L(\theta)\bigr{\}}

such that ff is expressed as f=𝔞(f,θ)fθ,𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}(f,\theta)}f_{\theta,\mathfrak{a}} on W(f,θ)W(f,\theta), where fθ,𝔞f_{\theta,\mathfrak{a}} has a single growth order 𝔞\mathfrak{a}, and R(θ)R(\theta) and L(θ)L(\theta) are positive numbers. We assume that the expression f=𝔞(f,θ)fθ,𝔞f=\sum_{\mathfrak{a}\in\mathcal{I}(f,\theta)}f_{\theta,\mathfrak{a}} is reduced, i.e., for two distinct elements 𝔞,𝔟(f,θ)\mathfrak{a},\mathfrak{b}\in\mathcal{I}(f,\theta), neither 𝔞θ𝔟\mathfrak{a}\prec_{\theta}\mathfrak{b} nor 𝔟θ𝔞\mathfrak{b}\prec_{\theta}\mathfrak{a} holds. (See Lemma 4.22.)

Lemma 5.11.

If 0<θ<π0<\theta<\pi, we obtain deg(𝔞1w)<1\deg(\mathfrak{a}-\sqrt{-1}w)<1 for any 𝔞(f,θ)\mathfrak{a}\in\mathcal{I}(f,\theta). For each 𝔟=κ>0𝔟κwκ(f,0)\mathfrak{b}=\sum_{\kappa>0}\mathfrak{b}_{\kappa}w^{\kappa}\in\mathcal{I}(f,0), the following holds.

  • We have deg(𝔟)1\deg(\mathfrak{b})\geq 1. If deg(𝔟)>1\deg(\mathfrak{b})>1, we obtain 1𝔟deg(𝔟)>0-\sqrt{-1}\mathfrak{b}_{\deg(\mathfrak{b})}>0. If deg(𝔟)=1\deg(\mathfrak{b})=1, we obtain 1𝔟deg(𝔟)1-\sqrt{-1}\mathfrak{b}_{\deg(\mathfrak{b})}\geq 1.

For each 𝔠=κ>0𝔠κwκ(f,π)\mathfrak{c}=\sum_{\kappa>0}\mathfrak{c}_{\kappa}w^{\kappa}\in\mathcal{I}(f,\pi), the following holds.

  • We have deg(𝔠)1\deg(\mathfrak{c})\geq 1. If deg(𝔠)>1\deg(\mathfrak{c})>1, we obtain

    1𝔠deg(𝔠)eπ1deg(𝔠)<0.-\sqrt{-1}\mathfrak{c}_{\deg(\mathfrak{c})}e^{\pi\sqrt{-1}\deg(\mathfrak{c})}<0.

    If deg(𝔠)=1\deg(\mathfrak{c})=1, we obtain 1𝔠deg(𝔠)1\sqrt{-1}\mathfrak{c}_{\deg(\mathfrak{c})}\leq-1.

Proof   Let 0<θ<π0<\theta<\pi and 𝔞(f,θ)\mathfrak{a}\in\mathcal{I}(f,\theta). We have the expression:

𝔞=0<κdeg(𝔞)𝔞κwκ.\mathfrak{a}=\sum_{0<\kappa\leq\deg(\mathfrak{a})}\mathfrak{a}_{\kappa}w^{\kappa}.

Take θ<θ<θ+\theta_{-}<\theta<\theta_{+} such that |θθ±||\theta-\theta_{\pm}| is sufficiently small. We have the expressions

𝔞(f,θ±)=1w+0<κ<1𝔞(f,θ±)κwκ.\mathfrak{a}(f,\theta_{\pm})=\sqrt{-1}w+\sum_{0<\kappa<1}\mathfrak{a}(f,\theta_{\pm})_{\kappa}w^{\kappa}.

If deg(𝔞)>1\deg(\mathfrak{a})>1, we obtain Re(𝔞deg(𝔞)e1θ±deg(𝔞))<0\mathop{\rm Re}\nolimits(\mathfrak{a}_{\deg(\mathfrak{a})}e^{\sqrt{-1}\theta_{\pm}\deg(\mathfrak{a})})<0 from 𝔞θ±𝔞(f,θ±)\mathfrak{a}\prec_{\theta_{\pm}}\mathfrak{a}(f,\theta_{\pm}). It implies that

Re(𝔞deg(𝔞)e1θdeg(𝔞))<0,\mathop{\rm Re}\nolimits(\mathfrak{a}_{\deg(\mathfrak{a})}e^{\sqrt{-1}\theta\deg(\mathfrak{a})})<0,

and hence we obtain 𝔞θ𝔞(f,θ±)\mathfrak{a}\prec_{\theta}\mathfrak{a}(f,\theta_{\pm}), which contradicts that 𝔞,𝔞(f,θ±)(f,θ)\mathfrak{a},\mathfrak{a}(f,\theta_{\pm})\in\mathcal{I}(f,\theta) and that the expression f=fθ,𝔞f=\sum f_{\theta,\mathfrak{a}} is reduced. If deg(𝔞)<1\deg(\mathfrak{a})<1, we obtain 𝔞(f,θ±)θ𝔞\mathfrak{a}(f,\theta_{\pm})\prec_{\theta}\mathfrak{a}, which contradicts that 𝔞,𝔞(f,θ±)(f,θ)\mathfrak{a},\mathfrak{a}(f,\theta_{\pm})\in\mathcal{I}(f,\theta) and that the expression f=fθ,𝔞f=\sum f_{\theta,\mathfrak{a}} is reduced. Hence, we obtain deg(𝔞)=1\deg(\mathfrak{a})=1. Because 𝔞,𝔞(f,θ±)(f,θ)\mathfrak{a},\mathfrak{a}(f,\theta_{\pm})\in\mathcal{I}(f,\theta), we obtain Re(𝔞1e1θ)=Re(1e1θ)\mathop{\rm Re}\nolimits(\mathfrak{a}_{1}e^{\sqrt{-1}\theta})=\mathop{\rm Re}\nolimits(\sqrt{-1}e^{\sqrt{-1}\theta}). If 𝔞11\mathfrak{a}_{1}\neq\sqrt{-1}, then either one of Re(𝔞1e1θ+)>Re(1e1θ+)\mathop{\rm Re}\nolimits(\mathfrak{a}_{1}e^{\sqrt{-1}\theta_{+}})>\mathop{\rm Re}\nolimits(\sqrt{-1}e^{\sqrt{-1}\theta+}) or Re(𝔞1e1θ)>Re(1e1θ)\mathop{\rm Re}\nolimits(\mathfrak{a}_{1}e^{\sqrt{-1}\theta_{-}})>\mathop{\rm Re}\nolimits(\sqrt{-1}e^{\sqrt{-1}\theta-}) holds. It contradicts 𝔞θ±𝔞(f,θ±)\mathfrak{a}\prec_{\theta_{\pm}}\mathfrak{a}(f,\theta_{\pm}). Therefore, we obtain the first claim deg(𝔞1w)<1\deg(\mathfrak{a}-\sqrt{-1}w)<1.

Let us study the second claim. We have the expansion of 𝔟(f,0)\mathfrak{b}\in\mathcal{I}(f,0):

𝔟=0<κ𝔟κwκ.\mathfrak{b}=\sum_{0<\kappa}\mathfrak{b}_{\kappa}w^{\kappa}.

Choose a sufficiently small θ+>0\theta_{+}>0. Note that

𝔞(f,θ+)=1w+0<κ<1𝔞(f,θ+)κwκ(f,0).\mathfrak{a}(f,\theta_{+})=\sqrt{-1}w+\sum_{0<\kappa<1}\mathfrak{a}(f,\theta_{+})_{\kappa}w^{\kappa}\in\mathcal{I}(f,0).

If deg(𝔟)<1\deg(\mathfrak{b})<1, we obtain 𝔞(f,θ+)θ+𝔟\mathfrak{a}(f,\theta_{+})\prec_{\theta_{+}}\mathfrak{b}, which contradicts the choice of 𝔞(f,θ+)\mathfrak{a}(f,\theta_{+}). Hence, we obtain deg(𝔟)1\deg(\mathfrak{b})\geq 1. Because 𝔟,𝔞(f,θ+)(f,0)\mathfrak{b},\mathfrak{a}(f,\theta_{+})\in\mathcal{I}(f,0), we obtain Re(𝔟deg(𝔟))=0\mathop{\rm Re}\nolimits(\mathfrak{b}_{\deg(\mathfrak{b})})=0. Note that 𝔟θ+𝔞(f,θ+)\mathfrak{b}\prec_{\theta_{+}}\mathfrak{a}(f,\theta_{+}). Hence, if deg(𝔟)>1\deg(\mathfrak{b})>1, we obtain Re(𝔟deg𝔟e1deg(𝔟)θ+)<0\mathop{\rm Re}\nolimits(\mathfrak{b}_{\deg\mathfrak{b}}e^{\sqrt{-1}\deg(\mathfrak{b})\theta_{+}})<0. If deg(𝔟)=1\deg(\mathfrak{b})=1, we obtain Re((𝔟deg𝔟1)e1deg(𝔟)θ+)0\mathop{\rm Re}\nolimits\bigl{(}(\mathfrak{b}_{\deg\mathfrak{b}}-\sqrt{-1})e^{\sqrt{-1}\deg(\mathfrak{b})\theta_{+}}\bigr{)}\leq 0. Therefore, 𝔟deg(𝔟)\mathfrak{b}_{\deg(\mathfrak{b})} satisfies the desired condition. We can check the third claim in a similar way. ∎

For each 𝔞=𝔞kwκ(f,θ)\mathfrak{a}=\sum\mathfrak{a}_{k}w^{\kappa}\in\mathcal{I}(f,\theta), we set 𝔰(𝔞):={κ>0|𝔞κ0}\mathfrak{s}(\mathfrak{a}):=\{\kappa\in{\mathbb{R}}_{>0}\,|\,\mathfrak{a}_{\kappa}\neq 0\}, and we set

S(f)=δ0<θ<π+δ0𝔞(f,θ)𝔰(𝔞).S(f)=\bigcup_{-\delta_{0}<\theta<\pi+\delta_{0}}\bigcup_{\mathfrak{a}\in\mathcal{I}(f,\theta)}\mathfrak{s}(\mathfrak{a})\subset{\mathbb{R}}.

Note that S(f)S(f) is a finite subset of {\mathbb{R}}. We choose 0<ϵ0<1/100<\epsilon_{0}<1/10 satisfying the following condition:

ϵ0<min{|κ1κ2|/10|κ1,κ2S,κ1κ2}.\epsilon_{0}<\min\bigl{\{}|\kappa_{1}-\kappa_{2}|/10\,\big{|}\,\kappa_{1},\kappa_{2}\in S,\,\,\kappa_{1}\neq\kappa_{2}\bigr{\}}. (93)

5.5.2. Another coordinate gg and some estimates

Let AA be a complex number such that Re(1A)>0\mathop{\rm Re}\nolimits(\sqrt{-1}A)>0 and Re(1Ae1(1ϵ0)π)>0\mathop{\rm Re}\nolimits(\sqrt{-1}Ae^{\sqrt{-1}(1-\epsilon_{0})\pi})>0. We put g=w+Aw1ϵ0g=w+Aw^{1-\epsilon_{0}} on {|w|>R0,|arg(w)π/2|<π}\{|w|>R_{0},\,\,|\arg(w)-\pi/2|<\pi\}. The following lemma is easy to see.

Lemma 5.12.

For any 0<L2<L1<L0<π0<L_{2}<L_{1}<L_{0}<\pi, there exist R0<R1<R2R_{0}<R_{1}<R_{2} and an open subset

{|w|>R2,|arg(w)π/2|<L2}𝒰{|w|>R0,|arg(w)π/2|<L0},\{|w|>R_{2},|\arg(w)-\pi/2|<L_{2}\}\subset\mathcal{U}\subset\{|w|>R_{0},|\arg(w)-\pi/2|<L_{0}\},

such that gg induces an isomorphism 𝒰{ζ||ζ|>R1,|arg(ζ)π/2|<L1}\mathcal{U}\simeq\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,|\zeta|>R_{1},\,\,|\arg(\zeta)-\pi/2|<L_{1}\bigr{\}}.

Proof   We set 𝒰0:={|w|>R0,|arg(w)π/2|<L0}\mathcal{U}_{0}:=\{|w|>R_{0},\,|\arg(w)-\pi/2|<L_{0}\}. Let 𝒰¯0\overline{\mathcal{U}}_{0} denote the closure of 𝒰0\mathcal{U}_{0} in ~1\widetilde{\mathbb{P}}^{1}_{\infty}. Note that gg induces a continuous map 𝒰¯0~1\overline{\mathcal{U}}_{0}\longrightarrow\widetilde{\mathbb{P}}^{1}_{\infty}, which is also denoted by gg. For any Qϖ1()𝒰¯0Q\in\varpi_{\infty}^{-1}(\infty)\cap\overline{\mathcal{U}}_{0}, we obtain g(Q)=Qg(Q)=Q. Moreover, for the real coordinate systems (|w|1,arg(w))(|w|^{-1},\arg(w)), the map gg is C1C^{1} around QQ, and the derivative at QQ is the identity. Then, the claim of the lemma follows from the inverse function theorem. ∎

Lemma 5.13.

There exist Ci>0C_{i}>0 (i=1,2)(i=1,2) such that

|f|C1eIm(g)|f|\leq C_{1}e^{-\mathop{\rm Im}\nolimits(g)}

on {0Im(g),|g|>C2}\{0\leq\mathop{\rm Im}\nolimits(g),\,\,|g|>C_{2}\}.

Proof   It follows from Lemma 5.18 and Lemma 5.19 below. ∎

Lemma 5.14.

There exists B0>0B_{0}>0 such that gg induces a holomorphic isomorphism

{wX(R0,δ0)|Im(g(w))B0}{ζ|Im(ζ)B0}.\bigl{\{}w\in X(R_{0},\delta_{0})\,\bigr{|}\,\mathop{\rm Im}\nolimits(g(w))\geq B_{0}\bigr{\}}\simeq\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,\mathop{\rm Im}\nolimits(\zeta)\geq B_{0}\bigr{\}}.

Proof   Take δ0<δ1<δ2<π/2\delta_{0}<\delta_{1}<\delta_{2}<\pi/2. There exist R0<R1<R2R_{0}<R_{1}<R_{2} and an open subset X(R2,δ0)𝒰X(R0,δ2)X(R_{2},\delta_{0})\subset\mathcal{U}\subset X(R_{0},\delta_{2}) such that gg induces an isomorphism 𝒰X(R1,δ1)\mathcal{U}\simeq X(R_{1},\delta_{1}). In particular, gg induces an injection on X(R2,δ0)X(R_{2},\delta_{0}). There exists R3>R2R_{3}>R_{2} such that

g1(g(X(R0,δ0)X(R2,δ0)))X(R3,δ0)=.g^{-1}\Bigl{(}g\bigl{(}X(R_{0},\delta_{0})\setminus X(R_{2},\delta_{0})\bigr{)}\Bigr{)}\cap X(R_{3},\delta_{0})=\emptyset.

We obtain g1(g(X(R3,δ0)))X(R0,δ0)=X(R3,δ0)g^{-1}\bigl{(}g(X(R_{3},\delta_{0}))\bigr{)}\cap X(R_{0},\delta_{0})=X(R_{3},\delta_{0}). By Lemma 5.12, there exists R4>0R_{4}>0 such that X(R4,0)g(X(R3,δ0))X(R_{4},0)\subset g(X(R_{3},\delta_{0})). Then, any B0>R4B_{0}>R_{4} satisfies the desired condition. ∎

5.5.3. Some sectors

For any R>0R>0, BB\in{\mathbb{R}} and 0<L<π/40<L<\pi/4, we set

S0(g,R,B,L):={wX(R0,δ0)||g(w)|>R,Im(g(w))>B,arg(g(w))<L},S_{0}(g,R,B,L):=\\ \bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,|g(w)|>R,\,\mathop{\rm Im}\nolimits(g(w))>B,\,\arg(g(w))<L\bigr{\}}, (94)
Sπ(g,R,B,L):={wX(R0,δ0)||g(w)|>R,Im(g(w))>B,arg(g(w))>πL}.S_{\pi}(g,R,B,L):=\\ \bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,|g(w)|>R,\,\mathop{\rm Im}\nolimits(g(w))>B,\,\arg(g(w))>\pi-L\bigr{\}}. (95)

Let B0B_{0} be a positive constant as in Lemma 5.14. By Lemma 5.13, there exists C>0C>0 such that

|f|CeImg|f|\leq Ce^{-\mathop{\rm Im}\nolimits g} (96)

on {wX(R0,δ0)|Img(w)B0}\{w\in X(R_{0},\delta_{0})\,|\,\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}. Let R(0)>0R^{(0)}>0, B(0)>B0B^{(0)}>B_{0} and 0<L(0)<π/40<L^{(0)}<\pi/4 such that R(0)sin(L(0))>10B(0)R^{(0)}\sin(L^{(0)})>10B^{(0)} and

S0(g,R(0),B(0),L(0))W(f,0),Sπ(g,R(0),B(0),L(0))W(f,π),S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\subset W(f,0),\quad S_{\pi}(g,R^{(0)},B^{(0)},L^{(0)})\subset W(f,\pi), (97)
𝒵(f){0<θ<2L(0)}=,𝒵(f){π2L(0)<θ<π}=.\mathcal{Z}(f)\cap\{0<\theta<2L^{(0)}\}=\emptyset,\quad\mathcal{Z}(f)\cap\{\pi-2L^{(0)}<\theta<\pi\}=\emptyset. (98)
Lemma 5.15.

For any B1(0)>B(0)B^{(0)}_{1}>B^{(0)} and 0<L1(0)<L(0)0<L^{(0)}_{1}<L^{(0)}, there exist R1(0)>R(0)R^{(0)}_{1}>R^{(0)} and a subset 𝒞0S0(g,R(0),B(0),L(0))\mathcal{C}_{0}\subset S_{0}(g,R^{(0)},B^{(0)},L^{(0)}) such that the following holds.

  • There exist Ci(0)>0C^{(0)}_{i}>0 (i=1,2)(i=1,2) such that

    |f|C1(0)exp(Re(𝔞(f,θ0))C2(0)|g|ϵ0/2)|f|\geq C^{(0)}_{1}\exp\Bigl{(}\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(f,\theta_{0})\bigr{)}-C^{(0)}_{2}|g|^{\epsilon_{0}/2}\Bigr{)}

    on S0(g,R(0),B(0),L(0))𝒞0S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\setminus\mathcal{C}_{0}, where θ0>0\theta_{0}>0 is sufficiently small.

  • Let 𝒟\mathcal{D} be any connected component of 𝒞0\mathcal{C}_{0} such that

    𝒟S0(g,R1(0),B1(0),L1(0)).\mathcal{D}\cap S_{0}(g,R^{(0)}_{1},B^{(0)}_{1},L^{(0)}_{1})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in S0(g,R(0),B(0),L(0))S_{0}(g,R^{(0)},B^{(0)},L^{(0)}).

  • There exists an increasing sequence of positive numbers T0,iT_{0,i} such that (i) limT0,i=\lim T_{0,i}=\infty, (ii) 𝒞0{|g(w)|=T0,i}=\mathcal{C}_{0}\cap\{|g(w)|=T_{0,i}\}=\emptyset.

Proof   Note that 𝔞(f,θ0)\mathfrak{a}(f,\theta_{0}) is contained in (f,0)\mathcal{I}(f,0). Consider

𝔟(f,0){𝔞(f,θ0)}.\mathfrak{b}\in\mathcal{I}(f,0)\setminus\{\mathfrak{a}(f,\theta_{0})\}.

According to Lemma 5.11, there exist the following three cases; (a) deg(𝔟)>1\deg(\mathfrak{b})>1 and 1𝔟deg𝔟>0-\sqrt{-1}\mathfrak{b}_{\deg\mathfrak{b}}>0, (b) deg(𝔟)=1\deg(\mathfrak{b})=1 and 1𝔟1>1-\sqrt{-1}\mathfrak{b}_{1}>1, (c) deg(𝔟)=1\deg(\mathfrak{b})=1 and 1𝔟1=1-\sqrt{-1}\mathfrak{b}_{1}=1.

According to the first claim of Lemma 5.19 below, if either (a) or (b) is satisfied, there exist 0<L2(0)(𝔟)L(0)0<L^{(0)}_{2}(\mathfrak{b})\leq L^{(0)}, C1(𝔟)>0C_{1}(\mathfrak{b})>0 and C1(𝔟)>0C^{\prime}_{1}(\mathfrak{b})>0 such that Re(𝔟𝔞(f,θ0))C1(𝔟)|g|ϵ0+C1(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{1}(\mathfrak{b})|g|^{\epsilon_{0}}+C^{\prime}_{1}(\mathfrak{b}) holds on S0(g,R(0),B(0),L2(0)(𝔟))S_{0}(g,R^{(0)},B^{(0)},L^{(0)}_{2}(\mathfrak{b})). Because (a) Re(𝔟deg(𝔟)e1deg(𝔟)θ)<0\mathop{\rm Re}\nolimits(\mathfrak{b}_{\deg(\mathfrak{b})}e^{\sqrt{-1}\deg(\mathfrak{b})\theta})<0 or (b) Re((𝔟deg(𝔟)1)e1deg(𝔟)θ)<0\mathop{\rm Re}\nolimits((\mathfrak{b}_{\deg(\mathfrak{b})}-\sqrt{-1})e^{\sqrt{-1}\deg(\mathfrak{b})\theta})<0 holds for any 0<θL(0)0<\theta\leq L^{(0)}, there exist 0<C2(𝔟)<C1(𝔟)0<C_{2}(\mathfrak{b})<C_{1}(\mathfrak{b}) and C2(𝔟)>C1(𝔟)C^{\prime}_{2}(\mathfrak{b})>C^{\prime}_{1}(\mathfrak{b}) such that Re(𝔟𝔞(f,θ0))C2(𝔟)|g|ϵ0+C2(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{2}(\mathfrak{b})|g|^{\epsilon_{0}}+C^{\prime}_{2}(\mathfrak{b}) holds on S0(g,R(0),B(0),L(0)){arg(g(w))>L2(0)(𝔟)/2}S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\cap\{\arg(g(w))>L^{(0)}_{2}(\mathfrak{b})/2\}. Therefore, we obtain Re(𝔟𝔞(f,θ0))C2(𝔟)|g|ϵ0+C2(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{2}(\mathfrak{b})|g|^{\epsilon_{0}}+C_{2}^{\prime}(\mathfrak{b}) holds on S0(g,R(0),B(0),L(0))S_{0}(g,R^{(0)},B^{(0)},L^{(0)}).

In the case (c), by the third claim of Lemma 5.19, there exist 0<L2(0)(𝔟)<L(0)0<L^{(0)}_{2}(\mathfrak{b})<L^{(0)}, C1(𝔟)>0C_{1}(\mathfrak{b})>0 and C1(𝔟)>0C_{1}^{\prime}(\mathfrak{b})>0 such that either Re(𝔟𝔞(f,θ0))C1(𝔟)|g|ϵ0+C1(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{1}(\mathfrak{b})|g|^{\epsilon_{0}}+C_{1}^{\prime}(\mathfrak{b}) or Re(𝔞(f,θ0)𝔟)C1(𝔟)|g|ϵ0+C1(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{a}(f,\theta_{0})-\mathfrak{b})\leq-C_{1}(\mathfrak{b})|g|^{\epsilon_{0}}+C_{1}^{\prime}(\mathfrak{b}) holds on S0(g,R(0),B(0),L2(0)(𝔟))S_{0}(g,R^{(0)},B^{(0)},L^{(0)}_{2}(\mathfrak{b})). Note that 𝔟θ𝔞(f,θ0)\mathfrak{b}\prec_{\theta}\mathfrak{a}(f,\theta_{0}) for 0<θL(0)0<\theta\leq L^{(0)} by Lemma 4.20 and (98). Hence, Re(𝔟𝔞(f,θ0))C1(𝔟)|g|ϵ0+C1(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{1}(\mathfrak{b})|g|^{\epsilon_{0}}+C_{1}^{\prime}(\mathfrak{b}) holds, indeed. Moreover, because 𝔟θ𝔞(f,θ0)\mathfrak{b}\prec_{\theta}\mathfrak{a}(f,\theta_{0}) for 0<θL(0)0<\theta\leq L^{(0)}, and because deg(𝔞(f,θ)𝔟)>2ϵ0\deg(\mathfrak{a}(f,\theta)-\mathfrak{b})>2\epsilon_{0}, there exist 0<C2(𝔟)<C1(𝔟)0<C_{2}(\mathfrak{b})<C_{1}(\mathfrak{b}) and C2(𝔟)>C1(𝔟)C_{2}^{\prime}(\mathfrak{b})>C_{1}^{\prime}(\mathfrak{b}) such that Re(𝔟𝔞(f,θ0))C2(𝔟)|g|ϵ0+C2(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{2}(\mathfrak{b})|g|^{\epsilon_{0}}+C_{2}^{\prime}(\mathfrak{b}) holds on S0(g,R(0),B(0),L(0)){arg(w)>L2(0)(𝔟)/2}S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\cap\{\arg(w)>L^{(0)}_{2}(\mathfrak{b})/2\}. Hence, we obtain Re(𝔟𝔞(f,θ0))C2(𝔟)|g|ϵ0+C2(𝔟)\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{2}(\mathfrak{b})|g|^{\epsilon_{0}}+C_{2}^{\prime}(\mathfrak{b}) on S0(g,R(0),B(0),L(0))S_{0}(g,R^{(0)},B^{(0)},L^{(0)}).

In all, there exist C3>0C_{3}>0 and C3>0C_{3}^{\prime}>0 such that

Re(𝔟𝔞(f,θ0))C3|g|ϵ0+C3\mathop{\rm Re}\nolimits(\mathfrak{b}-\mathfrak{a}(f,\theta_{0}))\leq-C_{3}|g|^{\epsilon_{0}}+C_{3}^{\prime}

on S0(g,R(0),B(0),L(0))S_{0}(g,R^{(0)},B^{(0)},L^{(0)}) for any 𝔟(f,0){𝔞(f,θ0)}\mathfrak{b}\in\mathcal{I}(f,0)\setminus\{\mathfrak{a}(f,\theta_{0})\}. Hence, there exists Ci>0C_{i}>0 (i=4,5)(i=4,5) such that

|𝔟𝔞(f,θ0)f𝔟|C4exp(Re(𝔞(f,θ0))C5|g|ϵ0).\Bigl{|}\sum_{\mathfrak{b}\neq\mathfrak{a}(f,\theta_{0})}f_{\mathfrak{b}}\Bigr{|}\leq C_{4}\exp\Bigl{(}\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(f,\theta_{0})\bigr{)}-C_{5}|g|^{\epsilon_{0}}\Bigr{)}.

Then, we obtain the claim of the lemma by applying Lemma 4.38 with ρ=ϵ0/2\rho=\epsilon_{0}/2. ∎

The following lemma is similar to Lemma 5.15.

Lemma 5.16.

For any B1(0)>B(0)B^{(0)}_{1}>B^{(0)}, and 0<L1(0)<L(0)0<L^{(0)}_{1}<L^{(0)}, there exist R1(0)>R(0)R^{(0)}_{1}>R^{(0)}, and a subset 𝒞πSπ(g,R(0),B(0),L(0))\mathcal{C}_{\pi}\subset S_{\pi}(g,R^{(0)},B^{(0)},L^{(0)}) such that the following holds.

  • There exists Ci(0)>0C^{(0)}_{i}>0 (i=1,2)(i=1,2) such that

    |f|C1(0)exp(Re(𝔞(f,θπ))C2(0)|g|ϵ0/2)|f|\geq C^{(0)}_{1}\exp\Bigl{(}\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(f,\theta_{\pi})\bigr{)}-C^{(0)}_{2}|g|^{\epsilon_{0}/2}\Bigr{)}

    on Sπ(g,R(0),B(0),L(0))𝒞πS_{\pi}(g,R^{(0)},B^{(0)},L^{(0)})\setminus\mathcal{C}_{\pi}, where πθπ>0\pi-\theta_{\pi}>0 is sufficiently small.

  • Let 𝒟\mathcal{D} be any connected component of 𝒞π\mathcal{C}_{\pi} such that

    𝒟Sπ(g,R1(0),B1(0),L1(0)).\mathcal{D}\cap S_{\pi}(g,R^{(0)}_{1},B^{(0)}_{1},L^{(0)}_{1})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in Sπ(g,R(0),B(0),L(0))S_{\pi}(g,R^{(0)},B^{(0)},L^{(0)}).

  • There exists an increasing sequence of positive numbers Tπ,iT_{\pi,i} such that (i) limTπ,i=\lim T_{\pi,i}=\infty, (ii) 𝒞π{|g(w)|=Tπ,i}=\mathcal{C}_{\pi}\cap\{|g(w)|=T_{\pi,i}\}=\emptyset. ∎

5.5.4. Relatively compact open subsets

We introduce a condition for a relatively compact open subset YY in X(R0,δ0)X(R_{0},\delta_{0}), which we shall use in §5.6.4–§5.6.7.

Let B0B_{0} be a constant as in Lemma 5.14. We take R(0)>0R^{(0)}>0, B(0)>B0B^{(0)}>B_{0} and 0<L(0)<π/40<L^{(0)}<\pi/4 as in §5.5.3. We set B1(0)=2B(0)B_{1}^{(0)}=2B^{(0)} and L1(0)=12L(0)L^{(0)}_{1}=\frac{1}{2}L^{(0)}, and let R1(0)>R(0)R_{1}^{(0)}>R^{(0)} be as in Lemma 5.15 and Lemma 5.16. We take B0<B1(0)<B(0)B_{0}<B^{(0)}_{-1}<B^{(0)}, 0<R1(0)<R(0)0<R^{(0)}_{-1}<R^{(0)}, and L(0)<L1(0)<π/4L^{(0)}<L^{(0)}_{-1}<\pi/4 such that

S0(g,R1(0),B1(0),L1(0))W(f,0),Sπ(g,R1(0),B1(0),L1(0))W(f,π).S_{0}(g,R^{(0)}_{-1},B^{(0)}_{-1},L^{(0)}_{-1})\subset W(f,0),\quad S_{\pi}(g,R^{(0)}_{-1},B^{(0)}_{-1},L^{(0)}_{-1})\subset W(f,\pi).

We also assume R1(0)sin(L(0))>10B(0)R^{(0)}_{-1}\sin(L^{(0)})>10B^{(0)}.

Condition 5.17.

  • The boundary Y\partial Y is smooth.

  • There exists T0,i(0)>2R1(0)T_{0,i(0)}>2R^{(0)}_{-1} such that

    YS0(g,R1(0),B1(0),L1(0))=S0(g,R1(0),B1(0),L1(0)){|g(w)|T0,i(0)}.Y\cap S_{0}(g,R_{-1}^{(0)},B_{-1}^{(0)},L_{-1}^{(0)})=S_{0}(g,R_{-1}^{(0)},B_{-1}^{(0)},L_{-1}^{(0)})\cap\{|g(w)|\leq T_{0,i(0)}\}.
  • There exists Tπ,i(π)>2R1(0)T_{\pi,i(\pi)}>2R^{(0)}_{-1} such that

    YSπ(g,R1(0),B1(0),L1(0))=Sπ(g,R1(0),B1(0),L1(0)){|g(w)|Tπ,i(π)}.Y\cap S_{\pi}(g,R_{-1}^{(0)},B_{-1}^{(0)},L_{-1}^{(0)})=S_{\pi}(g,R_{-1}^{(0)},B_{-1}^{(0)},L_{-1}^{(0)})\cap\{|g(w)|\leq T_{\pi,i(\pi)}\}.
  • {wX(R0,δ0)|Img(w)B1(0),|g(w)|2R1(0)}Y\bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,\mathop{\rm Im}\nolimits g(w)\geq B_{-1}^{(0)},\,|g(w)|\leq 2R_{-1}^{(0)}\bigr{\}}\subset Y.

We shall use an exhaustive family satisfying Condition 5.17 to construct harmonic metrics with the desired property in Theorem 5.3 and Theorem 6.1.

5.5.5. Appendix: A choice of complex coordinate

Let R>0R>0 and 0<δ<π/20<\delta<\pi/2. Let 𝔞i\mathfrak{a}_{i} (i=1,,n(1))(i=1,\ldots,n(1)) be mutually distinct holomorphic functions on X(R,δ)X(R,\delta) expressed as a finite sum

𝔞i=1w+0<κ<1𝔞i,κwκ.\mathfrak{a}_{i}=\sqrt{-1}w+\sum_{0<\kappa<1}\mathfrak{a}_{i,\kappa}w^{\kappa}.

Let 𝔟i\mathfrak{b}_{i} (i=1,,n(2))(i=1,\ldots,n(2)) be holomorphic functions on X(R,δ)X(R,\delta) expressed as a finite sum

𝔟i=0<κρ(𝔟i)𝔟i,κwκ,\mathfrak{b}_{i}=\sum_{0<\kappa\leq\rho(\mathfrak{b}_{i})}\mathfrak{b}_{i,\kappa}w^{\kappa},

where one of the following holds.

  • ρ(𝔟i)>1\rho(\mathfrak{b}_{i})>1 and 1𝔟i,ρ(𝔟i)>0-\sqrt{-1}\mathfrak{b}_{i,\rho(\mathfrak{b}_{i})}>0.

  • ρ(𝔟i)=1\rho(\mathfrak{b}_{i})=1 and 1𝔟i,1>1-\sqrt{-1}\mathfrak{b}_{i,1}>1.

Let 𝔠i\mathfrak{c}_{i} (i=1,,n(3))(i=1,\ldots,n(3)) be holomorphic functions on X(R,δ)X(R,\delta) expressed as a finite sum

𝔠i=0<κρ(𝔠i)𝔠i,κwκ,\mathfrak{c}_{i}=\sum_{0<\kappa\leq\rho(\mathfrak{c}_{i})}\mathfrak{c}_{i,\kappa}w^{\kappa},

where one of the following holds.

  • ρ(𝔠i)>1\rho(\mathfrak{c}_{i})>1 and 1𝔠i,ρ(𝔠i)eπ1ρ(𝔠i)<0-\sqrt{-1}\mathfrak{c}_{i,\rho(\mathfrak{c}_{i})}e^{\pi\sqrt{-1}\rho(\mathfrak{c}_{i})}<0.

  • ρ(𝔠i)=1\rho(\mathfrak{c}_{i})=1 and 1𝔠i,1<1\sqrt{-1}\mathfrak{c}_{i,1}<-1.

We set

S={0,1}i{κ|𝔞i,κ0}i{κ|𝔟i,κ0}i{κ|𝔠i,κ0}.S=\{0,1\}\cup\bigcup_{i}\{\kappa\,|\,\mathfrak{a}_{i,\kappa}\neq 0\}\cup\bigcup_{i}\{\kappa\,|\,\mathfrak{b}_{i,\kappa}\neq 0\}\cup\bigcup_{i}\{\kappa\,|\,\mathfrak{c}_{i,\kappa}\neq 0\}.

Take ϵ>0\epsilon>0 such that 2ϵ<|κ1κ2|2\epsilon<|\kappa_{1}-\kappa_{2}| for any two distinct elements κ1,κ2S\kappa_{1},\kappa_{2}\in S.

Let AA be a complex number such that

Re(1A)>0,Re(1Ae1(1ϵ)π)>0.\mathop{\rm Re}\nolimits(\sqrt{-1}A)>0,\quad\mathop{\rm Re}\nolimits(\sqrt{-1}Ae^{\sqrt{-1}(1-\epsilon)\pi})>0.

We set g(w)=w+Aw1ϵg(w)=w+Aw^{1-\epsilon} on X(R,δ)X(R,\delta). In the proof of Lemma 5.14, we observed that there exists R0>RR_{0}>R such that (i) g1(g(X(R0,δ)))=X(R0,δ)g^{-1}\bigl{(}g(X(R_{0},\delta))\bigr{)}=X(R_{0},\delta), (ii) gg induces an isomorphism X(R0,δ)g(X(R0,δ))X(R_{0},\delta)\simeq g(X(R_{0},\delta)). According to Lemma 5.12, there exists R0>R0R_{0}^{\prime}>R_{0} such that X(R0,δ/2)g(X(R0,δ))X(R_{0}^{\prime},\delta/2)\subset g(X(R_{0},\delta)).

Lemma 5.18.

There exist positive constants Rj>R0R_{j}>R^{\prime}_{0} (j=1,2)(j=1,2) and C1C_{1} such that the following conditions are satisfied:

  • Re(𝔞p)Re(1g)<C1|w|1ϵ\mathop{\rm Re}\nolimits(\mathfrak{a}_{p})-\mathop{\rm Re}\nolimits(\sqrt{-1}g)<-C_{1}|w|^{1-\epsilon} for any pp on X(R1,0)={|w|R1,Im(w)0}X(R_{1},0)=\{|w|\geq R_{1},\,\,\,\mathop{\rm Im}\nolimits(w)\geq 0\}.

  • {wX(R,δ)||g(w)|>R2,Img(w)0}X(R1,0)\bigl{\{}w\in X(R,\delta)\,\big{|}\,|g(w)|>R_{2},\,\,\mathop{\rm Im}\nolimits g(w)\geq 0\bigr{\}}\subset X(R_{1},0).

Proof   Note that 1g𝔞p=1Aw1ϵ+0<κ<1ϵαp,κwκ\sqrt{-1}g-\mathfrak{a}_{p}=\sqrt{-1}Aw^{1-\epsilon}+\sum_{0<\kappa<1-\epsilon}\alpha_{p,\kappa}w^{\kappa} for αp,κ\alpha_{p,\kappa}\in{\mathbb{C}}. Because {(p,κ)|αp,κ0}\{(p,\kappa)\,|\,\alpha_{p,\kappa}\neq 0\} is finite, there exist R1R_{1} and C1C_{1} as in the first condition.

By our choice of R0<R0R_{0}<R_{0}^{\prime}, the set

𝒰1:={wX(R,δ)||g(w)|>R0,Img(w)0}{ζ||ζ|>R0,Imζ0}\mathcal{U}_{1}:=\bigl{\{}w\in X(R,\delta)\,\big{|}\,|g(w)|>R^{\prime}_{0},\,\,\mathop{\rm Im}\nolimits g(w)\geq 0\bigr{\}}\simeq\\ \bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,|\zeta|>R^{\prime}_{0},\,\mathop{\rm Im}\nolimits\zeta\geq 0\bigr{\}} (99)

is connected. Clearly, we obtain 𝒰1X(R,0)\mathcal{U}_{1}\cap X(R,0)\neq\emptyset. Because Img(w)=Im(w)Re(1Aw1ϵ)\mathop{\rm Im}\nolimits g(w)=\mathop{\rm Im}\nolimits(w)-\mathop{\rm Re}\nolimits(\sqrt{-1}Aw^{1-\epsilon}), the intersection

{wX(R,δ)||g(w)|>R0,Img(w)0}{wX(R,δ)|Im(w)=0}\bigl{\{}w\in X(R,\delta)\,\big{|}\,|g(w)|>R^{\prime}_{0},\mathop{\rm Im}\nolimits g(w)\geq 0\bigr{\}}\cap\{w\in X(R,\delta)\,|\,\mathop{\rm Im}\nolimits(w)=0\}

is empty. Hence, if R2>R0R_{2}>R_{0}^{\prime} is sufficiently large, we obtain {wX(R,δ)||g(w)|>R2,Img(w)0}X(R1,0)\bigl{\{}w\in X(R,\delta)\,\big{|}\,|g(w)|>R_{2},\,\,\mathop{\rm Im}\nolimits g(w)\geq 0\bigr{\}}\subset X(R_{1},0). ∎

Lemma 5.19.

There exist positive constants LL, R3R_{3} and CjC_{j} (j=2,3)(j=2,3) such that the following holds:

  • Re(𝔟i)Re(𝔞p)<C2|g|ρ(𝔟i)ϵ\mathop{\rm Re}\nolimits(\mathfrak{b}_{i})-\mathop{\rm Re}\nolimits(\mathfrak{a}_{p})<-C_{2}|g|^{\rho(\mathfrak{b}_{i})-\epsilon} for any ii and pp on

    𝒰0:={wX(R,δ)|Reg(w)>0,|g(w)|R3,  0argg(w)L}.\mathcal{U}_{0}:=\bigl{\{}w\in X(R,\delta)\,\big{|}\,\mathop{\rm Re}\nolimits g(w)>0,\,\,|g(w)|\geq R_{3},\,\,0\leq\arg g(w)\leq L\bigr{\}}.
  • Re(𝔠i)Re(𝔞p)<C2|g|ρ(𝔠i)ϵ\mathop{\rm Re}\nolimits(\mathfrak{c}_{i})-\mathop{\rm Re}\nolimits(\mathfrak{a}_{p})<-C_{2}|g|^{\rho(\mathfrak{c}_{i})-\epsilon} for any ii and pp on

    𝒰π:={wX(R,δ)|Reg(w)<0,|g(w)|R3,πLargg(w)π}.\mathcal{U}_{\pi}:=\bigl{\{}w\in X(R,\delta)\,\big{|}\,\mathop{\rm Re}\nolimits g(w)<0,\,\,|g(w)|\geq R_{3},\,\,\pi-L\leq\arg g(w)\leq\pi\bigr{\}}.
  • For pqp\neq q, we obtain C3|g|ϵ|Re(𝔞p𝔞q)|C_{3}|g|^{\epsilon}\leq\bigl{|}\mathop{\rm Re}\nolimits(\mathfrak{a}_{p}-\mathfrak{a}_{q})\bigr{|} on 𝒰0𝒰π\mathcal{U}_{0}\cup\mathcal{U}_{\pi}.

Proof   We use the following estimate on X(R0,δ)X(R_{0},\delta) for some ϵ>0\epsilon^{\prime}>0:

w(gAg1ϵ)=O(|w|1ϵϵ).w-\bigl{(}g-Ag^{1-\epsilon}\bigr{)}=O(|w|^{1-\epsilon-\epsilon^{\prime}}). (100)

Let us study the first claim. To simplify the notation, we set ρ(i):=ρ(𝔟i)\rho(i):=\rho(\mathfrak{b}_{i}). If ρ(i)>1\rho(i)>1, we obtain the following expression

𝔟i𝔞p=𝔟i,ρ(i)wρ(i)+0<κ<ρ(i)𝔟(i,p),κwκ.\mathfrak{b}_{i}-\mathfrak{a}_{p}=\mathfrak{b}_{i,\rho(i)}w^{\rho(i)}+\sum_{0<\kappa<\rho(i)}\mathfrak{b}_{(i,p),\kappa}w^{\kappa}.

By (100), there exists δ(i,p)>0\delta(i,p)>0 such that

𝔟i𝔞p(𝔟i,ρ(i)gρ(i)ρ(i)𝔟i,ρ(i)Agρ(i)ϵ)=O(|g|ρ(i)ϵδ(i,p)).\mathfrak{b}_{i}-\mathfrak{a}_{p}-\Bigl{(}\mathfrak{b}_{i,\rho(i)}g^{\rho(i)}-\rho(i)\cdot\mathfrak{b}_{i,\rho(i)}Ag^{\rho(i)-\epsilon}\Bigr{)}=O(|g|^{\rho(i)-\epsilon-\delta(i,p)}).

By the assumption, there exists b>0b>0 such that 𝔟i,ρ(i)=1b\mathfrak{b}_{i,\rho(i)}=\sqrt{-1}b. We obtain

Re(𝔟i,ρ(i)gρ(i)ρ(i)𝔟i,ρ(i)Agρ(i)ϵ)=b|g|ρ(i)sin(ρ(i)arg(g))bρ(i)|g|ρ(i)ϵRe(1Ae1(ρ(i)ϵ)arg(g)).\mathop{\rm Re}\nolimits\Bigl{(}\mathfrak{b}_{i,\rho(i)}g^{\rho(i)}-\rho(i)\cdot\mathfrak{b}_{i,\rho(i)}Ag^{\rho(i)-\epsilon}\Bigr{)}=\\ -b|g|^{\rho(i)}\sin\bigl{(}\rho(i)\arg(g)\bigr{)}-b\rho(i)|g|^{\rho(i)-\epsilon}\mathop{\rm Re}\nolimits\bigl{(}\sqrt{-1}Ae^{\sqrt{-1}(\rho(i)-\epsilon)\arg(g)}\bigr{)}. (101)

There exists L1(i,p)>0L_{1}(i,p)>0 such that

sin(ρ(i)arg(g))0,Re(1Ae1(ρ(i)ϵ)arg(g))>0\sin\bigl{(}\rho(i)\arg(g)\bigr{)}\geq 0,\quad\mathop{\rm Re}\nolimits\bigl{(}\sqrt{-1}Ae^{\sqrt{-1}(\rho(i)-\epsilon)\arg(g)}\bigr{)}>0

on {0arg(g)L1(i,p)}\bigl{\{}0\leq\arg(g)\leq L_{1}(i,p)\bigr{\}}. Hence, there exist R2>0R_{2}^{\prime}>0 and C3(i,p)>0C_{3}(i,p)>0 such that

Re(𝔟i𝔞p)C3(i,p)|g|ρ(i)ϵ\mathop{\rm Re}\nolimits(\mathfrak{b}_{i}-\mathfrak{a}_{p})\leq-C_{3}(i,p)|g|^{\rho(i)-\epsilon} (102)

on {wX(R,δ)||g(w)|>R2,  0arg(g)L1(i,p)}\bigl{\{}w\in X(R,\delta)\,\big{|}\,|g(w)|>R_{2}^{\prime},\,\,0\leq\arg(g)\leq L_{1}(i,p)\bigr{\}}. If ρ(i)=1\rho(i)=1, there exists b>1b>1 such that 𝔟i,1=1b\mathfrak{b}_{i,1}=\sqrt{-1}b. We obtain

𝔟i𝔞p=(𝔟i,11)w+0<κ<1𝔟(i,p),κwκ=1(b1)w+0<κ<1𝔟(i,p),κwκ.\mathfrak{b}_{i}-\mathfrak{a}_{p}=(\mathfrak{b}_{i,1}-\sqrt{-1})w+\sum_{0<\kappa<1}\mathfrak{b}_{(i,p),\kappa}w^{\kappa}\\ =\sqrt{-1}(b-1)w+\sum_{0<\kappa<1}\mathfrak{b}_{(i,p),\kappa}w^{\kappa}. (103)

By the same argument, we can prove that there exist L1(i,p)>0L_{1}(i,p)>0, R2>0R_{2}^{\prime}>0 and C3(i,p)>0C_{3}(i,p)>0 such that (102) holds on {wX(R,δ)||g(w)|>R2,  0arg(g)L1(i,p)}\bigl{\{}w\in X(R,\delta)\,\big{|}\,|g(w)|>R_{2}^{\prime},\,\,0\leq\arg(g)\leq L_{1}(i,p)\bigr{\}} with ρ(i)=1\rho(i)=1. Thus, we obtain the first claim. The second claim can be proved similarly.

Let us prove the third claim. For pqp\neq q, we set ρ(p,q):=deg(𝔞p𝔞q)>0\rho(p,q):=\deg(\mathfrak{a}_{p}-\mathfrak{a}_{q})>0. We obtain the following expression:

𝔞p𝔞q=0<κρ(p,q)𝔞(p,q),κwκ.\mathfrak{a}_{p}-\mathfrak{a}_{q}=\sum_{0<\kappa\leq\rho(p,q)}\mathfrak{a}_{(p,q),\kappa}w^{\kappa}.

Here, 𝔞(p,q),ρ(p,q)0\mathfrak{a}_{(p,q),\rho(p,q)}\neq 0. There exists δ(p,q)>0\delta(p,q)>0 such that the following holds:

Re(𝔞p𝔞q)Re(𝔞(p,q),ρ(p,q)(gρ(p,q)ρ(p,q)Agρ(p,q)ϵ))=O(|g|ρ(p,q)ϵδ(p,q)).\mathop{\rm Re}\nolimits(\mathfrak{a}_{p}-\mathfrak{a}_{q})-\mathop{\rm Re}\nolimits\Bigl{(}\mathfrak{a}_{(p,q),\rho(p,q)}\bigl{(}g^{\rho(p,q)}-\rho(p,q)Ag^{\rho(p,q)-\epsilon}\bigr{)}\Bigr{)}\\ =O\bigl{(}|g|^{\rho(p,q)-\epsilon-\delta(p,q)}\bigr{)}. (104)

If 𝔞p,q,ρ(p,q)1\mathfrak{a}_{p,q,\rho(p,q)}\not\in\sqrt{-1}{\mathbb{R}}, there exist L3(p,q)>0L_{3}(p,q)>0 and C4(p,q)>0C_{4}(p,q)>0 such that

|Re(𝔞(p,q),ρ(p,q)gρ(p,q))|>C4(p,q)|g|ρ(p,q)\Bigl{|}\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}_{(p,q),\rho(p,q)}g^{\rho(p,q)}\bigr{)}\Bigr{|}>C_{4}(p,q)|g|^{\rho(p,q)} (105)

on {wX(R0,δ)| 0arg(g(w))L3(p,q)}\bigl{\{}w\in X(R_{0},\delta)\,\big{|}\,0\leq\arg(g(w))\leq L_{3}(p,q)\bigr{\}}. If 𝔞p,q,ρ(p,q)1{0}\mathfrak{a}_{p,q,\rho(p,q)}\in\sqrt{-1}{\mathbb{R}}\setminus\{0\}, there exists b0b\neq 0 such that 𝔞p,q,ρ(p,q)=1b\mathfrak{a}_{p,q,\rho(p,q)}=\sqrt{-1}b. Note that

|Re(𝔞(p,q),ρ(p,q)(gρ(p,q)ρ(p,q)Agρ(p,q)ϵ))|=|b|×||g|ρ(p,q)sin(ρ(p,q)argg)+ρ(p,q)|g|ρ(p,q)ϵRe(1Ae1(ρ(p,q)ϵ)arg(g))|.\Bigl{|}\mathop{\rm Re}\nolimits\Bigl{(}\mathfrak{a}_{(p,q),\rho(p,q)}\bigl{(}g^{\rho(p,q)}-\rho(p,q)Ag^{\rho(p,q)-\epsilon}\bigr{)}\Bigr{)}\Bigr{|}=|b|\times\\ \Bigl{|}|g|^{\rho(p,q)}\sin\bigl{(}\rho(p,q)\arg g\bigr{)}+\rho(p,q)|g|^{\rho(p,q)-\epsilon}\mathop{\rm Re}\nolimits\bigl{(}\sqrt{-1}Ae^{\sqrt{-1}(\rho(p,q)-\epsilon)\arg(g)}\bigr{)}\Bigr{|}. (106)

Hence, there exist L3(p,q)>0L_{3}(p,q)>0 and C4(p,q)>0C_{4}(p,q)>0

|Re(𝔞(p,q),ρ(p,q)gρ(p,q))|>C4(p,q)|g|ρ(p,q)ϵ.\Bigl{|}\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}_{(p,q),\rho(p,q)}g^{\rho(p,q)}\bigr{)}\Bigr{|}>C_{4}(p,q)|g|^{\rho(p,q)-\epsilon}.

on {wX(R0,δ)| 0arg(g(w))L3(p,q)}\bigl{\{}w\in X(R_{0},\delta)\,\big{|}\,0\leq\arg(g(w))\leq L_{3}(p,q)\bigr{\}}. Then, we can easily deduce the third claim. ∎

5.6. Proof of Theorem 5.3 and Proposition 5.4

Take R0>10R_{0}>10 and 0<δ0<π/20<\delta_{0}<\pi/2. Let ff be a holomorphic function on X(R0,δ0)X(R_{0},\delta_{0}) as in §5.5.1. We set q=f(dw)rq=f\,(dw)^{r}. We first study GrG_{r}-invariant harmonic metrics of (𝕂X(R0,δ0),θ(q))(\mathbb{K}_{X(R_{0},\delta_{0})},\theta(q)) in §5.6.1–§5.6.6. Then, we shall derive Theorem 5.3 and Proposition 5.4 in §5.6.7.

5.6.1. The associated parabolic weights

Let ϵ0>0\epsilon_{0}>0 be as in §5.5.1.

Proposition 5.20.

Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). Then, there exists 𝐚(h)𝒫{\boldsymbol{a}}(h)\in\mathcal{P} satisfying the following estimate on any {w||w|>R1,|arg(w)π/2|<(1δ)π/2}\bigl{\{}w\in{\mathbb{C}}\,\big{|}\,|w|>R_{1},|\arg(w)-\pi/2|<(1-\delta)\pi/2\bigr{\}} for any R1>R0R_{1}>R_{0} and δ>0\delta>0:

log|(dw)(r+1)/2i|h=ai(h)Im(w)+O(|w|1ϵ0).\log\bigl{|}(dw)^{(r+1)/2-i}\bigr{|}_{h}=a_{i}(h)\cdot\mathop{\rm Im}\nolimits(w)+O\Bigl{(}|w|^{1-\epsilon_{0}}\Bigr{)}. (107)

Proof   For R>0R>0, B>0B>0 and 0<L<π/40<L<\pi/4, let S~(w,R,B,L)\widetilde{S}(w,R,B,L) denote the following set:

{w||w|>R,Im(w)B,arg(w)<L}{w||w|>R,Im(w)B,arg(w)>πL}.\bigl{\{}w\in{\mathbb{C}}\,\big{|}\,|w|>R,\,\mathop{\rm Im}\nolimits(w)\geq B,\,\arg(w)<L\bigr{\}}\cup\\ \bigl{\{}w\in{\mathbb{C}}\,\big{|}\,|w|>R,\,\mathop{\rm Im}\nolimits(w)\geq B,\,\arg(w)>\pi-L\bigr{\}}. (108)

By using the function g=w+Aw1ϵ0g=w+Aw^{1-\epsilon_{0}} as in §5.5.2, from now on we use ww to refer to gg for simplicity. By the results in §5.5.2, the following condition is satisfied.

Condition 5.21.

There exists C0(1)>0C^{(1)}_{0}>0 such that |f|C0(1)eIm(w)|f|\leq C^{(1)}_{0}e^{-\mathop{\rm Im}\nolimits(w)} on {|w|>R0,Im(w)0}\{|w|>R_{0},\,\,\mathop{\rm Im}\nolimits(w)\geq 0\}. Moreover, there exist 2R0<R(1)<R1(1)2R_{0}<R^{(1)}<R^{(1)}_{1}, 2R0<B(1)<B1(1)2R_{0}<B^{(1)}<B^{(1)}_{1}, 0<L1(1)<L(1)0<L^{(1)}_{1}<L^{(1)}, and 𝒞S~(w,R(1),B(1),L(1))\mathcal{C}\subset\widetilde{S}(w,R^{(1)},B^{(1)},L^{(1)}) such that the following conditions are satisfied.

  • There exists Ci(1)>0C^{(1)}_{i}>0 (i=1,2)(i=1,2) such that

    |f|C1(1)exp(Im(w)C2(1)|w|1ϵ0)|f|\geq C^{(1)}_{1}\exp(-\mathop{\rm Im}\nolimits(w)-C^{(1)}_{2}|w|^{1-\epsilon_{0}})

    on S~(w,R(1),B(1),L(1))𝒞\widetilde{S}(w,R^{(1)},B^{(1)},L^{(1)})\setminus\mathcal{C}.

  • Let 𝒟\mathcal{D} be any connected component of 𝒞\mathcal{C} such that

    𝒟S~(w,R1(1),B1(1),L1(1)).\mathcal{D}\cap\widetilde{S}(w,R^{(1)}_{1},B^{(1)}_{1},L^{(1)}_{1})\neq\emptyset.

    Then, 𝒟\mathcal{D} is relatively compact in S~(w,R(1),B(1),L(1))\widetilde{S}(w,R^{(1)},B^{(1)},L^{(1)}).

Let θ(q)\theta(q) denote the Higgs field of 𝕂X(R0,δ0),r\mathbb{K}_{X(R_{0},\delta_{0}),r} associated with qq. For w1w_{1} with Im(w1)2R0\mathop{\rm Im}\nolimits(w_{1})\geq 2R_{0}, {|ww1|<Im(w1)R0/2}\{|w-w_{1}|<\mathop{\rm Im}\nolimits(w_{1})-R_{0}/2\} is contained in {Im(w)>R0}\{\mathop{\rm Im}\nolimits(w)>R_{0}\}. There exists an isomorphism

Φw1:{|η|<1}{|ww1|<Im(w1)R0/2}\Phi_{w_{1}}:\{|\eta|<1\}\simeq\{|w-w_{1}|<\mathop{\rm Im}\nolimits(w_{1})-R_{0}/2\}

given by Φw1(η)=w1+(Im(w1)R0/2)η\Phi_{w_{1}}(\eta)=w_{1}+(\mathop{\rm Im}\nolimits(w_{1})-R_{0}/2)\eta. Note that there exists C3(1)>0C^{(1)}_{3}>0 such that |Φw1(q)|<C3(1)|\Phi_{w_{1}}^{\ast}(q)|<C^{(1)}_{3} independently from w1w_{1}. By applying Proposition 3.7 to Φw1(𝕂X(R0,δ0),r,θ(q),h)\Phi_{w_{1}}^{\ast}(\mathbb{K}_{X(R_{0},\delta_{0}),r},\theta(q),h), we obtain that the sup norm of Φw1(θ(q))\Phi_{w_{1}}^{\ast}(\theta(q)) with respect to Φw1(h)\Phi_{w_{1}}^{\ast}(h) on {|η|<1/2}\{|\eta|<1/2\} is dominated by a constant independently from w1w_{1}. Because Φw1(dw)=(Im(w1)R0/2)dη\Phi_{w_{1}}^{\ast}(dw)=(\mathop{\rm Im}\nolimits(w_{1})-R_{0}/2)d\eta, we obtain |θ(q)|h=O(Im(w)1)|\theta(q)|_{h}=O\bigl{(}\mathop{\rm Im}\nolimits(w)^{-1}\bigr{)} on {Im(w)2R0}\{\mathop{\rm Im}\nolimits(w)\geq 2R_{0}\}. By the Hitchin equation, we also obtain |R(h)|h=O(Im(w)2)|R(h)|_{h}=O\bigl{(}\mathop{\rm Im}\nolimits(w)^{-2}\bigr{)} on {Im(w)2R0}\{\mathop{\rm Im}\nolimits(w)\geq 2R_{0}\}.

Lemma 5.22.

The following estimates hold on S~(w,R1(1),B1(1),L1(1))\widetilde{S}(w,R^{(1)}_{1},B^{(1)}_{1},L^{(1)}_{1}) for i=1,,ri=1,\ldots,r:

log|(dw)(r+1)/2i|h=O(Im(w)+|w|1ϵ0).\log\bigl{|}(dw)^{(r+1)/2-i}\bigr{|}_{h}=O\Bigl{(}\mathop{\rm Im}\nolimits(w)+|w|^{1-\epsilon_{0}}\Bigr{)}. (109)

Proof   We obtain the estimate (109) on S~(w,R(1),B(1),L(1))𝒞\widetilde{S}(w,R^{(1)},B^{(1)},L^{(1)})\setminus\mathcal{C} by the estimate |θ(q)|h=O(Im(w)1)|\theta(q)|_{h}=O\bigl{(}\mathop{\rm Im}\nolimits(w)^{-1}\bigr{)}, Condition 5.21 and Corollary 3.5. Because |R(h)|=O(Im(w)2)|R(h)|=O(\mathop{\rm Im}\nolimits(w)^{-2}) on {Im(w)2R0}\{\mathop{\rm Im}\nolimits(w)\geq 2R_{0}\}, there exist functions βi\beta_{i} on {Im(w)2R0}\{\mathop{\rm Im}\nolimits(w)\geq 2R_{0}\} such that |βi|=O(log(Im(w)))|\beta_{i}|=O\bigl{(}\log(\mathop{\rm Im}\nolimits(w))\bigr{)} and log|(dw)(r+12i)/2|hβi\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}-\beta_{i} are harmonic functions on {Im(w)2R0}\{\mathop{\rm Im}\nolimits(w)\geq 2R_{0}\}. (See Lemma 4.5.) Then, by using Condition 5.21 again, we obtain the estimate (109) on S~(w,R1(1),B1(1),L1(1))\widetilde{S}(w,R^{(1)}_{1},B^{(1)}_{1},L^{(1)}_{1}). ∎

Lemma 5.23.

Let KK be any compact subset of ]0,π[]0,\pi[, which we regard as a closed subset in ϖ1()X¯(R0,δ0)\varpi_{\infty}^{-1}(\infty)\cap\overline{X}(R_{0},\delta_{0}). Then, there exists a neighbourhood 𝒰\mathcal{U} of KK in X¯(R0,δ0)\overline{X}(R_{0},\delta_{0}) such that the estimate (109) holds on 𝒰ϖ1()\mathcal{U}\setminus\varpi_{\infty}^{-1}(\infty).

Proof   It is enough to study the case where KK consists of a point. If K]0,π[𝒵(f)K\subset]0,\pi[\setminus\mathcal{Z}(f), we obtain the claim from Lemma 5.7. If K𝒵(f)]0,π[K\subset\mathcal{Z}(f)\cap]0,\pi[, we obtain the claim from Lemma 5.10 and Corollary 4.3. (See the proof of Theorem 5.2 in §5.3.2 for a more detailed argument.) ∎

By Lemma 5.22 and Lemma 5.23, the estimate (109) holds on {|w|>R1(1),Im(w)B1(1)}\bigl{\{}|w|>R^{(1)}_{1},\mathop{\rm Im}\nolimits(w)\geq B^{(1)}_{1}\bigr{\}}. We also have the following estimate on {Im(w)2R0}\{\mathop{\rm Im}\nolimits(w)\geq 2R_{0}\}:

¯log|(dw)(r+1)/2i|h=O(Im(w)2).\partial\overline{\partial}\log\bigl{|}(dw)^{(r+1)/2-i}\bigr{|}_{h}=O\bigl{(}\mathop{\rm Im}\nolimits(w)^{-2}\bigr{)}.

By Proposition 4.4, there exist ai(h)a_{i}(h)\in{\mathbb{R}} (i=1,,r)(i=1,\ldots,r) such that

log|(dw)(r+12i)/2|hai(h)Im(w)=O(|w|1ϵ0)\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}-a_{i}(h)\mathop{\rm Im}\nolimits(w)=O(|w|^{1-\epsilon_{0}})

on {|w|>R1(1),Im(w)B1(1)}\{|w|>R^{(1)}_{1},\,\mathop{\rm Im}\nolimits(w)\geq B_{1}^{(1)}\}. Because the above ww is referring to gg, it means

log|(dg(w))(r+12i)/2|hai(h)Im(g(w))=O(|g(w)|1ϵ0)\log\bigl{|}(dg(w))^{(r+1-2i)/2}\bigr{|}_{h}-a_{i}(h)\mathop{\rm Im}\nolimits(g(w))=O(|g(w)|^{1-\epsilon_{0}})

on {|g(w)|>R1(1),Img(w)B1(1)}\{|g(w)|>R^{(1)}_{1},\,\mathop{\rm Im}\nolimits g(w)\geq B_{1}^{(1)}\}. By Lemma 5.12, for any δ>0\delta>0, there exists R1>0R_{1}>0 such that {|w|>R1,|arg(w)π/2|<(1δ)π/2}{|g(w)|>R1(1),Img(w)B1(1)}\{|w|>R_{1},\,\,|\arg(w)-\pi/2|<(1-\delta)\pi/2\}\subset\{|g(w)|>R^{(1)}_{1},\mathop{\rm Im}\nolimits g(w)\geq B_{1}^{(1)}\}. Using dg/dw1=O(|w|ϵ0)dg/dw-1=O(|w|^{-\epsilon_{0}}) and Img(w)Imw=O(|w|1ϵ0)\mathop{\rm Im}\nolimits g(w)-\mathop{\rm Im}\nolimits w=O(|w|^{1-\epsilon_{0}}), we obtain

log|(dw)(r+12i)/2|hai(h)Im(w)=O(|w|1ϵ0)\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}-a_{i}(h)\mathop{\rm Im}\nolimits(w)=O(|w|^{1-\epsilon_{0}})

on {|w|>R1,|arg(w)π/2|<(1δ)π/2}\{|w|>R_{1},\,\,|\arg(w)-\pi/2|<(1-\delta)\pi/2\}. Because

θ(q)(dw)(r+12i)/2=(dw)(r+12(i+1))/2dw(i=1,,r1)\theta(q)(dw)^{(r+1-2i)/2}=(dw)^{(r+1-2(i+1))/2}\,dw\quad(i=1,\ldots,r-1)

and |θ(q)|h=O(Im(w)1)|\theta(q)|_{h}=O\bigl{(}\mathop{\rm Im}\nolimits(w)^{-1}\bigr{)}, we obtain ai(h)ai+1(h)a_{i}(h)\geq a_{i+1}(h). We also obtain ar(h)a1(h)1a_{r}(h)\geq a_{1}(h)-1 by using

θ(q)(dw)(r+1)/2=f(dw)(r1)/2(dw)\theta(q)(dw)^{(-r+1)/2}=f(dw)^{(r-1)/2}(dw)

and the estimate of ff around any point of ({}×]0,π[)𝒵(f)(\{\infty\}\times]0,\pi[)\setminus\mathcal{Z}(f) as in Condition 5.5. Thus, the proof of Proposition 5.20 is completed. ∎

5.6.2. Mutually boundedness

Proposition 5.24.

Let hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) such that 𝐚(h1)=𝐚(h2){\boldsymbol{a}}(h_{1})={\boldsymbol{a}}(h_{2}). Then, for any R1>R0R_{1}>R_{0} and 0<δ1<δ00<\delta_{1}<\delta_{0}, h1h_{1} and h2h_{2} are mutually bounded on X(R1,δ1)X(R_{1},\delta_{1}).

Proof   Take any 0<δ1<δ00<\delta_{1}<\delta_{0} and R1>R0R_{1}>R_{0}. It is enough to prove that there exists R1>R1R^{\prime}_{1}>R_{1} such that h1h_{1} and h2h_{2} are mutually bounded on X(R1,δ1)X(R^{\prime}_{1},\delta_{1}) because X(R1,δ1)X(R1,δ1)X(R_{1},\delta_{1})\setminus X(R_{1}^{\prime},\delta_{1}) is compact.

Let ss be determined by h2=h1sh_{2}=h_{1}\cdot s. Because det(s)=1\det(s)=1, we obtain Tr(s)r\mathop{\rm Tr}\nolimits(s)\geq r. By Proposition 5.1, Tr(s)\mathop{\rm Tr}\nolimits(s) is bounded on X(R1,δ1)X(R1,δ3)X(R_{1},\delta_{1})\setminus X(R_{1},\delta_{3}) for any 0<δ3<δ10<\delta_{3}<\delta_{1}.

By the assumption 𝒂(h1)=𝒂(h2){\boldsymbol{a}}(h_{1})={\boldsymbol{a}}(h_{2}), the following estimate holds on {|w|>R1,|arg(w)π/2|<(1δ)π/2}\bigl{\{}|w|>R_{1},\,\,|\arg(w)-\pi/2|<(1-\delta)\pi/2\bigr{\}} for any δ>0\delta>0:

logTr(s)=O(|w|1ϵ0).\log\mathop{\rm Tr}\nolimits(s)=O\bigl{(}|w|^{1-\epsilon_{0}}\bigr{)}.

By Lemma 5.10, there exist N>0N>0 and a subset Z1>0Z_{1}\subset{\mathbb{R}}_{>0} with Z1𝑑t/t<\int_{Z_{1}}dt/t<\infty such that

logTr(s)=O(|w|N)\log\mathop{\rm Tr}\nolimits(s)=O(|w|^{N})

on {|w|>R1,δ<arg(w)<δ,|w|Z1}\bigl{\{}|w|>R_{1},\,\,-\delta<\arg(w)<\delta,\,\,|w|\not\in Z_{1}\bigr{\}} for any δ>0\delta>0. By Corollary 4.3, we obtain that

logTr(s)=O(|w|1ϵ0)\log\mathop{\rm Tr}\nolimits(s)=O\bigl{(}|w|^{1-\epsilon_{0}}\bigr{)}

on {|w|>R1,δ1<arg(w)<δ1}\bigl{\{}|w|>R_{1},\,\,-\delta_{1}<\arg(w)<\delta_{1}\bigr{\}}. Similarly, we obtain

logTr(s)=O(|w|1ϵ0)\log\mathop{\rm Tr}\nolimits(s)=O\bigl{(}|w|^{1-\epsilon_{0}}\bigr{)}

on {|w|>R1,πδ1<arg(w)<π+δ1}\{|w|>R_{1},\,\,\pi-\delta_{1}<\arg(w)<\pi+\delta_{1}\}. If 0<δ4δ10<\delta_{4}\leq\delta_{1} is sufficiently small, we obtain that logTr(s)\log\mathop{\rm Tr}\nolimits(s) is bounded on X(R1,δ4)X(R_{1},\delta_{4}) by Corollary 4.2. Because logTr(s)\log\mathop{\rm Tr}\nolimits(s) is bounded on X(R1,δ1)X(R1,δ4/2)X(R_{1},\delta_{1})\setminus X(R_{1},\delta_{4}/2), we obtain that logTr(s)\log\mathop{\rm Tr}\nolimits(s) is bounded on X(R1,δ1)X(R_{1},\delta_{1}). ∎

5.6.3. Auxiliary metrics

To introduce an auxiliary metric h0h_{0}, we set q0:=e1g(w)(dg(w))rq_{0}:=e^{\sqrt{-1}g(w)}(dg(w))^{r} on X(R0,δ0)X(R_{0},\delta_{0}).

Lemma 5.25.

For any 𝐚𝒫{\boldsymbol{a}}\in\mathcal{P}, there exists h0Harm(q0)h_{0}\in\mathop{\rm Harm}\nolimits(q_{0}) such that the following holds on 𝒰1:={wX(R0,δ0)|Im(g(w))B0}\mathcal{U}_{1}:=\bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,\mathop{\rm Im}\nolimits(g(w))\geq B_{0}\bigr{\}}:

log|(dw)(r+1)/2i|h0=aiIm(w)+O(|w|1ϵ0).\log\bigl{|}(dw)^{(r+1)/2-i}\bigr{|}_{h_{0}}=a_{i}\cdot\mathop{\rm Im}\nolimits(w)+O\Bigl{(}|w|^{1-\epsilon_{0}}\Bigr{)}.

We also obtain

|θ(q)|h0=O((Img(w))1),|F(h0,θ(q))|h0=O((Img(w))2)|\theta(q)|_{h_{0}}=O\Bigl{(}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-1}\Bigr{)},\quad|F(h_{0},\theta(q))|_{h_{0}}=O\Bigl{(}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-2}\Bigr{)}

on 𝒰1\mathcal{U}_{1}. Here, F(h0,θ(q))=R(h0)+[θ(q),θ(q)h0]F(h_{0},\theta(q))=R(h_{0})+[\theta(q),\theta(q)^{\dagger}_{h_{0}}].

Proof   We consider the map Φ:\Phi:{\mathbb{C}}\longrightarrow{\mathbb{C}}^{\ast} by Φ(w)=e1g(w)/r\Phi(w)=e^{\sqrt{-1}g(w)/r}. We set q1:=(1rdz)rq_{1}:=(-\sqrt{-1}rdz)^{r} on {\mathbb{C}}^{\ast}. We obtain q0=Φ(q1)q_{0}=\Phi^{\ast}(q_{1}). Note that dg/dw1=O(|w|ϵ0)dg/dw-1=O(|w|^{-\epsilon_{0}}) and Im(g(w))y=O(|w|1ϵ0)\mathop{\rm Im}\nolimits(g(w))-y=O(|w|^{1-\epsilon_{0}}). Then, the first claim is reduced to [30]. (See Proposition 3.25.) We obtain |θ(q0)|h0=O((Img(w))1)|\theta(q_{0})|_{h_{0}}=O\Bigl{(}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-1}\Bigr{)}. Because |f|=O(|e1g(w)|)|f|=O\bigl{(}|e^{\sqrt{-1}g(w)}|\bigr{)}, we obtain |θ(q)|h0=O(|θ(q0)|h0)|\theta(q)|_{h_{0}}=O\bigl{(}|\theta(q_{0})|_{h_{0}}\bigr{)}. By the Hitchin equation R(h0)+[θ(q0),θ(q0)h0]=0R(h_{0})+[\theta(q_{0}),\theta(q_{0})^{\dagger}_{h_{0}}]=0, we obtain R(h0)=O((Img(w))2)R(h_{0})=O\Bigl{(}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-2}\Bigr{)}. Hence, we obtain |F(h0,θ(q))|h0=O((Img(w))2)|F(h_{0},\theta(q))|_{h_{0}}=O\Bigl{(}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-2}\Bigr{)}. Thus, we obtain the second claim. ∎

5.6.4. Comparison with auxiliary metrics

Let h0h_{0} be as in Lemma 5.25. For any relatively compact open subset YX(R0,δ0)Y\subset X(R_{0},\delta_{0}) satisfying Condition 5.17, there uniquely exists hYHarm(q|Y)h_{Y}\in\mathop{\rm Harm}\nolimits(q_{|Y}) such that hY|Y=h0|Yh_{Y|\partial Y}=h_{0|\partial Y}, according to Proposition 2.1.

Lemma 5.26.

There exists C0(2)>0C^{(2)}_{0}>0 which is independent of YY, such that the following holds on the region S0(g,R(0),B(0),L(0))YS_{0}(g,R^{(0)},B^{(0)},L^{(0)})\cap Y:

|θ(q)|hYC0(2)Im(g(w))1.|\theta(q)|_{h_{Y}}\leq C^{(2)}_{0}\mathop{\rm Im}\nolimits(g(w))^{-1}. (110)

Proof   Because hY|Y=h0|Yh_{Y|\partial Y}=h_{0|\partial Y}, there exists C1(2)>0C^{(2)}_{1}>0, which is independent of YY, such that the following holds on YS0(g,R1(0),B1(0),L1(0))\partial Y\cap S_{0}(g,R_{-1}^{(0)},B_{-1}^{(0)},L_{-1}^{(0)}):

|θ(q)|hYC1(2)(Im(g(w))1).|\theta(q)|_{h_{Y}}\leq C^{(2)}_{1}(\mathop{\rm Im}\nolimits(g(w))^{-1}).

We set

D:=110min{B(0)B1(0),R(0)R1(0),R1(0)sin(L1(0)L(0)),R1(0)sin(L1(0))B(0)}.D:=\\ \frac{1}{10}\min\bigl{\{}B^{(0)}-B^{(0)}_{-1},R^{(0)}-R^{(0)}_{-1},R^{(0)}_{-1}\sin(L^{(0)}_{-1}-L^{(0)}),R^{(0)}_{-1}\sin(L^{(0)}_{-1})-B^{(0)}\bigr{\}}. (111)

Take any w1S0(g,R(0),B(0),L(0))Yw_{1}\in S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\cap Y such that

B(0)Im(g(w1))B(0)+D.B^{(0)}\leq\mathop{\rm Im}\nolimits(g(w_{1}))\leq B^{(0)}+D.

We set

(w1):={wX(R0,δ0)||g(w)g(w1)|<D}{ζ||ζg(w1)|<D}.\mathcal{B}(w_{1}):=\bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,|g(w)-g(w_{1})|<D\bigr{\}}\simeq\\ \bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,|\zeta-g(w_{1})|<D\bigr{\}}. (112)

Note that (w1)S0(g,R1(0),B1(0),L1(0))\mathcal{B}(w_{1})\subset S_{0}(g,R^{(0)}_{-1},B^{(0)}_{-1},L^{(0)}_{-1}). By applying Proposition 3.11 to hY|Y(w1)h_{Y|Y\cap\mathcal{B}(w_{1})}, we obtain that there exists C2(2)>0C_{2}^{(2)}>0, which is independent of YY and w1w_{1} as above, such that the following holds on Y{wX(R0,δ0)||g(w)g(w1)|<D/2}Y\cap\bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,|g(w)-g(w_{1})|<D/2\bigr{\}}:

|θ(q)|hYC2(2).\bigl{|}\theta(q)\bigr{|}_{h_{Y}}\leq C_{2}^{(2)}.

Take any w1S0(g,R(0),B(0),L(0))w_{1}\in S_{0}(g,R^{(0)},B^{(0)},L^{(0)}) such that Img(w1)>D+B(0)\mathop{\rm Im}\nolimits g(w_{1})>D+B^{(0)} and that argg(w1)<L1(0)/2\arg g(w_{1})<L^{(0)}_{-1}/2. We set ρ(w1):=(Img(w1)B(0))/2\rho(w_{1}):=(\mathop{\rm Im}\nolimits g(w_{1})-B^{(0)})/2. Then, (w1):={wX(R0,δ0)||g(w)g(w1)|<ρ(w1)}\mathcal{B}(w_{1}):=\bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,|g(w)-g(w_{1})|<\rho(w_{1})\bigr{\}} is contained in the following set:

S0(g,R1(0),B1(0),L1(0)){wX(R0,δ0)|Img(w)B1(0),|g(w)|2R1(0)}.S_{0}(g,R^{(0)}_{-1},B^{(0)}_{-1},L^{(0)}_{-1})\cup\\ \bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,\mathop{\rm Im}\nolimits g(w)\geq B^{(0)}_{-1},|g(w)|\leq 2R^{(0)}_{-1}\bigr{\}}. (113)

Let Ψw1\Psi_{w_{1}} be the isomorphism {ζ||ζg(w1)|<ρ(w1)}{|η|<1}\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,|\zeta-g(w_{1})|<\rho(w_{1})\bigr{\}}\simeq\{|\eta|<1\} determined by Ψw1(ζ)=ρ(w1)1(ζg(w1))\Psi_{w_{1}}(\zeta)=\rho(w_{1})^{-1}\bigl{(}\zeta-g(w_{1})\bigr{)}. Note that there exists C3(2)1C^{(2)}_{3}\geq 1, which is independent of YY and w1w_{1} as above, such that

(C3(2))1|(Ψw1g)(dη)/dw|Img(w1)B(0)C3(2).(C^{(2)}_{3})^{-1}\leq\frac{\bigl{|}(\Psi_{w_{1}}\circ g)^{\ast}(d\eta)/dw\bigr{|}}{\mathop{\rm Im}\nolimits g(w_{1})-B^{(0)}}\leq C^{(2)}_{3}.

There exists qw1q_{w_{1}} on {|η|<1}\bigl{\{}|\eta|<1\bigr{\}} such that (Ψw1g)(qw1)=q(\Psi_{w_{1}}\circ g)^{\ast}(q_{w_{1}})=q on (w1)\mathcal{B}(w_{1}). By (96), there exists C4(2)>0C^{(2)}_{4}>0, which is independent of YY and w1w_{1} as above, such that |qw1|C4(2)|q_{w_{1}}|\leq C^{(2)}_{4}. We also obtain

hw1Harm(qw1|{|η|<1}(Ψw1g)(Y))h_{w_{1}}\in\mathop{\rm Harm}\nolimits\bigl{(}q_{w_{1}|\{|\eta|<1\}\cap(\Psi_{w_{1}}\circ g)(Y)}\bigr{)}

such that (Ψw1g)(hw1)=hY(\Psi_{w_{1}}\circ g)^{\ast}(h_{w_{1}})=h_{Y} on Y(w1)Y\cap\mathcal{B}(w_{1}). By applying Proposition 3.11 to hw1|{|η|<1}(Ψw1g)(Y)h_{w_{1}|\{|\eta|<1\}\cap(\Psi_{w_{1}}\circ g)(Y)}, we obtain that there exists C5(2)>0C^{(2)}_{5}>0, which is independent of YY and w1w_{1} as above, such that the following holds on {|η|<1/2}(Ψw1g)(Y)\{|\eta|<1/2\}\cap(\Psi_{w_{1}}\circ g)(Y):

|θ(qw1)|hw1C5(2).\bigl{|}\theta(q_{w_{1}})\bigr{|}_{h_{w_{1}}}\leq C^{(2)}_{5}.

Hence, there exists C6(2)>0C^{(2)}_{6}>0, which is independent of YY and w1w_{1} as above, such that the following holds on {|g(w)g(w1)|<ρ(w1)/2}Y\bigl{\{}|g(w)-g(w_{1})|<\rho(w_{1})/2\bigr{\}}\cap Y:

|θ(q)|hYC6(2)(Img(w))1.\bigl{|}\theta(q)\bigr{|}_{h_{Y}}\leq C^{(2)}_{6}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-1}.

Take any w1S0(g,R(0),B(0),L(0))w_{1}\in S_{0}(g,R^{(0)},B^{(0)},L^{(0)}) such that argg(w1)>L(0)/2\arg g(w_{1})>L^{(0)}/2. We set

ρ(w1):=|g(w1)|sin((L1(0)argw1)/2).\rho(w_{1}):=|g(w_{1})|\sin\bigl{(}(L^{(0)}_{-1}-\arg w_{1})/2\bigr{)}.

Then, (w1)={wX(R0,δ0)||g(w)g(w1)|<ρ(w1)}\mathcal{B}(w_{1})=\Bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,|g(w)-g(w_{1})|<\rho(w_{1})\Bigr{\}} is contained in (113). Let Ψw1\Psi_{w_{1}} be the isomorphism {ζ||ζg(w1)|<ρ(w1)}{|η|<1}\bigl{\{}\zeta\in{\mathbb{C}}\,\big{|}\,|\zeta-g(w_{1})|<\rho(w_{1})\bigr{\}}\simeq\{|\eta|<1\} determined by Ψw1(ζ)=ρ(w1)1(ζg(w1))\Psi_{w_{1}}(\zeta)=\rho(w_{1})^{-1}(\zeta-g(w_{1})). Note that there exists C7(2)1C^{(2)}_{7}\geq 1, which is independent of YY and w1w_{1} such that

(C7(2))1|(Ψw1g)(dη)/dw||g(w1)|C7(2).(C^{(2)}_{7})^{-1}\leq\frac{\bigl{|}(\Psi_{w_{1}}\circ g)^{\ast}(d\eta)/dw\bigr{|}}{|g(w_{1})|}\leq C^{(2)}_{7}.

There exists qw1q_{w_{1}} on {|η|<1}\bigl{\{}|\eta|<1\bigr{\}} such that (Ψw1g)(qw1)=q(\Psi_{w_{1}}\circ g)^{\ast}(q_{w_{1}})=q on (w1)\mathcal{B}(w_{1}). By (96), there exists C8(2)>0C^{(2)}_{8}>0, which is independent of YY and w1w_{1}, such that |qw1|C8(2)|q_{w_{1}}|\leq C^{(2)}_{8}. We also obtain hw1Harm(qw1|{|η|<1}(Ψw1g)(Y))h_{w_{1}}\in\mathop{\rm Harm}\nolimits\bigl{(}q_{w_{1}|\{|\eta|<1\}\cap(\Psi_{w_{1}}\circ g)(Y)}\bigr{)} such that (Ψw1g)(hw1)=hY(\Psi_{w_{1}}\circ g)^{\ast}(h_{w_{1}})=h_{Y} on Y(w1)Y\cap\mathcal{B}(w_{1}). By applying Proposition 3.11 to hw1|{|η|<1}(Ψw1g)(Y)h_{w_{1}|\{|\eta|<1\}\cap(\Psi_{w_{1}}\circ g)(Y)}, we obtain that there exists C9(2)>0C^{(2)}_{9}>0, which is independent of YY and w1w_{1}, such that the following holds on {|η|<1/2}(Ψw1g)(Y)\{|\eta|<1/2\}\cap(\Psi_{w_{1}}\circ g)(Y):

|θ(qw1)|hw1C9(2).\bigl{|}\theta(q_{w_{1}})\bigr{|}_{h_{w_{1}}}\leq C^{(2)}_{9}.

Hence, there exists C10(2)>0C^{(2)}_{10}>0 which is independent of YY and w1w_{1} such that the following holds on {|g(w)g(w1)|<ρ(w1)/2}Y\bigl{\{}|g(w)-g(w_{1})|<\rho(w_{1})/2\bigr{\}}\cap Y:

|θ(q)|hYC10(2)|g(w)|1.\bigl{|}\theta(q)\bigr{|}_{h_{Y}}\leq C^{(2)}_{10}|g(w)|^{-1}.

Note that there exists C11(2)1C^{(2)}_{11}\geq 1, such that (C11(2))1|g(w)|Img(w)C11(2)|g(w)|(C^{(2)}_{11})^{-1}|g(w)|\leq\mathop{\rm Im}\nolimits g(w)\leq C^{(2)}_{11}|g(w)| on S0(g,R(0),B(0),L(0)){arg(g(w))>L(0)/2}S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\cap\{\arg(g(w))>L^{(0)}/2\}. Thus, we obtain the claim of Lemma 5.26. ∎

Lemma 5.27.

There exists an 0{\mathbb{R}}_{\geq 0}-valued harmonic function ϕ\phi on {Im(g(w))B0}\{\mathop{\rm Im}\nolimits(g(w))\geq B_{0}\} such that the following holds.

  • ϕ=O(|g(w)|1ϵ0)\phi=O\bigl{(}|g(w)|^{1-\epsilon_{0}}\bigr{)}.

  • For any YY satisfying Condition 5.17, let sYs_{Y} be the automorphism of 𝕂Y,r\mathbb{K}_{Y,r} determined by hY=h0|YsYh_{Y}=h_{0|Y}\cdot s_{Y}. Then, we obtain logTr(sY)ϕ\log\mathop{\rm Tr}\nolimits(s_{Y})\leq\phi on Y{Im(g(w))B1(0)}Y\cap\{\mathop{\rm Im}\nolimits(g(w))\geq B_{1}^{(0)}\}.

Proof   We recall that |F(h0,θ(q))|h0=O((Img(w))2)\bigl{|}F(h_{0},\theta(q))\bigr{|}_{h_{0}}=O\bigl{(}(\mathop{\rm Im}\nolimits g(w))^{-2}\bigr{)} by Lemma 5.25. Hence, there exists an 0{\mathbb{R}}_{\geq 0}-valued function ϕ1\phi_{1} on {Img(w)B0}\{\mathop{\rm Im}\nolimits g(w)\geq B_{0}\} such that (i) ϕ1=O(log(1+Img(w)))\phi_{1}=O\bigl{(}\log(1+\mathop{\rm Im}\nolimits g(w))\bigr{)}, (ii) 1Λ¯ϕ1=|F(h0,θ(q))|\sqrt{-1}\Lambda\overline{\partial}\partial\phi_{1}=\bigl{|}F(h_{0},\theta(q))\bigr{|}.

By Corollary 3.6, Lemma 5.15 and Lemma 5.26, there exist Ci(2)C^{(2)}_{i} (i=20,21)(i=20,21), which are independent of YY, such that the following holds on YS0(g,R(0),B(0),L(0))𝒞0Y\cap S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\setminus\mathcal{C}_{0}:

logTr(sY)2Re𝔞(f,θ0)+C20(2)|g(w)|ϵ0+C21(2).\log\mathop{\rm Tr}\nolimits(s_{Y})\leq-2\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{0})+C^{(2)}_{20}|g(w)|^{\epsilon_{0}}+C^{(2)}_{21}. (114)

Here, θ0>0\theta_{0}>0 is sufficiently small. There exists β\beta\in{\mathbb{C}}^{\ast} such that Re(βg(w)ϵ0)>0\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})>0 on {Img(w)B0}\{\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}. Then, there exist Ci(2)C^{(2)}_{i} (i=22,23)(i=22,23) which are independent of YY, such that the following inequality holds on YS0(g,R(0),B(0),L(0))𝒞0Y\cap S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\setminus\mathcal{C}_{0}:

logTr(sY)ϕ12Re𝔞(f,θ0)+C22(2)Re(βg(w)ϵ0)+C23(2).\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}\leq-2\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{0})+C^{(2)}_{22}\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})+C^{(2)}_{23}. (115)

Let 𝒟\mathcal{D} be a connected component of 𝒞0\mathcal{C}_{0} such that

𝒟S0(g,R1(0),B1(0),L1(0)).\mathcal{D}\cap S_{0}(g,R^{(0)}_{1},B^{(0)}_{1},L^{(0)}_{1})\neq\emptyset.

Then, 𝒟\mathcal{D} is relatively compact in YS0(g,R(0),B(0),L(0))Y\cap S_{0}(g,R^{(0)},B^{(0)},L^{(0)}). Because logTr(sY)ϕ1\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1} is subharmonic on YS0(g,R(0),B(0),L(0))Y\cap S_{0}(g,R^{(0)},B^{(0)},L^{(0)}), the estimate (115) holds on 𝒟\mathcal{D}. Hence, (115) holds on the closure of YS0(g,R1(0),B1(0),L1(0))Y\cap S_{0}(g,R^{(0)}_{1},B^{(0)}_{1},L_{1}^{(0)}). Similarly, there exist Ci(2)>0C^{(2)}_{i}>0 (i=24,25)(i=24,25) which are independent of YY such that the following holds on the closure of YSπ(g,R1(0),B1(0),L1(0))Y\cap S_{\pi}(g,R^{(0)}_{1},B_{1}^{(0)},L^{(0)}_{1}):

logTr(sY)ϕ12Re𝔞(f,θπ)+C24(2)Re(βg(w)ϵ0)+C25(2).\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}\leq-2\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{\pi})+C^{(2)}_{24}\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})+C^{(2)}_{25}. (116)

Here πθπ>0\pi-\theta_{\pi}>0 is sufficiently small.

As in Proposition 3.15, there exists C26(2)>0C^{(2)}_{26}>0 which is independent of YY such that the following holds on {wX(R0,δ0)|Img(w)=B1(0),|g(w)|2R1(0)}Y\bigl{\{}w\in X(R_{0},\delta_{0})\,\big{|}\,\mathop{\rm Im}\nolimits g(w)=B^{(0)}_{1},\,\,|g(w)|\leq 2R^{(0)}_{-1}\bigr{\}}\cap Y:

logTr(sY)ϕ1C26(2).\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}\leq C^{(2)}_{26}. (117)

There exists γ\gamma\in{\mathbb{C}}^{\ast} such that Re(γg(w)1ϵ0)>0\mathop{\rm Re}\nolimits(\gamma g(w)^{1-\epsilon_{0}})>0 on

{wX(R0,δ0)|Img(w)B0}.\{w\in X(R_{0},\delta_{0})\,|\,\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}.

There exists C27(2)>0C^{(2)}_{27}>0 which is independent of YY such that the following holds on Y{Img(w)=B1(0)}Y\cap\{\mathop{\rm Im}\nolimits g(w)=B^{(0)}_{1}\}:

|Re𝔞(f,θ0)|+|Re𝔞(f,θπ)|+|Re(βg(w)ϵ0)|C27(2)Re(γg(w)1ϵ0).\bigl{|}\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{0})\bigr{|}+\bigl{|}\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{\pi})\bigr{|}+\bigl{|}\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})\bigr{|}\leq C^{(2)}_{27}\mathop{\rm Re}\nolimits(\gamma g(w)^{1-\epsilon_{0}}). (118)

By (115), (116), (117) and (118), there exist Ci(2)>0C^{(2)}_{i}>0 (i=28,29)(i=28,29), which are independent of YY, such that the following holds on Y{Img(w)=B1(0)}Y\cap\{\mathop{\rm Im}\nolimits g(w)=B^{(0)}_{1}\}:

logTr(sY)ϕ1C28(2)Re(γg(w)1ϵ0)+C29(2)=:ϕ2.\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}\leq C^{(2)}_{28}\mathop{\rm Re}\nolimits(\gamma g(w)^{1-\epsilon_{0}})+C^{(2)}_{29}=:\phi_{2}.

We obtain the following on Y{Img(w)B0}Y\cap\{\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}:

ww¯(logTr(sY)ϕ1ϕ2)0.-\partial_{w}\partial_{\overline{w}}\bigl{(}\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}-\phi_{2}\bigr{)}\leq 0.

Note that logTr(sY)=logr\log\mathop{\rm Tr}\nolimits(s_{Y})=\log r on Y\partial Y, and hence

logTr(sY)ϕ1ϕ2logTr(sY)logr\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}-\phi_{2}\leq\log\mathop{\rm Tr}\nolimits(s_{Y})\leq\log r

on Y{Img(w)B1(0)}\partial Y\cap\{\mathop{\rm Im}\nolimits g(w)\geq B^{(0)}_{1}\}. We also obtain

logTr(sY)ϕ1ϕ20logr\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}-\phi_{2}\leq 0\leq\log r

on Y{Img(w)=B1(0)}Y\cap\{\mathop{\rm Im}\nolimits g(w)=B^{(0)}_{1}\}. Hence, we obtain

logTr(sY)ϕ1ϕ2logr\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi_{1}-\phi_{2}\leq\log r

on (Y{Img(w)B1(0)})\partial\bigl{(}Y\cap\{\mathop{\rm Im}\nolimits g(w)\geq B^{(0)}_{1}\}\bigr{)}. Hence, we obtain logTr(sY)logr+ϕ1+ϕ2\log\mathop{\rm Tr}\nolimits(s_{Y})\leq\log r+\phi_{1}+\phi_{2} on Y{Img(w)B1(0)}Y\cap\{\mathop{\rm Im}\nolimits g(w)\geq B^{(0)}_{1}\}. Because ϕ1=O(log(1+|g(w)|))\phi_{1}=O\bigl{(}\log(1+|g(w)|)\bigr{)}, there exists C30(2)>0C^{(2)}_{30}>0 such that ϕ1+ϕ2+logrC30(2)Re(γg(w)1ϵ0)=:ϕ\phi_{1}+\phi_{2}+\log r\leq C^{(2)}_{30}\mathop{\rm Re}\nolimits(\gamma g(w)^{1-\epsilon_{0}})=:\phi on {Img(w)B0}\{\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}. Thus, we obtain Lemma 5.27. ∎

5.6.5. Construction of harmonic metrics

Proposition 5.28.

For any 𝐚𝒫{\boldsymbol{a}}\in\mathcal{P}, there exists hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that 𝐚(h)=𝐚{\boldsymbol{a}}(h)={\boldsymbol{a}}.

Proof   For 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}, we take an auxiliary metric h0h_{0} as in Lemma 5.25. Let Y1Y2Y_{1}\subset Y_{2}\subset\cdots be a smooth exhaustive family of X(R0,δ0)X(R_{0},\delta_{0}) satisfying Condition 5.17. According to Proposition 2.1, we obtain a sequence hYiHarm(q|Yi)h_{Y_{i}}\in\mathop{\rm Harm}\nolimits(q_{|Y_{i}}) such that hYi|Yi=h0|Yih_{Y_{i}|\partial Y_{i}}=h_{0|\partial Y_{i}}. According to Proposition 2.6, by taking a subsequence, we may assume that the sequence hYih_{Y_{i}} is convergent to hHarm(q)h_{\infty}\in\mathop{\rm Harm}\nolimits(q). Lemma 5.25 and Lemma 5.27 imply 𝒂(h)=𝒂{\boldsymbol{a}}(h_{\infty})={\boldsymbol{a}}. Thus, the proof of Proposition 5.28 is completed. ∎

5.6.6. Comparison with harmonic metrics

Let B0<B1(0)<B0(0)<B1(0)B_{0}<B^{(0)}_{-1}<B^{(0)}_{0}<B^{(0)}_{1} be as in §5.5.4. We can prove the following proposition by the argument of Lemma 5.27.

Proposition 5.29.

There exist a harmonic function ϕ\phi on {wX(R0,δ0)|Img(w)B0}\{w\in X(R_{0},\delta_{0})\,|\,\mathop{\rm Im}\nolimits g(w)\geq B_{0}\} such that the following holds:

  • ϕ=O(|g(w)|1ϵ0)\phi=O\bigl{(}|g(w)|^{1-\epsilon_{0}}\bigr{)}.

  • Let hHarm(q|{Img(w)>B0})h\in\mathop{\rm Harm}\nolimits(q_{|\{\mathop{\rm Im}\nolimits g(w)>B_{0}\}}). Let YY be a relatively compact subset of X(R0,δ0)X(R_{0},\delta_{0}) satisfying Condition 5.17. Suppose that hYHarm(q|Y)h_{Y}\in\mathop{\rm Harm}\nolimits(q_{|Y}) satisfies

    hY|Y{Im(g(w))>B0}=h|Y{Img(w)>B0}.h_{Y|\partial Y\cap\{\mathop{\rm Im}\nolimits(g(w))>B_{0}\}}=h_{|\partial Y\cap\{\mathop{\rm Im}\nolimits g(w)>B_{0}\}}.

    Let sYs_{Y} be the automorphism of 𝕂Y{Img(w)>B0},r\mathbb{K}_{Y\cap\{\mathop{\rm Im}\nolimits g(w)>B_{0}\},r} determined by hY|Y{Img(w)>B0}=h|Y{Img(w)>B0}sYh_{Y|Y\cap\{\mathop{\rm Im}\nolimits g(w)>B_{0}\}}=h_{|Y\cap\{\mathop{\rm Im}\nolimits g(w)>B_{0}\}}\cdot s_{Y}. Then, we obtain logTr(sY)ϕ\log\mathop{\rm Tr}\nolimits(s_{Y})\leq\phi on Y{Im(g(w))>B1(0)}Y\cap\{\mathop{\rm Im}\nolimits(g(w))>B^{(0)}_{1}\}.

Proof   In this proof, constants are independent of hh and YY. Let hh, hYHarm(q|Y)h_{Y}\in\mathop{\rm Harm}\nolimits(q_{|Y}) and sYs_{Y} be as in the statement. By Proposition 3.7, there exists a constant C1>0C_{1}>0 such that the following holds on {wX(R0,δ0)|Img(w)B1(0)}\{w\in X(R_{0},\delta_{0})\,|\,\mathop{\rm Im}\nolimits g(w)\geq B^{(0)}_{-1}\}:

|θ(q)|hC1(Img(w))1.|\theta(q)|_{h}\leq C_{1}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-1}.

By the same argument as the proof of Lemma 5.26, we can prove that there exists a constant C2>0C_{2}>0 such that the following inequality holds on YS0(g,R(0),B(0),L(0))Y\cap S_{0}(g,R^{(0)},B^{(0)},L^{(0)}):

|θ(q)|hYC2(Img(w))1.|\theta(q)|_{h_{Y}}\leq C_{2}\bigl{(}\mathop{\rm Im}\nolimits g(w)\bigr{)}^{-1}.

By Corollary 3.5 and Lemma 5.15, there exist constants Ci>0C_{i}>0 (i=3,4)(i=3,4) such that the following inequality holds on YS0(g,R(0),B(0),L(0))𝒞0Y\cap S_{0}(g,R^{(0)},B^{(0)},L^{(0)})\setminus\mathcal{C}_{0}:

logTr(sY)2Re𝔞(f,θ0)+C3Re(βg(w)ϵ0)+C4.\log\mathop{\rm Tr}\nolimits(s_{Y})\leq-2\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{0})+C_{3}\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})+C_{4}. (119)

Here, θ0>0\theta_{0}>0 is sufficiently small, and β\beta is a complex number such that Re(βg(w)ϵ0)>0\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})>0 on {Img(w)B0}\{\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}. Note that logTr(sY)\log\mathop{\rm Tr}\nolimits(s_{Y}) is subharmonic by (21). Hence, the inequality (119) holds on the closure of YS0(g,R1(0),B1(0),L1(0))Y\cap S_{0}(g,R^{(0)}_{1},B^{(0)}_{1},L^{(0)}_{1}). Similarly, there exist constants Ci>0C_{i}>0 (i=5,6)(i=5,6) such that the following holds on the closure of YSπ(g,R1(0),B1(0),L1(0))Y\cap S_{\pi}(g,R^{(0)}_{1},B^{(0)}_{1},L^{(0)}_{1}):

logTr(sY)2Re𝔞(f,θπ)+C5Re(βg(w)ϵ0)+C6.\log\mathop{\rm Tr}\nolimits(s_{Y})\leq-2\mathop{\rm Re}\nolimits\mathfrak{a}(f,\theta_{\pi})+C_{5}\mathop{\rm Re}\nolimits(\beta g(w)^{\epsilon_{0}})+C_{6}. (120)

Here, πθπ>0\pi-\theta_{\pi}>0 is sufficiently small. As in Proposition 3.15, there exists a constant C7>0C_{7}>0 such that logTr(sY)C7\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C_{7} on {Img(w)=B1(0),|g(w)|2R1(0)}Y\{\mathop{\rm Im}\nolimits g(w)=B^{(0)}_{1},\,\,|g(w)|\leq 2R^{(0)}_{-1}\}\cap Y. Therefore, there exist positive constants CiC_{i} (i=8,9)(i=8,9) such that the following holds on Y{Img(w)=B1(0)}Y\cap\{\mathop{\rm Im}\nolimits g(w)=B^{(0)}_{1}\}:

logTr(sY)C8Re(γg(w)1ϵ0)+C9=:ϕ0.\log\mathop{\rm Tr}\nolimits(s_{Y})\leq C_{8}\mathop{\rm Re}\nolimits(\gamma g(w)^{1-\epsilon_{0}})+C_{9}=:\phi_{0}.

Here γ\gamma is a complex number such that Re(γg(w)1ϵ0)>0\mathop{\rm Re}\nolimits(\gamma g(w)^{1-\epsilon_{0}})>0 on {Img(w)B0}\{\mathop{\rm Im}\nolimits g(w)\geq B_{0}\}.

We set ϕ=ϕ0+log(r)\phi=\phi_{0}+\log(r). Note that logTr(sY)ϕ\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi is a subharmonic function. By the construction, we obtain logTr(sY)ϕ0\log\mathop{\rm Tr}\nolimits(s_{Y})-\phi\leq 0 on (Y{Img(w)B1(0)})\partial(Y\cap\{\mathop{\rm Im}\nolimits g(w)\geq B_{1}^{(0)}\}), and hence on Y{Img(w)B1(0)}Y\cap\{\mathop{\rm Im}\nolimits g(w)\geq B_{1}^{(0)}\}. Thus, we obtain Proposition 5.29. ∎

5.6.7. Proof of Theorem 5.3 and Proposition 5.4

We consider the map Φ:X(R,δ)\Phi:X(R,\delta)\longrightarrow{\mathbb{C}} defined by w=1αzρw=-\sqrt{-1}\alpha z^{-\rho}. If RR is sufficiently large and δ\delta is sufficiently small, we obtain Φ:X(R,δ)V\Phi:X(R,\delta)\longrightarrow V, and it is a holomorphic embedding. By applying Proposition 5.20 and Proposition 5.24 to Φ(q)=f(dw)r\Phi^{\ast}(q)=f\cdot(dw)^{r}, we obtain the first two claims of Theorem 5.3.

Let 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P}. By Proposition 5.28, there exists h0,𝒂Harm(Φ(q))h_{0,{\boldsymbol{a}}}\in\mathop{\rm Harm}\nolimits(\Phi^{\ast}(q)) such that 𝒂(h0,𝒂)=𝒂{\boldsymbol{a}}(h_{0,{\boldsymbol{a}}})={\boldsymbol{a}}. It induces h0,𝒂Harm(q|Φ(X(R,δ)))h_{0,{\boldsymbol{a}}}\in\mathop{\rm Harm}\nolimits(q_{|\Phi(X(R,\delta))}). We extend it to a GrG_{r}-invariant Hermitian metric of 𝕂Vϖ1(0),r\mathbb{K}_{V\setminus\varpi^{-1}(0),r} such that det(h1,𝒂)=1\det(h_{1,{\boldsymbol{a}}})=1. Let {Xi}\{X_{i}\} be a smooth exhaustive family of VV such that each Φ1(Xi)\Phi^{-1}(X_{i}) satisfies Condition 5.17. Let hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) such that hi|Xi=h1,𝒂|Xih_{i|\partial X_{i}}=h_{1,{\boldsymbol{a}}|\partial X_{i}}. According to Proposition 3.15, we may assume that the sequence hih_{i} is convergent, and we obtain hHarm(q)h_{\infty}\in\mathop{\rm Harm}\nolimits(q) as the limit of a subsequence of hih_{i}. By Proposition 5.29, hh_{\infty} satisfies 𝒂I(h)=𝒂{\boldsymbol{a}}_{I}(h_{\infty})={\boldsymbol{a}}. Thus, we obtain the first claim of Theorem 5.3.

For the proof of Proposition 5.4, we use the notation in §5.5.4. We set 𝒰I,1=Φ({Im(g(w))>B0})\mathcal{U}_{I,1}=\Phi(\{\mathop{\rm Im}\nolimits(g(w))>B_{0}\}) and 𝒰I,2=Φ({Im(g(w))>B1(0)})\mathcal{U}_{I,2}=\Phi(\{\mathop{\rm Im}\nolimits(g(w))>B_{1}^{(0)}\}). Let {Ki}\{K_{i}\} be a smooth exhaustive family of Vϖ1(0)V\setminus\varpi^{-1}(0) such that each Φ1(Ki)\Phi^{-1}(K_{i}) satisfies Condition 5.17. Let ϕI\phi_{I} denote the function on 𝒰I,1\mathcal{U}_{I,1} induced by ϕ\phi in Proposition 5.29. Then, we obtain Proposition 5.4 from Proposition 5.29. ∎

5.7. Refined estimate in an easy case

Let ff be a section of 𝔅\mathfrak{B} on V~V\subset\widetilde{{\mathbb{C}}}. Let II be an interval of Vϖ1(0)V\cap\varpi^{-1}(0) such that I¯V\overline{I}\subset V and that II is special with respect to ff. We assume the following.

Condition 5.30.

There exist a neighbourhood 𝒰\mathcal{U} of I¯\overline{I} in VV, a finite sum 𝔞I=0<κρ𝔞I,κzκ\mathfrak{a}_{I}=\sum_{0<\kappa\leq\rho}\mathfrak{a}_{I,\kappa}z^{\kappa} (𝔞I,ρ0)(\mathfrak{a}_{I,\rho}\neq 0), α\alpha\in{\mathbb{C}}^{\ast}, I\ell_{I}\in{\mathbb{R}} and C>1C>1 such that |e𝔞IzI(fαe𝔞IzI)|C|z|ϵ\bigl{|}e^{-\mathfrak{a}_{I}}z^{-\ell_{I}}(f-\alpha e^{\mathfrak{a}_{I}}z^{\ell_{I}})\bigr{|}\leq C|z|^{\epsilon} for ϵ>0\epsilon>0 on 𝒰ϖ1(0)\mathcal{U}\setminus\varpi^{-1}(0).

Recall deg(𝔞I)=ρ\deg(\mathfrak{a}_{I})=\rho. We set q=f(dz/z)rq=f\,(dz/z)^{r} and

𝔡I(q):=Ideg(𝔞I)+r.\mathfrak{d}_{I}(q):=\frac{\ell_{I}}{\deg(\mathfrak{a}_{I})}+r.

Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). We obtain 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} as in Theorem 5.3. In this easier case, we can obtain the estimate on the behaviour of hh up to boundedness. Let 𝒌I(h){\boldsymbol{k}}_{I}(h) be the tuple of integers obtained from 𝒂I(h){\boldsymbol{a}}_{I}(h) as in §3.6.1. We shall prove the following proposition in §5.7.15.7.2.

Proposition 5.31.

The following estimates holds as Re(𝔞I(z))+𝔡I(q)log|𝔞I(z)|-\mathop{\rm Re}\nolimits(\mathfrak{a}_{I}(z))+\mathfrak{d}_{I}(q)\log|\mathfrak{a}_{I}(z)|\to\infty:

log|(dz)(r+12i)/2|haI,i(h)(Re(𝔞I(z))+𝔡I(q)log|𝔞I(z)|)+(deg(𝔞I)+1)(r+12i)log|z|kI,i2log(Re(𝔞I(z))+𝔡I(f)log|𝔞I(z)|)=O(1).\log\Bigl{|}(dz)^{(r+1-2i)/2}\Bigr{|}_{h}-a_{I,i}(h)\Bigl{(}-\mathop{\rm Re}\nolimits(\mathfrak{a}_{I}(z))+\mathfrak{d}_{I}(q)\log|\mathfrak{a}_{I}(z)|\Bigr{)}\\ +(\deg(\mathfrak{a}_{I})+1)\left(\frac{r+1}{2}-i\right)\log|z|\\ -\frac{k_{I,i}}{2}\log\Bigl{(}-\mathop{\rm Re}\nolimits(\mathfrak{a}_{I}(z))+\mathfrak{d}_{I}(f)\log|\mathfrak{a}_{I}(z)|\Bigr{)}=O(1). (121)

5.7.1. Normalization

Let 𝒰\mathcal{U} denote an open neighbourhood of I¯\overline{I}. We define the map F:𝒰ϖ1(0)F:\mathcal{U}\setminus\varpi^{-1}(0)\longrightarrow{\mathbb{C}} by

F(z)=1(𝔞I(z)𝔡I(q)log(1𝔞I(z))).F(z)=-\sqrt{-1}\Bigl{(}\mathfrak{a}_{I}(z)-\mathfrak{d}_{I}(q)\log\bigl{(}-\sqrt{-1}\mathfrak{a}_{I}(z)\bigr{)}\Bigr{)}.

Let ϖ:~11\varpi_{\infty}:\widetilde{\mathbb{P}}^{1}_{\infty}\longrightarrow\mathbb{P}^{1} denote the oriented real blowing up at \infty. We regard {\mathbb{C}} as an open subset of ~1\widetilde{\mathbb{P}}^{1}_{\infty}. Let ww denote the standard coordinate system on {\mathbb{C}}. By using the polar decomposition of ww, we identify ~1{0}\widetilde{\mathbb{P}}^{1}_{\infty}\setminus\{0\} with (>0{})×S1({\mathbb{R}}_{>0}\cup\{\infty\})\times S^{1}.

Lemma 5.32.

There exist a neighbourhood 𝒰\mathcal{U}^{\prime} of I¯\overline{I} in 𝒰\mathcal{U} and a neighbourhood 𝒱\mathcal{V} of {}×[0,π]\{\infty\}\times[0,\pi] in ~1\widetilde{\mathbb{P}}^{1}_{\infty} such that FF induces a homeomorphism 𝒰𝒱\mathcal{U}^{\prime}\simeq\mathcal{V}.

Proof   It is easy to see that FF extends to a continuous map 𝒰~1\mathcal{U}\longrightarrow\widetilde{\mathbb{P}}^{1}_{\infty}, which is also denoted by FF. Moreover, we can easily check that FF induces a homeomorphism of a neighbourhood of I¯\overline{I} in ϖ1(0)\varpi^{-1}(0) and a neighbourhood of {}×[0,π]\{\infty\}\times[0,\pi] in ϖ1()\varpi_{\infty}^{-1}(\infty).

We set ρ:=deg(𝔞I)\rho:=\deg(\mathfrak{a}_{I}). On a neighbourhood of I¯\overline{I}, we use the real coordinate system given by (|z|ρ,arg(z))(|z|^{\rho},\arg(z)). On a neighbourhood of {}×[0,π]\{\infty\}\times[0,\pi], we use the real coordinate system given by (|w|1,arg(w))(|w|^{-1},\arg(w)). Then, it is easy to check that FF is of C1C^{1}-class, and the tangent map at each point of I¯\overline{I} is an isomorphism. Then, the claim follows from the inverse function theorem. ∎

Then, (F1)(q)(F^{-1})^{\ast}(q) is expressed as follows:

(F1)(q)=βe1w(1+υ(w))(dw)r.(F^{-1})^{\ast}(q)=\beta e^{\sqrt{-1}w}(1+\upsilon(w))\,(dw)^{r}.

Here, υ\upsilon is a holomorphic function satisfying υ(w)=O(|w|ϵ)\upsilon(w)=O(|w|^{-\epsilon}) for some ϵ>0\epsilon>0, and β\beta is a non-zero complex number.

5.7.2. Comparison with the model metric

Let δ>0\delta>0 and R>0R>0. Let f0f_{0} be a nowhere vanishing holomorphic function on X(R,δ)X(R,\delta) such that log|f0|=Imw+O(1)\log|f_{0}|=-\mathop{\rm Im}\nolimits w+O(1). We set q0:=f0(dw)rq_{0}:=f_{0}\,(dw)^{r}. According to Theorem 5.3, for any h0Harm(q0)h_{0}\in\mathop{\rm Harm}\nolimits(q_{0}), there exist 𝒂(h0)𝒫{\boldsymbol{a}}(h_{0})\in\mathcal{P} and 0<ρ<10<\rho<1 such that the following estimates hold on {|w|>R,|arg(w)π/2|<(1δ1)π/2}\bigl{\{}|w|>R,\,\,\big{|}\arg(w)-\pi/2|<(1-\delta_{1})\pi/2\bigr{\}} for any 0<δ1<10<\delta_{1}<1:

log|(dw)(r+12i)/2|h0ai(h0)Im(w)=O(|w|ρ).\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h_{0}}-a_{i}(h_{0})\mathop{\rm Im}\nolimits(w)=O\bigl{(}|w|^{\rho}\bigr{)}.

Note that log|f0|=O(1)\log|f_{0}|=O(1) on {wX(R+1,δ)||Im(w)|<A}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,|\mathop{\rm Im}\nolimits(w)|<A\bigr{\}} for any A>0A>0 in this case. By Corollary 3.8, we obtain that log|(dw)(r+12i)/2|h0\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h_{0}} is bounded on {wX(R+1,δ)||Im(w)|<A}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,|\mathop{\rm Im}\nolimits(w)|<A\bigr{\}} for any A>0A>0. By Proposition 4.4, we obtain the following estimate on {wX(R+1,δ)|Im(w)1}\{w\in X(R+1,\delta)\,|\,\mathop{\rm Im}\nolimits(w)\geq 1\}:

log|(dw)(r+12i)/2|h0ai(h0)Im(w)=O(log(1+Im(w))).\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h_{0}}-a_{i}(h_{0})\mathop{\rm Im}\nolimits(w)=O\bigl{(}\log(1+\mathop{\rm Im}\nolimits(w))\bigr{)}.

Let f1f_{1} be a holomorphic function on X(R,δ)X(R,\delta) such that there exist C1>0C_{1}>0 and κ>0\kappa>0 such that

|f0f1|C1|w|κ|f0|.|f_{0}-f_{1}|\leq C_{1}|w|^{-\kappa}|f_{0}|.

We obtain log|f1|=Im(w)+O(1)\log|f_{1}|=-\mathop{\rm Im}\nolimits(w)+O(1). We set q1:=f1(dw)rq_{1}:=f_{1}\,(dw)^{r}. Similarly, for h1Harm(q1)h_{1}\in\mathop{\rm Harm}\nolimits(q_{1}), there exists 𝒂(h1)𝒫{\boldsymbol{a}}(h_{1})\in\mathcal{P} such that the following estimate holds on {wX(R+1,δ)|Im(w)1}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,\mathop{\rm Im}\nolimits(w)\geq 1\bigr{\}}:

log|(dw)(r+12i)/2|h1ai(h1)Im(w)=O(log(1+Im(w))).\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h_{1}}-a_{i}(h_{1})\mathop{\rm Im}\nolimits(w)=O\bigl{(}\log(1+\mathop{\rm Im}\nolimits(w))\bigr{)}.
Proposition 5.33.

Suppose that there exist h0Harm(q0)h_{0}\in\mathop{\rm Harm}\nolimits(q_{0}) and h1Harm(q1)h_{1}\in\mathop{\rm Harm}\nolimits(q_{1}) such that 𝐚(h0)=𝐚(h1){\boldsymbol{a}}(h_{0})={\boldsymbol{a}}(h_{1}). Then, h0h_{0} and h1h_{1} are mutually bounded on {wX(R+1,δ)|Im(w)1}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,\mathop{\rm Im}\nolimits(w)\geq 1\bigr{\}}.

Proof   We set ei:=(dw)(r+12i)/2e_{i}:=(dw)^{(r+1-2i)/2} for i=1,,ri=1,\ldots,r. Note that θ(q1)ei=θ(q0)ei\theta(q_{1})e_{i}=\theta(q_{0})e_{i} (i=1,,r)(i=1,\ldots,r), θ(q1)er=θ(q0)er+(f1f0)e1dw\theta(q_{1})e_{r}=\theta(q_{0})e_{r}+(f_{1}-f_{0})e_{1}\,dw and θ(q0)(er)=f0e1dw\theta(q_{0})(e_{r})=f_{0}\,e_{1}\,dw. Because |f1f0|C1|w|κ|f0||f_{1}-f_{0}|\leq C_{1}|w|^{-\kappa}|f_{0}|, there exists C2>0C_{2}>0 such that |θ(q1)θ(q0)|h0C2|w|κ|θ(q0)|h0|\theta(q_{1})-\theta(q_{0})|_{h_{0}}\leq C_{2}|w|^{-\kappa}|\theta(q_{0})|_{h_{0}}. Hence, on {wX(R+1,δ)|Im(w)1}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,\mathop{\rm Im}\nolimits(w)\geq 1\bigr{\}}, we obtain |θ(q1)θ(q0)|h0=O((1+Im(w))1κ)|\theta(q_{1})-\theta(q_{0})|_{h_{0}}=O\bigl{(}(1+\mathop{\rm Im}\nolimits(w))^{-1-\kappa}\bigr{)} and |θ(q1)|h=O((1+Im(w))1)|\theta(q_{1})|_{h}=O\bigl{(}(1+\mathop{\rm Im}\nolimits(w))^{-1}\bigr{)}. Because R(h0)+[θ(q0),θ(q0)h0]=0R(h_{0})+\bigl{[}\theta(q_{0}),\theta(q_{0})_{h_{0}}^{\dagger}\bigr{]}=0, we obtain

|R(h0)+[θ(q1),θ(q1)h0]|h0=O((1+Im(w))2κ).\bigl{|}R(h_{0})+\bigl{[}\theta(q_{1}),\theta(q_{1})_{h_{0}}^{\dagger}\bigr{]}\bigr{|}_{h_{0}}=O\bigl{(}(1+\mathop{\rm Im}\nolimits(w))^{-2-\kappa}\bigr{)}.

Let ss be the automorphism of 𝕂X(R,δ),r\mathbb{K}_{X(R,\delta),r} determined by h1=h0sh_{1}=h_{0}s. By (21), we obtain the following on {wX(R+1,δ)|Im(w)0}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,\mathop{\rm Im}\nolimits(w)\geq 0\bigr{\}}:

2zz¯logTr(s)|R(h0)+[θ(q1),θ(q1)h0]|h0=O((1+Im(w))2κ).-2\partial_{z}\partial_{\overline{z}}\log\mathop{\rm Tr}\nolimits(s)\leq\bigl{|}R(h_{0})+\bigl{[}\theta(q_{1}),\theta(q_{1})_{h_{0}}^{\dagger}\bigr{]}\bigr{|}_{h_{0}}=O\bigl{(}(1+\mathop{\rm Im}\nolimits(w))^{-2-\kappa}\bigr{)}.

Note that Tr(s)\mathop{\rm Tr}\nolimits(s) is bounded on {wX(R+1,δ)||Im(w)|<A}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,|\mathop{\rm Im}\nolimits(w)|<A\bigr{\}}. By Proposition 4.4, there exists aa\in{\mathbb{R}} such that |logTr(s)aIm(w)||\log\mathop{\rm Tr}\nolimits(s)-a\mathop{\rm Im}\nolimits(w)| is bounded on {wX(R+1,δ)|Im(w)0}\bigl{\{}w\in X(R+1,\delta)\,\big{|}\,\mathop{\rm Im}\nolimits(w)\geq 0\bigr{\}}. Because 𝒂(h0)=𝒂(h1){\boldsymbol{a}}(h_{0})={\boldsymbol{a}}(h_{1}), we obtain a=0a=0, which implies the boundedness of logTr(s)\log\mathop{\rm Tr}\nolimits(s). ∎

Let υ(w)\upsilon(w) be a holomorphic function on X(R,δ)X(R,\delta) such that υ(w)=O(|w|κ)\upsilon(w)=O(|w|^{-\kappa}). Let α\alpha be a non-zero complex number. We set q=α(1+υ)e1w(dw)rq=\alpha(1+\upsilon)e^{\sqrt{-1}w}(dw)^{r}. Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). We obtain 𝒂(h)𝒫{\boldsymbol{a}}(h)\in\mathcal{P} and a tuple of integers 𝒌(h){\boldsymbol{k}}(h) as in §3.6.1.

Corollary 5.34.

We obtain the following estimates for i=1,,ri=1,\ldots,r on {wX(R,δ)|Im(w)0}\bigl{\{}w\in X(R,\delta)\,\big{|}\,\mathop{\rm Im}\nolimits(w)\geq 0\bigr{\}}:

log|(dw)(r+12i)/2|hai(h)Im(w)ki(h)2log(1+Im(w))=O(1).\log\bigl{|}(dw)^{(r+1-2i)/2}\bigr{|}_{h}-a_{i}(h)\mathop{\rm Im}\nolimits(w)-\frac{k_{i}(h)}{2}\log\bigl{(}1+\mathop{\rm Im}\nolimits(w)\bigr{)}=O(1).

Proof   In the case υ=0\upsilon=0, the claim follows from Proposition 3.25 and Theorem 5.3. The general case follows from Proposition 5.33. ∎

We obtain Proposition 5.31 by applying the normalization in §5.7.1 and Corollary 5.34. ∎

6. Global results

6.1. Holomorphic rr-differentials with multiple growth orders on a punctured disc

6.1.1. General case

Let UU be a neighbourhood of 0 in {\mathbb{C}}. Let ϖ:U~U\varpi:\widetilde{U}\longrightarrow U denote the oriented real blowing up. Let ff be a section of 𝔅U~\mathfrak{B}_{\widetilde{U}} on U~\widetilde{U}. Let 𝒮(f)\mathcal{S}(f) denote the set of the intervals which are special with respect to ff. For any I𝒮(f)I\in\mathcal{S}(f), we fix a branch of logz\log z around II. Then, ρ(I)>0\rho(I)\in{\mathbb{R}}_{>0} and αI\alpha_{I}\in{\mathbb{C}}^{\ast} are determined as

deg(𝔞(f,θ)αIzρ(I))<ρ(I)\deg(\mathfrak{a}(f,\theta)-\alpha_{I}z^{-\rho(I)})<\rho(I)

for θI𝒵(f)\theta\in I\setminus\mathcal{Z}(f). We set q=f(dz)rq=f\,(dz)^{r}. We set 𝒮(q):=𝒮(f)\mathcal{S}(q):=\mathcal{S}(f). Let U0UU_{0}\subset U be a relatively compact neighbourhood of 0 in UU. We set U=U{0}U^{\ast}=U\setminus\{0\} and U0=U0{0}U_{0}^{\ast}=U_{0}\setminus\{0\}.

Theorem 6.1.

  • For any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) and for any I𝒮(q)I\in\mathcal{S}(q), there exist 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} and ϵ>0\epsilon>0 such that the following estimates hold as |z|0|z|\to 0 on {|arg(αIzρ(I))π|<(1δ)π/2}\bigl{\{}|\arg(\alpha_{I}z^{-\rho(I)})-\pi|<(1-\delta)\pi/2\bigr{\}} for any δ>0\delta>0:

    log|(dz)(r+1)/2i|h+ai(h)Re(αIzρ(I))=O(|z|ρ(I)+ϵ).\log\bigl{|}(dz)^{(r+1)/2-i}\bigr{|}_{h}+a_{i}(h)\mathop{\rm Re}\nolimits\bigl{(}\alpha_{I}z^{-\rho(I)}\bigr{)}=O\bigl{(}|z|^{-\rho(I)+\epsilon}\bigr{)}.
  • If hiHarm(q)h_{i}\in\mathop{\rm Harm}\nolimits(q) (i=1,2)(i=1,2) satisfy 𝒂I(h1)=𝒂I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) for any I𝒮(q)I\in\mathcal{S}(q), then h1h_{1} and h2h_{2} are mutually bounded on U0U_{0}^{\ast}.

  • For any (𝒂I)I𝒮(q)I𝒮(q)𝒫({\boldsymbol{a}}_{I})_{I\in\mathcal{S}(q)}\in\prod_{I\in\mathcal{S}(q)}\mathcal{P}, there exists hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) such that 𝒂I(h)=𝒂I{\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}_{I}.

Proof   We obtain the first claims from Theorem 5.3. We obtain the second claim from Proposition 5.1, Theorem 5.2 and Theorem 5.3.

Let us prove the third claim. Let (𝒂I)I𝒮(q)I𝒮(q)𝒫({\boldsymbol{a}}_{I})_{I\in\mathcal{S}(q)}\in\prod_{I\in\mathcal{S}(q)}\mathcal{P}. For each I𝒮(q)I\in\mathcal{S}(q), let VIV_{I} be a relatively compact neighbourhood of I¯\overline{I} in U~\widetilde{U}. There exist relatively compact neighbourhoods 𝒰I,i\mathcal{U}_{I,i} (i=1,2)(i=1,2), a function ϕI\phi_{I}, and a smooth exhaustive family {KI,i}\{K_{I,i}\}, as in Proposition 5.4. By Theorem 5.3, there exists a GrG_{r}-invariant Hermitian metric h0h_{0} of 𝕂U{0},r\mathbb{K}_{U\setminus\{0\},r} such that det(h0)=1\det(h_{0})=1 and that h0|𝒰I,1Harm(q|𝒰I,1)h_{0|\mathcal{U}_{I,1}}\in\mathop{\rm Harm}\nolimits(q_{|\mathcal{U}_{I,1}}) with 𝒂I(h0)=𝒂I{\boldsymbol{a}}_{I}(h_{0})={\boldsymbol{a}}_{I} for I𝒮(q)I\in\mathcal{S}(q). Let {Xi}\{X_{i}\} be a smooth exhaustive family for U{0}U\setminus\{0\} such that (i) IKI,iXi\coprod_{I}K_{I,i}\subset X_{i}, (ii) KI,i𝒰I,1=Xi𝒰I,1K_{I,i}\cap\mathcal{U}_{I,1}=X_{i}\cap\mathcal{U}_{I,1} for any II. We obtain hiHarm(q|Xi)h_{i}\in\mathop{\rm Harm}\nolimits(q_{|X_{i}}) such that hi|Xi=h0|Xih_{i|\partial X_{i}}=h_{0|\partial X_{i}}. We may assume that the sequence {hi}\{h_{i}\} is convergent, and we obtain hHarm(q)h_{\infty}\in\mathop{\rm Harm}\nolimits(q) as the limit. By Proposition 5.4, we obtain 𝒂I(h)=𝒂I{\boldsymbol{a}}_{I}(h_{\infty})={\boldsymbol{a}}_{I} for I𝒮(q)I\in\mathcal{S}(q). ∎

6.1.2. Refined estimate in the nowhere vanishing case

Let us state a refined result in the case where qq is nowhere vanishing.

Lemma 6.2.

Let qq be a holomorphic rr-differential with multiple growth orders on (U,0)(U,0). If qq is nowhere vanishing, then there exist an integer (q)\ell(q) and a meromorphic function 𝔞\mathfrak{a} on (U,0)(U,0) such that q=z(q)e𝔞(z)(dz/z)rq=z^{\ell(q)}e^{\mathfrak{a}(z)}(dz/z)^{r}.

Proof   We describe qq as q=β(dz/z)rq=\beta(dz/z)^{r} for a holomorphic function β\beta. There exists an integer \ell such that 𝔞=log(zβ)\mathfrak{a}=\log(z^{-\ell}\beta) is well defined as a single-valued holomorphic function on UU^{\ast}. Because β\beta induces a section of 𝔅\mathfrak{B}, we obtain Re(𝔞)=O(|z|ρ)\mathop{\rm Re}\nolimits(\mathfrak{a})=O\bigl{(}|z|^{-\rho}\bigr{)} for some ρ>0\rho>0. Because Re(𝔞)\mathop{\rm Re}\nolimits(\mathfrak{a}) is harmonic, there exists ρ1>0\rho_{1}>0 such that xRe(𝔞)=O(|z|ρ1)\partial_{x}\mathop{\rm Re}\nolimits(\mathfrak{a})=O(|z|^{-\rho_{1}}) and yRe(𝔞)=O(|z|ρ1)\partial_{y}\mathop{\rm Re}\nolimits(\mathfrak{a})=O(|z|^{-\rho_{1}}), where (x,y)(x,y) is the real coordinate system induced by z=x+1yz=x+\sqrt{-1}y. By the Cauchy-Riemann equation, we obtain that xIm(𝔞)=O(|z|ρ1)\partial_{x}\mathop{\rm Im}\nolimits(\mathfrak{a})=O(|z|^{-\rho_{1}}) and yIm(𝔞)=O(|z|ρ1)\partial_{y}\mathop{\rm Im}\nolimits(\mathfrak{a})=O(|z|^{-\rho_{1}}). Hence, there exists ρ2>0\rho_{2}>0 such that Im(𝔞)=O(|z|ρ2)\mathop{\rm Im}\nolimits(\mathfrak{a})=O(|z|^{-\rho_{2}}). Thus, we obtain that 𝔞\mathfrak{a} is meromorphic at z=0z=0. ∎

For the description q=z(q)e𝔞(z)(dz/z)rq=z^{\ell(q)}e^{\mathfrak{a}(z)}(dz/z)^{r}, we obtain the integer m(q)m(q) such that zm(q)𝔞(z)z^{m(q)}\mathfrak{a}(z) is holomorphic at z=0z=0, and α:=(zm(q)𝔞(z))z=00\alpha:=(z^{m(q)}\mathfrak{a}(z))_{z=0}\neq 0. We set 𝔡(q):=(q)m(q)+r\mathfrak{d}(q):=\frac{\ell(q)}{m(q)}+r. We also set

T(q):={0θ<2π|m(q)θ+arg(α)π,mod2π}.T(q):=\{0\leq\theta<2\pi\,|\,-m(q)\theta+\arg(\alpha)\equiv\pi,\mod 2\pi\}.

Note that special intervals with respect to qq are {|θθ0|<π/2m(q)}\bigl{\{}|\theta-\theta_{0}|<\pi/2m(q)\bigr{\}} for θ0T(q)\theta_{0}\in T(q).

Proposition 6.3.

Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). For any θ0T(q)\theta_{0}\in T(q), we obtain the tuple of real numbers

𝒂(h,θ0)=(a0(h,θ0),,ar1(h,θ0))𝒫{\boldsymbol{a}}(h,\theta_{0})=(a_{0}(h,\theta_{0}),\ldots,a_{r-1}(h,\theta_{0}))\in\mathcal{P}

determined by the following estimates for i=1,,ri=1,\ldots,r as Re(𝔞(z))+𝔡(q)log|𝔞(z)|-\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(z)\bigr{)}+\mathfrak{d}(q)\log|\mathfrak{a}(z)|\to\infty on {|arg(z)θ0|<π/2m(q)}\bigl{\{}|\arg(z)-\theta_{0}|<\pi/2m(q)\bigr{\}}:

log|(dz)(r+1)/2i|hai(h,θ0)(Re(𝔞(z))+𝔡(q)log|𝔞(z)|)+(m(q)+1)(r+12i)log|z|=O(log(Re(𝔞(z))+𝔡(q)log|𝔞(z)|)).\log\Bigl{|}(dz)^{(r+1)/2-i}\Bigr{|}_{h}-a_{i}(h,\theta_{0})\Bigl{(}-\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(z)\bigr{)}\\ +\mathfrak{d}(q)\log|\mathfrak{a}(z)|\Bigr{)}+(m(q)+1)\left(\frac{r+1}{2}-i\right)\log|z|\\ =O\left(\log\Bigl{(}-\mathop{\rm Re}\nolimits(\mathfrak{a}(z))+\mathfrak{d}(q)\log|\mathfrak{a}(z)|\Bigr{)}\right). (122)

¿From 𝐚(h,θ0)𝒫{\boldsymbol{a}}(h,\theta_{0})\in\mathcal{P}, we determine the integers ki(h,θ0)k_{i}(h,\theta_{0}) (i=0,,r1)(i=0,\ldots,r-1) as in §3.6.1. Then, we obtain the following estimates as Re(𝔞(z))+𝔡(q)log|𝔞(z)|-\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(z)\bigr{)}+\mathfrak{d}(q)\log|\mathfrak{a}(z)|\to\infty on {|arg(z)θ0|<π/2m(q)}\bigl{\{}|\arg(z)-\theta_{0}|<\pi/2m(q)\bigr{\}}:

log|(dz)(r+1)/2i|hai(h,θ0)(Re(𝔞(z))+𝔡(q)log|𝔞(z)|)+(m(q)+1)(r+12i)log|z|ki(h,θ0)2log(Re(𝔞(z))+𝔡(q)log|𝔞(z)|)=O(1).\log\Bigl{|}(dz)^{(r+1)/2-i}\Bigr{|}_{h}-a_{i}(h,\theta_{0})\Bigl{(}-\mathop{\rm Re}\nolimits\bigl{(}\mathfrak{a}(z)\bigr{)}+\mathfrak{d}(q)\log|\mathfrak{a}(z)|\Bigr{)}\\ +(m(q)+1)\left(\frac{r+1}{2}-i\right)\log|z|\\ -\frac{k_{i}(h,\theta_{0})}{2}\log\Bigl{(}-\mathop{\rm Re}\nolimits(\mathfrak{a}(z))+\mathfrak{d}(q)\log|\mathfrak{a}(z)|\Bigr{)}=O\bigl{(}1\bigr{)}. (123)

Proof   It follows from Proposition 5.31. ∎

6.2. The case of punctured Riemann surfaces

6.2.1. Statement

Let XX be a Riemann surface. Let DD be a finite subset of XX. For each PDP\in D, let (XP,zP)(X_{P},z_{P}) be a holomorphic coordinate neighbourhood such that zP(P)=0z_{P}(P)=0. Set XP:=XP{P}X_{P}^{\ast}:=X_{P}\setminus\{P\}. Let ϖP:X~PXP\varpi_{P}:\widetilde{X}_{P}\to X_{P} denote the oriented blow up along PP.

Let qq be an rr-differential on XX. We obtain the expression q|XP=fP(dzP/zP)rq_{|X_{P}^{\ast}}=f_{P}\,(dz_{P}/z_{P})^{r}. Let DmeroD_{\mathop{\rm mero}\nolimits} be the points PP of DD such that fPf_{P} are meromorphic at PP. Let D>0D_{>0} denote the set of PP such that fPf_{P} is holomorphic at PP and that fP(P)=0f_{P}(P)=0. We set D0:=DmeroD>0D_{\leq 0}:=D_{\mathop{\rm mero}\nolimits}\setminus D_{>0}.

For each PD>0P\in D_{>0}, we have the description qP=wPmPfP(dwP/wP)rq_{P}=w_{P}^{m_{P}}f_{P}(dw_{P}/w_{P})^{r}, where fPf_{P} is holomorphic at PP such that fP(P)0f_{P}(P)\neq 0. Let 𝒫(q,P)\mathcal{P}(q,P) denote the set of 𝒃=(b1,b2,,br)r{\boldsymbol{b}}=(b_{1},b_{2},\ldots,b_{r})\in{\mathbb{R}}^{r} satisfying

b1b2brb1mP,bi+r(r+1)/2=0.b_{1}\geq b_{2}\geq\cdots\geq b_{r}\geq b_{1}-m_{P},\quad\sum b_{i}+r(r+1)/2=0.

We set Dess:=DDmeroD_{\mathop{\rm ess}\nolimits}:=D\setminus D_{\mathop{\rm mero}\nolimits}. We assume that for any PDessP\in D_{\mathop{\rm ess}\nolimits}, fPf_{P} is a section of 𝔅X~P\mathfrak{B}_{\widetilde{X}_{P}}. At each PDessP\in D_{\mathop{\rm ess}\nolimits}, we obtain the set of the intervals 𝒮(q,P)\mathcal{S}(q,P) in ϖP1(P)\varpi_{P}^{-1}(P) which are special with respect to q|XPq_{|X_{P}^{\ast}}.

Let hHarm(q)h\in\mathop{\rm Harm}\nolimits(q). For any PDessP\in D_{\mathop{\rm ess}\nolimits} and I𝒮(q,P)I\in\mathcal{S}(q,P), we obtain 𝒂P,I(h)𝒫{\boldsymbol{a}}_{P,I}(h)\in\mathcal{P} as in Theorem 6.1. For any PD>0P\in D_{>0}, we obtain 𝒃(h)𝒫(q,P){\boldsymbol{b}}(h)\in\mathcal{P}(q,P) as in Proposition 3.21. Thus, we obtain the map

Harm(q)PDessI𝒮(q,P)𝒫×PD>0𝒫(q,P).\mathop{\rm Harm}\nolimits(q)\longrightarrow\prod_{P\in D_{\mathop{\rm ess}\nolimits}}\prod_{I\in\mathcal{S}(q,P)}\mathcal{P}\times\prod_{P\in D_{>0}}\mathcal{P}(q,P). (124)

We introduce a boundary condition at infinity of XX when XX is non-compact. Because (KXD(r+12i)/2)1KXD(r+12(i+1))/2KXD1(K_{X\setminus D}^{(r+1-2i)/2})^{-1}\otimes K_{X\setminus D}^{(r+1-2(i+1))/2}\simeq K_{X\setminus D}^{-1}, the restrictions of h|KXD(r+12i/2)h_{|K_{X\setminus D}^{(r+1-2i/2)}} and h|KXD(r+12(i+1))/2h_{|K_{X\setminus D}^{(r+1-2(i+1))/2}} induce a Kähler metric g(h)ig(h)_{i} of XDX\setminus D (i=1,,r1)(i=1,\ldots,r-1).

Definition 6.4.

We say that hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) is complete at infinity of XX if there exists a relatively compact open neighbourhood NN of DD such that g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are complete. Let Harm(q;D,c)\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) denote the set of hHarm(q)h\in\mathop{\rm Harm}\nolimits(q) which is complete at infinity of XX.

Lemma 6.5.

Let NN be any relatively compact open neighbourhood of DD in XX.

  • For any hHarm(q;D,c)h\in\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits), g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are complete. Moreover, g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded, and |q|g(h)i|XN|q|_{g(h)_{i}|X\setminus N} are bounded.

  • Any hjHarm(q;D,c)h_{j}\in\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) (j=1,2)(j=1,2) are mutually bounded on XNX\setminus N.

Proof   The first claim follows from [24, Proposition 3.27]. The second follows from [24, Proposition 3.29]. ∎

We obtain the map

Harm(q;D,c)PDessI𝒮(q,P)𝒫×PD>0𝒫(q,P).\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits)\longrightarrow\prod_{P\in D_{\mathop{\rm ess}\nolimits}}\prod_{I\in\mathcal{S}(q,P)}\mathcal{P}\times\prod_{P\in D_{>0}}\mathcal{P}(q,P). (125)

We set Harm(q;D,c):=Harm(q)Harm(q;D,c)\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q;D,\mathop{\rm c}\nolimits):=\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q)\cap\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits). Let 𝒫(q,P)\mathcal{P}^{{\mathbb{R}}}(q,P) denote the set of 𝒃𝒫(q,P){\boldsymbol{b}}\in\mathcal{P}(q,P) such that bi+br+1i=r1b_{i}+b_{r+1-i}=-r-1. Let 𝒫\mathcal{P}^{{\mathbb{R}}} denote the set of 𝒂𝒫{\boldsymbol{a}}\in\mathcal{P} such that ai+ar+1i=0a_{i}+a_{r+1-i}=0. As the restriction of (124) and (125), we obtain the following maps:

Harm(q)PDessI𝒮(q,P)𝒫×PD>0𝒫(q,P),\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q)\longrightarrow\prod_{P\in D_{\mathop{\rm ess}\nolimits}}\prod_{I\in\mathcal{S}(q,P)}\mathcal{P}^{{\mathbb{R}}}\times\prod_{P\in D_{>0}}\mathcal{P}^{{\mathbb{R}}}(q,P), (126)
Harm(q;D,c)PDessI𝒮(q,P)𝒫×PD>0𝒫(q,P).\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q;D,\mathop{\rm c}\nolimits)\longrightarrow\prod_{P\in D_{\mathop{\rm ess}\nolimits}}\prod_{I\in\mathcal{S}(q,P)}\mathcal{P}^{{\mathbb{R}}}\times\prod_{P\in D_{>0}}\mathcal{P}^{{\mathbb{R}}}(q,P). (127)
Theorem 6.6.

The maps (125) and (127) are bijective. In particular, if XX is compact, the maps (124) and (126) are bijective.

The theorem follows from Lemma 6.8, Lemma 6.10 and Lemma 6.11 below.

Remark 6.7.

If the zero set of qq is finite, we obtain a refined estimate around each PDessP\in D_{\mathop{\rm ess}\nolimits} as in Proposition 6.3.

6.2.2. Uniqueness

We set 𝒮(q):=PDess𝒮(q,P)\mathcal{S}(q):=\coprod_{P\in D_{\mathop{\rm ess}\nolimits}}\mathcal{S}(q,P).

Lemma 6.8.

Suppose that hjHarm(q;D,c)h_{j}\in\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) (j=1,2)(j=1,2) satisfy 𝐛P(h1)=𝐛P(h2){\boldsymbol{b}}_{P}(h_{1})={\boldsymbol{b}}_{P}(h_{2}) for any PD>0P\in D_{>0} and 𝐚I(h1)=𝐚I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) for I𝒮(q)I\in\mathcal{S}(q). Then, we obtain h1=h2h_{1}=h_{2}.

Proof   As in Theorem 1.4, there uniquely exists hcHarm(q)h^{\mathop{\rm c}\nolimits}\in\mathop{\rm Harm}\nolimits(q) such that g(hc)ig(h^{\mathop{\rm c}\nolimits})_{i} are complete on XDX\setminus D. Recall that g(hc)ig(h^{\mathop{\rm c}\nolimits})_{i} are mutually bounded, and that the Gaussian curvature of g(hc)ig(h^{\mathop{\rm c}\nolimits})_{i} are bounded from below [24, Lemma 3.15].

Let gg be a Kähler metric of XX such that (i) the Gaussian curvature is bounded from below, (ii) the following condition is satisfied for a relatively compact neighbourhood NN of DD.

Condition 6.9.

g|XNg_{|X\setminus N} is mutually bounded with the metrics g(hc)ig(h^{\mathop{\rm c}\nolimits})_{i} (i=1,,r1)(i=1,\ldots,r-1).

Let ss be the automorphism of 𝕂XD,r\mathbb{K}_{X\setminus D,r} determined by h2=h1sh_{2}=h_{1}\cdot s. Let Λ\Lambda denote the adjoint of the multiplication of the Kähler form associated with gg. By (20), we obtain the following on XDX\setminus D:

1Λ¯Tr(s)=|[θ(q),s]s1/2|2h1,g|¯E(s)s1/2|2h1,g|[θ(q),s]s1/2|2h1,g.\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s)=-\bigl{|}[\theta(q),s]s^{-1/2}\bigr{|}^{2}_{h_{1},g}-\bigl{|}\overline{\partial}_{E}(s)s^{-1/2}\bigr{|}^{2}_{h_{1},g}\\ \leq-\bigl{|}[\theta(q),s]s^{-1/2}\bigr{|}^{2}_{h_{1},g}. (128)

By Proposition 3.19, Proposition 3.21, Theorem 6.1, and Lemma 6.5, we obtain the boundedness of Tr(s)\mathop{\rm Tr}\nolimits(s). Hence, the inequality (128) holds across DD. (See [39, Lemma 2.2].) It particularly implies that Tr(s)\mathop{\rm Tr}\nolimits(s) is subharmonic on XX.

Let NN be a relatively compact neighbourhood of DD. If

supQXDTr(s)(Q)=supQNDTr(s)(Q),\sup_{Q\in X\setminus D}\mathop{\rm Tr}\nolimits(s)(Q)=\sup_{Q\in N\setminus D}\mathop{\rm Tr}\nolimits(s)(Q),

the maximum principle for subharmonic functions implies that Tr(s)\mathop{\rm Tr}\nolimits(s) is constant on XDX\setminus D, and hence [θ(q),s]=0[\theta(q),s]=0 on XDX\setminus D. Together with det(s)=1\det(s)=1, we obtain s=ids=\mathop{\rm id}\nolimits.

Let us study the case supQXDTr(s)(Q)>supQNDTr(s)(Q)\sup_{Q\in X\setminus D}\mathop{\rm Tr}\nolimits(s)(Q)>\sup_{Q\in N\setminus D}\mathop{\rm Tr}\nolimits(s)(Q). By Omori-Yau maximum principle [24, Lemma 3.2], there exists a sequence QXNQ_{\ell}\in X\setminus N such that

21Λ¯Tr(s)(Q)1,Tr(s)(Q)>supQXDTr(s)(Q)1.-2\sqrt{-1}\Lambda\overline{\partial}\partial\mathop{\rm Tr}\nolimits(s)(Q_{\ell})\leq\ell^{-1},\quad\mathop{\rm Tr}\nolimits(s)(Q_{\ell})>\sup_{Q\in X\setminus D}\mathop{\rm Tr}\nolimits(s)(Q_{\ell})-\ell^{-1}.

By (128), there exists C1>0C_{1}>0 such that the following holds for any \ell:

|[s,θ(q)]s1/2|2h1,g(Q)C11.\bigl{|}[s,\theta(q)]s^{-1/2}\bigr{|}^{2}_{h_{1},g}(Q_{\ell})\leq C_{1}\ell^{-1}.

Let zz_{\ell} be a holomorphic coordinate around QQ_{\ell} such that |dz|2g(Q)=2|dz_{\ell}|^{2}_{g}(Q_{\ell})=2. By Condition 6.9, there exists B>0B>0 such that |(dz)(r+12i)/2|h1|(dz)(r+12(i+1))/2|h11B\bigl{|}(dz_{\ell})^{(r+1-2i)/2}\bigr{|}_{h_{1}}\cdot\bigl{|}(dz_{\ell})^{(r+1-2(i+1))/2}\bigr{|}_{h_{1}}^{-1}\leq B for any \ell. Hence, by [24, Lemma 3.9] there exists C2>0C_{2}>0 and 2\ell_{2} such that Tr(s)(Q)r(1+C21/2)\mathop{\rm Tr}\nolimits(s)(Q_{\ell})\leq r(1+C_{2}\ell^{-1/2}) for any >2\ell>\ell_{2}. Hence, we obtain supQXDTr(s)(Q)r\sup_{Q\in X\setminus D}\mathop{\rm Tr}\nolimits(s)(Q)\leq r. Because det(s)=1\det(s)=1, we obtain Tr(s)r\mathop{\rm Tr}\nolimits(s)\geq r for any QQ. Therefore, Tr(s)\mathop{\rm Tr}\nolimits(s) is constantly rr, and we obtain s=ids=\mathop{\rm id}\nolimits. ∎

Lemma 6.10.

Suppose that hHarm(q;D,c)h\in\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) satisfies 𝐛P(h)𝒫(q,P){\boldsymbol{b}}_{P}(h)\in\mathcal{P}^{{\mathbb{R}}}(q,P) for any PD>0P\in D_{>0} and 𝐚I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P}^{{\mathbb{R}}} for any I𝒮(q)I\in\mathcal{S}(q). Then, we obtain hHarm(q;D,c)h\in\mathop{\rm Harm}\nolimits^{{\mathbb{R}}}(q;D,\mathop{\rm c}\nolimits).

Proof   We obtain hHarm(q)h^{\lor}\in\mathop{\rm Harm}\nolimits(q) as the dual of hh by using the natural identification of 𝕂XD,r\mathbb{K}_{X\setminus D,r} with its dual. Clearly, hh^{\lor} is also complete at infinity of XX. Because 𝒃P(h)𝒫(q,P){\boldsymbol{b}}_{P}(h)\in\mathcal{P}^{{\mathbb{R}}}(q,P) (PD>0)(P\in D_{>0}) and 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P}^{{\mathbb{R}}} (I𝒮(q))(I\in\mathcal{S}(q)), we obtain 𝒃P(h)=𝒃P(h){\boldsymbol{b}}_{P}(h)={\boldsymbol{b}}_{P}(h^{\lor}) and 𝒂I(h)=𝒂I(h){\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}_{I}(h^{\lor}). By Lemma 6.8, we obtain h=hh=h^{\lor}. ∎

6.2.3. Existence

Lemma 6.11.

For any 𝐛P𝒫(q,P){\boldsymbol{b}}_{P}\in\mathcal{P}(q,P) (PD>0)(P\in D_{>0}) and 𝐚I𝒫{\boldsymbol{a}}_{I}\in\mathcal{P} (I𝒮(q))(I\in\mathcal{S}(q)), there exists hHarm(q;D,c)h\in\mathop{\rm Harm}\nolimits(q;D,\mathop{\rm c}\nolimits) such that 𝐛P(h)=𝐛P{\boldsymbol{b}}_{P}(h)={\boldsymbol{b}}_{P} for any PD>0P\in D_{>0} and that 𝐚I(h)=𝐚I{\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}_{I} for any I𝒮(q)I\in\mathcal{S}(q).

Proof   Let hch^{\mathop{\rm c}\nolimits} be as in the proof of Lemma 6.8. For any (𝒂I)I𝒮(q)({\boldsymbol{a}}_{I})_{I\in\mathcal{S}(q)} and (𝒃P)PD>0({\boldsymbol{b}}_{P})_{P\in D_{>0}}, by Proposition 3.23 and Theorem 6.1, there exists a GrG_{r}-invariant Hermitian metric h0h_{0} of 𝕂XD,r\mathbb{K}_{X\setminus D,r} such that the following holds.

  • There exists a relatively compact neighbourhood N1N_{1} of DD such that h0|XN1=hc|XN1h_{0|X\setminus N_{1}}=h^{\mathop{\rm c}\nolimits}_{|X\setminus N_{1}}.

  • There exists a relatively compact neighbourhood N2N1N_{2}\subset N_{1} of DD such that h0|N2DHarm(q|N2D)h_{0|N_{2}\setminus D}\in\mathop{\rm Harm}\nolimits(q_{|N_{2}\setminus D}) with 𝒃P(h0)=𝒃P{\boldsymbol{b}}_{P}(h_{0})={\boldsymbol{b}}_{P} for PD>0P\in D_{>0}, and 𝒂I(h0)=𝒂I{\boldsymbol{a}}_{I}(h_{0})={\boldsymbol{a}}_{I} for I𝒮(q)I\in\mathcal{S}(q).

  • det(h0)=1\det(h_{0})=1.

Then, according to Proposition 3.18, there exists hHarm(q;D,hc)h\in\mathop{\rm Harm}\nolimits(q;D,h^{\mathop{\rm c}\nolimits}) which is mutually bounded with h0h_{0}. Thus, we obtain Lemma 6.11, and the proof of Theorem 6.6 is completed. ∎

6.3. Examples on {\mathbb{C}}

Let ϖ:~11\varpi_{\infty}:\widetilde{\mathbb{P}}^{1}_{\infty}\longrightarrow\mathbb{P}^{1} be the oriented real blow up at \infty. We identify ϖ1()\varpi_{\infty}^{-1}(\infty) with S1S^{1} by the polar decomposition z=|z|e1θz=|z|e^{\sqrt{-1}\theta}.

6.3.1. Classification in the case of γ(z)e𝔞(z)(dz)r\gamma(z)e^{\mathfrak{a}(z)}(dz)^{r}

Let ρ\rho be a positive integer. Let α\alpha be a non-zero complex number. Let Λ(α,ρ)\Lambda(\alpha,\rho) denote the set of the connected components of

{e1θS1|Re(αe1ρθ)<0}.\bigl{\{}e^{\sqrt{-1}\theta}\in S^{1}\,\big{|}\,\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\rho\theta})<0\bigr{\}}.

Note that |Λ(α,ρ)|=ρ|\Lambda(\alpha,\rho)|=\rho. Let 𝔞\mathfrak{a} be a polynomial of the form

𝔞=αzρ+j=1ρ1𝔞jzj.\mathfrak{a}=\alpha z^{\rho}+\sum_{j=1}^{\rho-1}\mathfrak{a}_{j}z^{j}.

Let γ(z)\gamma(z) be a non-zero polynomial. We set q=γ(z)e𝔞(dz)rq=\gamma(z)e^{\mathfrak{a}}(dz)^{r}. Clearly, Λ(α,ρ)\Lambda(\alpha,\rho) is the set of the intervals special with respect to qq. For any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), we obtain 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} (IΛ(α,ρ))(I\in\Lambda(\alpha,\rho)). The following proposition is a special case of Theorem 6.6.

Proposition 6.12.

The map Harm(q)𝒫Λ(α,ρ)\mathop{\rm Harm}\nolimits(q)\longrightarrow\mathcal{P}^{\Lambda(\alpha,\rho)} is bijective. ∎

Remark 6.13.

We can obtain a refined estimate around \infty as in Proposition 6.3.

6.3.2. The case of i=1nγi(z)e𝔞i(z)(dz)r\sum_{i=1}^{n}\gamma_{i}(z)e^{\mathfrak{a}_{i}(z)}\,(dz)^{r}

Let n2n\geq 2. Let 𝔞i(z)\mathfrak{a}_{i}(z) (i=1,2,,n)(i=1,2,\ldots,n) be mutually distinct polynomials such that 𝔞i(0)=0\mathfrak{a}_{i}(0)=0. If 𝔞i0\mathfrak{a}_{i}\neq 0, we set ρi:=deg(𝔞i)\rho_{i}:=\deg(\mathfrak{a}_{i}), and let αi0\alpha_{i}\neq 0 denote the coefficients of the top term of 𝔞i\mathfrak{a}_{i}, i.e.,

𝔞i(z)=αizρi+j=1ρi1𝔞i,jzj.\mathfrak{a}_{i}(z)=\alpha_{i}z^{\rho_{i}}+\sum_{j=1}^{\rho_{i}-1}\mathfrak{a}_{i,j}z^{j}.

If 𝔞i=0\mathfrak{a}_{i}=0, we set ρi=0\rho_{i}=0 and αi=0\alpha_{i}=0. Let γi(z)\gamma_{i}(z) (i=1,,n)(i=1,\ldots,n) be non-zero polynomials. We set q=i=1nγi(z)e𝔞i(z)(dz)rq=\sum_{i=1}^{n}\gamma_{i}(z)e^{\mathfrak{a}_{i}(z)}(dz)^{r}. The following proposition is a special case of Theorem 6.6.

Proposition 6.14.

  • Suppose that there exist ρ>0\rho>0 and α\alpha\in{\mathbb{C}}^{\ast} such that ρi=ρ\rho_{i}=\rho and αi/α>0\alpha_{i}/\alpha\in{\mathbb{R}}_{>0} for any ii. Then, Λ(α,ρ)\Lambda(\alpha,\rho) is the set of the intervals which are special with respect to qq. Hence, for any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), we obtain 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} (IΛ(α,ρ))(I\in\Lambda(\alpha,\rho)), which induces a bijection Harm(q)𝒫Λ(α,ρ)\mathop{\rm Harm}\nolimits(q)\longrightarrow\mathcal{P}^{\Lambda(\alpha,\rho)}.

  • Otherwise, we obtain Harm(q)={hc(q)}\mathop{\rm Harm}\nolimits(q)=\{h^{\mathop{\rm c}\nolimits}(q)\}.

Proof   Let us prove the first claim. We set κ0:=max{αi/α}\kappa_{0}:=\max\{\alpha_{i}/\alpha\} and κ1:=min{αi/α}\kappa_{1}:=\min\{\alpha_{i}/\alpha\}. Let ff be the entire function determined by q=f(dz)rq=f(dz)^{r}. We set 𝒵(f)={θS1=/2π|Re(αe1θ)=0}\mathcal{Z}(f)=\bigl{\{}\theta\in S^{1}={\mathbb{R}}/2\pi{\mathbb{Z}}\,\big{|}\,\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\theta})=0\bigr{\}}. Let θS1𝒵(q)\theta\in S^{1}\setminus\mathcal{Z}(q). There exists 𝔞i\mathfrak{a}_{i} such that 𝔞i=𝔞(f,θ)\mathfrak{a}_{i}=\mathfrak{a}(f,\theta) (See Definition 4.17 for 𝔞(f,θ)\mathfrak{a}(f,\theta).) Note that Re(αe1ρθ)0\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\rho\theta})\neq 0. If Re(αe1ρθ)>0\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\rho\theta})>0, we obtain αi/α=κ0\alpha_{i}/\alpha=\kappa_{0}. If Re(αe1ρθ)<0\mathop{\rm Re}\nolimits(\alpha e^{\sqrt{-1}\rho\theta})<0, we obtain αi/α=κ1\alpha_{i}/\alpha=\kappa_{1}. Then, the first claim is clear.

Let us prove the second claim. Note that qq is not meromorphic at \infty under the assumption that n2n\geq 2. Suppose that there exists a special interval IS1I\subset S^{1} with respect to ff. There exists ρ>0\rho>0 such that the length of II is π/ρ\pi/\rho. Because there exists 𝔞i\mathfrak{a}_{i} such that ρ=ρi\rho=\rho_{i}, ρ\rho is a positive integer. Suppose that there exists 𝔞j\mathfrak{a}_{j} such that ρj>ρ\rho_{j}>\rho. Then, there exists θI𝒵(f)\theta\in I\setminus\mathcal{Z}(f) such that Re(αje1ρjθ)>0\mathop{\rm Re}\nolimits(\alpha_{j}e^{\sqrt{-1}\rho_{j}\theta})>0. It contradicts that II is negative. Suppose that there exists ρj<ρ\rho_{j}<\rho. For any θI𝒵(f)\theta\in I\setminus\mathcal{Z}(f), there exists 𝔞i\mathfrak{a}_{i} such that ρi=ρ\rho_{i}=\rho and 𝔞i=𝔞(f,θ)\mathfrak{a}_{i}=\mathfrak{a}(f,\theta). But, we obtain Re(αjsρje1ρjθ)>Re(αisρie1ρiθ)\mathop{\rm Re}\nolimits(\alpha_{j}s^{\rho_{j}}e^{\sqrt{-1}\rho_{j}\theta})>\mathop{\rm Re}\nolimits(\alpha_{i}s^{\rho_{i}}e^{\sqrt{-1}\rho_{i}\theta}) for any sufficiently large ss, which contradicts 𝔞i=𝔞(f,θ)\mathfrak{a}_{i}=\mathfrak{a}(f,\theta). Therefore, we obtain ρi=ρ\rho_{i}=\rho for any ii. There exists α\alpha\in{\mathbb{C}}^{\ast} such that for any θI𝒵(f)\theta\in I\setminus\mathcal{Z}(f) we obtain deg(𝔞(f,θ)αzρ)<ρ\deg(\mathfrak{a}(f,\theta)-\alpha z^{\rho})<\rho. Suppose that there exists 𝔞j\mathfrak{a}_{j} such that αj/α>0\alpha_{j}/\alpha\not\in{\mathbb{R}}_{>0}. Then, there exists θI\theta\in I such that Re(αje1ρθ)>0\mathop{\rm Re}\nolimits(\alpha_{j}e^{\sqrt{-1}\rho\theta})>0. It contradicts that II is negative. Hence, we obtain αj/α>0\alpha_{j}/\alpha\in{\mathbb{R}}_{>0} for any jj. ∎

Corollary 6.15.

Let PP be any non-zero polynomial. and let QQ be any non-constant polynomial. We set qcos=Pcos(Q)(dz)rq_{\cos}=P\cos(Q)(dz)^{r}, qsin=Psin(Q)(dz)rq_{\sin}=P\sin(Q)(dz)^{r} , qcosh=Pcosh(Q)(dz)rq_{\cosh}=P\cosh(Q)(dz)^{r} and qsinh=Psinh(Q)(dz)rq_{\sinh}=P\sinh(Q)(dz)^{r}. Because Pcos(Q)=(Pe1Q+Pe1Q)/2P\cos(Q)=(Pe^{\sqrt{-1}Q}+Pe^{-\sqrt{-1}Q})/2, we obtain Harm(qcos)={hc(qcos)}\mathop{\rm Harm}\nolimits(q_{\cos})=\{h^{\mathop{\rm c}\nolimits}(q_{\cos})\}. Similarly, we obtain Harm(qsin)={hc(qsin)}\mathop{\rm Harm}\nolimits(q_{\sin})=\{h^{\mathop{\rm c}\nolimits}(q_{\sin})\}, Harm(qcosh)={hc(qcosh)}\mathop{\rm Harm}\nolimits(q_{\cosh})=\{h^{\mathop{\rm c}\nolimits}(q_{\cosh})\} and Harm(qsinh)={hc(qsinh)}\mathop{\rm Harm}\nolimits(q_{\sinh})=\{h^{\mathop{\rm c}\nolimits}(q_{\sinh})\}. ∎

6.3.3. Airy function

Let Γ\Gamma be a continuous map {\mathbb{R}}\longrightarrow{\mathbb{C}} such that Γ(t)=t(1+3)/2\Gamma(t)=t(1+\sqrt{-3})/2 for t<1t<-1 and Γ(t)=t(1+3)/2\Gamma(t)=t(-1+\sqrt{-3})/2 for t>1t>1. Recall that an Airy function is defined as

Ai(z):=Γeztt3/3dt.\mathop{\rm Ai}\nolimits(z):=\int_{\Gamma}e^{zt-t^{3}/3}dt. (129)

(For example, see [47, §22].) It is an entire function. Recall that it is a solution of the linear differential equation z2uzu=0\partial_{z}^{2}u-zu=0, and hence it induces a section of 𝔅~1\mathfrak{B}_{\widetilde{\mathbb{P}}^{1}}. On |arg(z)|<π|\arg(z)|<\pi, it satisfies the following estimate (for example, see [47, §23]):

e23z3/2z1/4(Ai(z)z1/42πe23z3/2)=O(|z|3/2).e^{\frac{2}{3}z^{3/2}}z^{1/4}\Bigl{(}\mathop{\rm Ai}\nolimits(z)-\frac{z^{-1/4}}{2\sqrt{\pi}}e^{-\frac{2}{3}z^{3/2}}\Bigr{)}=O(|z|^{-3/2}). (130)

Let γ(z)\gamma(z) be any non-zero polynomial. We set q=γ(z)Ai(z)(dz)rq=\gamma(z)\mathop{\rm Ai}\nolimits(z)\,(dz)^{r}. Because of (130), the special interval with respect to qq is I={π/3<θ<π/3}I=\{-\pi/3<\theta<\pi/3\}. We obtain the following proposition from Theorem 6.6.

Proposition 6.16.

For any hHarm(q)h\in\mathop{\rm Harm}\nolimits(q), we obtain 𝐚I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P}, which induces a bijection Harm(q)𝒫\mathop{\rm Harm}\nolimits(q)\longrightarrow\mathcal{P}. ∎

Remark 6.17.

More generally, let ff be a solution of a differential equation (zn+j=0n1aj(z)zj)f=0(\partial_{z}^{n}+\sum_{j=0}^{n-1}a_{j}(z)\partial_{z}^{j})f=0, where aj(z)a_{j}(z) are polynomials. Then, ff induces a section of 𝔅~1\mathfrak{B}_{\widetilde{\mathbb{P}}^{1}_{\infty}} according to the classical asymptotic analysis. (For example, see [28, §II.1].) Hence, Theorem 6.6 allows us to classify Harm(f(dz)r)\mathop{\rm Harm}\nolimits(f\,(dz)^{r}) in terms of the asymptotic expansion of ff.

7. A generalization

7.1. Preliminary

7.1.1. Meromorphic extensions

Let XX be a Riemann surface with a discrete subset SS. Let 𝒪X(S)\mathcal{O}_{X}(\ast S) denote the sheaf of meromorphic functions on XX which may have poles along SS. Let ι:XSX\iota:X\setminus S\longrightarrow X denote the inclusion. For any locally free 𝒪XS\mathcal{O}_{X\setminus S}-module \mathcal{M}, we obtain the ι𝒪XS\iota_{\ast}\mathcal{O}_{X\setminus S}-module ι\iota_{\ast}\mathcal{M}.

Definition 7.1.

A meromorphic extension of \mathcal{M} on (X,S)(X,S) is a locally free 𝒪X(S)\mathcal{O}_{X}(\ast S)-submodule ~ι\widetilde{\mathcal{M}}\subset\iota_{\ast}\mathcal{M} such that ι~=\iota^{\ast}\widetilde{\mathcal{M}}=\mathcal{M}.

7.1.2. Meromorphic extensions of cyclic Higgs bundles on a punctured disc

Let UU be a neighbourhood of 0 in {\mathbb{C}}. We set U:=U{0}U^{\circ}:=U\setminus\{0\}. Let 𝔏i\mathfrak{L}_{i} (i=1,r)(i=1,\ldots r) be locally free 𝒪U\mathcal{O}_{U^{\circ}}-modules of rank one. Let ψi:𝔏i𝔏i+1KU\psi_{i}:\mathfrak{L}_{i}\longrightarrow\mathfrak{L}_{i+1}\otimes K_{U^{\circ}} (i=1,,r1)(i=1,\ldots,r-1) and ψr:𝔏r𝔏1KU\psi_{r}:\mathfrak{L}_{r}\longrightarrow\mathfrak{L}_{1}\otimes K_{U^{\circ}} be 𝒪U\mathcal{O}_{U^{\circ}}-morphisms. We obtain a holomorphic section qr1:=ψr1ψ1q_{\leq r-1}:=\psi_{r-1}\circ\cdots\circ\psi_{1} of Hom(𝔏1,𝔏r)KU(r1)\mathop{\rm Hom}\nolimits(\mathfrak{L}_{1},\mathfrak{L}_{r})\otimes K_{U^{\circ}}^{\otimes(r-1)}. We assume the following.

  • The zero set of qr1q_{\leq r-1} is finite.

We set 𝔙:=i=1r𝔏i\mathfrak{V}:=\bigoplus_{i=1}^{r}\mathfrak{L}_{i}. Let θ\theta be the Higgs field of 𝔙\mathfrak{V} induced by ψi\psi_{i} (i=1,r)(i=1\ldots,r).

Proposition 7.2.

For any meromorphic extension det(𝔙)~\widetilde{\det(\mathfrak{V})} of det(𝔙)\det(\mathfrak{V}) on (U,0)(U,0), there uniquely exist meromorphic extensions 𝔏~i\widetilde{\mathfrak{L}}_{i} of 𝔏i\mathfrak{L}_{i} (i=1,,r)(i=1,\ldots,r) on (U,0)(U,0) such that (i) θ(𝔏~i)𝔏~i+1KU\theta(\widetilde{\mathfrak{L}}_{i})\subset\widetilde{\mathfrak{L}}_{i+1}\otimes K_{U} (i=1,,r1)(i=1,\ldots,r-1), (ii) det(𝔙~)=det(𝔙)~\det(\widetilde{\mathfrak{V}})=\widetilde{\det(\mathfrak{V})}, where 𝔙~=𝔏~i\widetilde{\mathfrak{V}}=\bigoplus\widetilde{\mathfrak{L}}_{i}.

Proof   By shrinking UU, we may assume that qr1q_{\leq r-1} is nowhere vanishing on UU^{\circ}. We obtain the morphisms βi:𝔏i𝔏i+1\beta_{i}:\mathfrak{L}_{i}\longrightarrow\mathfrak{L}_{i+1} (i=1,,r1)(i=1,\ldots,r-1) by ψi=βi(dz/z)\psi_{i}=\beta_{i}\,(dz/z).

Because UU^{\circ} is Stein, there exists a global frame u1u^{\star}_{1} of 𝔏1\mathfrak{L}_{1}. We set ui:=βi1β1(u1)u^{\star}_{i}:=\beta_{i-1}\circ\cdots\beta_{1}(u^{\star}_{1}) on UU^{\circ} for i=2,,ri=2,\ldots,r. We obtain the meromorphic extensions 𝔏~i=𝒪Uui\widetilde{\mathfrak{L}}^{\star}_{i}=\mathcal{O}_{U}\cdot u^{\star}_{i} of 𝔏i\mathfrak{L}_{i}.

Let v0v_{0} be a frame of det(𝔙)~\widetilde{\det(\mathfrak{V})}. Let ff be the holomorphic function on UU^{\circ} determined by u1ur=fv0u^{\star}_{1}\wedge\cdots\wedge u^{\star}_{r}=f\cdot v_{0}. There exist an integer \ell and a holomorphic function gg on UU^{\circ} such that f=zegf=z^{\ell}e^{g}. We set ui:=eg/ruiu_{i}:=e^{-g/r}u_{i}^{\star} and we obtain meromorphic extensions 𝔏~i:=𝒪U(0)ui\widetilde{\mathfrak{L}}_{i}:=\mathcal{O}_{U}(\ast 0)u_{i}. We set 𝔙~=𝔏~i\widetilde{\mathfrak{V}}=\bigoplus\widetilde{\mathfrak{L}}_{i}. Then, by the construction, we obtain θ(𝔏~i)𝔏~i+1KU\theta(\widetilde{\mathfrak{L}}_{i})\subset\widetilde{\mathfrak{L}}_{i+1}\otimes K_{U} for i=1,,r1i=1,\ldots,r-1, and det(𝔙~)=det(𝔙)~\det(\widetilde{\mathfrak{V}})=\widetilde{\det(\mathfrak{V})}.

Let 𝔏~i\widetilde{\mathfrak{L}}^{\sharp}_{i} be meromorphic extensions of 𝔏i\mathfrak{L}_{i} such that (i) θ𝔏~i𝔏~i+1KU\theta\widetilde{\mathfrak{L}}^{\sharp}_{i}\subset\widetilde{\mathfrak{L}}^{\sharp}_{i+1}\otimes K_{U} (i=1,,r1)(i=1,\ldots,r-1), (ii) det(𝔙~)=det(𝔙)~=det(𝔙~)\det(\widetilde{\mathfrak{V}}^{\sharp})=\widetilde{\det(\mathfrak{V})}=\det(\widetilde{\mathfrak{V}}), where 𝔙~=𝔏~i\widetilde{\mathfrak{V}}^{\sharp}=\bigoplus\widetilde{\mathfrak{L}}^{\sharp}_{i}. There exist frames uiu^{\sharp}_{i} of 𝔏~i\widetilde{\mathfrak{L}}^{\sharp}_{i} such that θ(ui)=ui+1dz/z\theta(u^{\sharp}_{i})=u^{\sharp}_{i+1}dz/z for i=1,,r1i=1,\ldots,r-1.

Let γ\gamma be the holomorphic function on UU^{\circ} determined by u1=γu1u^{\sharp}_{1}=\gamma u_{1}. Then, we obtain ui=γuiu^{\sharp}_{i}=\gamma u_{i} for i=1,,ri=1,\ldots,r. Because both u1uru^{\sharp}_{1}\wedge\cdots\wedge u^{\sharp}_{r} and u1uru_{1}\wedge\cdots\wedge u_{r} are sections of det(𝔙)~\widetilde{\det(\mathfrak{V})}, γr\gamma^{r} is meromorphic at 0. Hence, we obtain that γ\gamma is meromorphic, i.e., 𝔏~i=𝔏~i\widetilde{\mathfrak{L}}_{i}=\widetilde{\mathfrak{L}}_{i}^{\sharp}. ∎

7.2. Cyclic Higgs bundles

Let XX be a Riemann surface with a finite subset DXD\subset X. Assume that XDX\setminus D is an open Riemann surface, i.e., XX is open, or XX is compact and DD\neq\emptyset. (See Remark 7.12.) For each PDP\in D, let (XP,zP)(X_{P},z_{P}) be a holomorphic coordinate neighbourhood such that zP(P)=0z_{P}(P)=0. Set XP:=XP{P}X_{P}^{\ast}:=X_{P}\setminus\{P\}.

Let r2r\geq 2. Let LiL_{i} (i=1,,r)(i=1,\ldots,r) be holomorphic line bundles on XDX\setminus D. Let ψi:LiLi+1KXD\psi_{i}:L_{i}\longrightarrow L_{i+1}\otimes K_{X\setminus D} and ψr:LrL1KXD\psi_{r}:L_{r}\longrightarrow L_{1}\otimes K_{X\setminus D} be non-zero morphisms. We set E:=LiE:=\bigoplus L_{i}. Let θ\theta be the cyclic Higgs field of EE induced by ψi\psi_{i} (i=1,,r)(i=1,\ldots,r).

We obtain the holomorphic section q:=ψrψ1=(1)r1det(θ)q:=\psi_{r}\circ\cdots\circ\psi_{1}=(-1)^{r-1}\det(\theta) of KXDrK_{X\setminus D}^{\otimes r}, and the holomorphic section qr1:=ψr1ψ1q_{\leq r-1}:=\psi_{r-1}\circ\cdots\circ\psi_{1} of Hom(L1,Lr)KXD(r1)\mathop{\rm Hom}\nolimits(L_{1},L_{r})\otimes K_{X\setminus D}^{\otimes(r-1)}. We assume the following.

  • The zero set of qr1q_{\leq r-1} is finite.

  • qq is not constantly 0.

  • Let fPf_{P} (PD)(P\in D) be holomorphic functions on XPX_{P}^{\ast} obtained as q|XP=fP(dzP/zP)rq_{|X_{P}^{\ast}}=f_{P}\,(dz_{P}/z_{P})^{r}. Then, fPf_{P} have multiple growth orders at PP.

Let DmeroD_{\mathop{\rm mero}\nolimits} denote the set of PDP\in D such that fPf_{P} is meromorphic at PP. We put Dess:=DDmeroD_{\mathop{\rm ess}\nolimits}:=D\setminus D_{\mathop{\rm mero}\nolimits}. For PDmeroP\in D_{\mathop{\rm mero}\nolimits}, we describe fP=zPmPβPf_{P}=z_{P}^{m_{P}}\beta_{P}, where βP\beta_{P} is a nowhere vanishing holomorphic function on XPX_{P}. Let D>0D_{>0} denote the set of PDmeroP\in D_{\mathop{\rm mero}\nolimits} such that mP>0m_{P}>0. We set D0:=DmeroD>0D_{\leq 0}:=D_{\mathop{\rm mero}\nolimits}\setminus D_{>0}.

7.2.1. Flat metrics on the determinant and local frames

Let hdet(E)h_{\det(E)} be a Hermitian metric of det(E)\det(E) such that the Chern connection of (det(E),hdet(E))\bigl{(}\det(E),h_{\det(E)}\bigr{)} is flat. Note that such a metric exists because det(E)𝒪XD\det(E)\simeq\mathcal{O}_{X\setminus D} under the assumption that XDX\setminus D is non-compact. (For example, see [12, Theorem 30.3].)

Lemma 7.3.

For each PDP\in D, there exist frames 𝐯P=(vP,i|i=1,,r){\boldsymbol{v}}_{P}=(v_{P,i}\,|\,i=1,\ldots,r) of Li|XPL_{i|X_{P}^{\ast}} and a real number c(𝐯P)c({\boldsymbol{v}}_{P}) such that the following condition is satisfied.

  • θ(vP,i)=vP,i+1(dzP/zP)\theta(v_{P,i})=v_{P,i+1}(dz_{P}/z_{P}) for i=1,,r1i=1,\ldots,r-1.

  • We set ωP=vP,1vP,r\omega_{P}=v_{P,1}\wedge\cdots\wedge v_{P,r}. Then, we obtain |zP|c(𝒗P)|ωP|hdet(E)=1|z_{P}|^{c({\boldsymbol{v}}_{P})}\cdot|\omega_{P}|_{h_{\det(E)}}=1.

Proof   There exists a frame wPw_{P} of det(E)|XP\det(E)_{|X_{P}^{\ast}} such that |wP|hdet(E)=|zP|dP|w_{P}|_{h_{\det(E)}}=|z_{P}|^{d_{P}} for a real number dPd_{P}. By Proposition 7.2, there exists a frame vP,iv^{\prime}_{P,i} of Li|XPL_{i|X_{P}^{\ast}} (i=1,,r)(i=1,\ldots,r) such that θ(vP,i)=vP,i+1(dzP/zP)\theta(v^{\prime}_{P,i})=v^{\prime}_{P,i+1}(dz_{P}/z_{P}) for i=1,,r1i=1,\ldots,r-1, and vP,1vP,r=zPnPβPwPv^{\prime}_{P,1}\wedge\cdots\wedge v^{\prime}_{P,r}=z_{P}^{n_{P}}\beta_{P}\cdot w_{P} for an integer nPn_{P} and a nowhere vanishing holomorphic function βP\beta_{P} on XPX_{P}. By fixing an rr-th root βP1/r\beta_{P}^{1/r} of βP\beta_{P}, and by setting vP,i:=βP1/rvP,iv_{P,i}:=\beta_{P}^{-1/r}v_{P,i}^{\prime}, we obtain the claim of the lemma. ∎

For PD>0P\in D_{>0}, let 𝒫(q,P,𝒗P)\mathcal{P}(q,P,{\boldsymbol{v}}_{P}) denote the set of 𝒃=(bi)r{\boldsymbol{b}}=(b_{i})\in{\mathbb{R}}^{r} satisfying the following conditions:

i=1rbi=c(𝒗P),bibi+1(i=1,,r1),brb1mP.\sum_{i=1}^{r}b_{i}=c({\boldsymbol{v}}_{P}),\quad b_{i}\geq b_{i+1}\,\,(i=1,\ldots,r-1),\quad b_{r}\geq b_{1}-m_{P}.
Remark 7.4.

If (E,θ)=(𝕂XD,r,θ(q))(E,\theta)=(\mathbb{K}_{X\setminus D,r},\theta(q)) for an rr-differential qq, we usually choose hdet(E)=1h_{\det(E)}=1 and

vP,i=zPi(dzP)(r+12i)/2(i=1,,r),v_{P,i}=z_{P}^{i}(dz_{P})^{(r+1-2i)/2}\quad(i=1,\ldots,r),

for which we obtain c(𝐯P)=r(r+1)/2c({\boldsymbol{v}}_{P})=-r(r+1)/2.

7.2.2. Harmonic metrics and the associated parabolic weights

We consider the GrG_{r}-action on LiL_{i} by aui=aiuia\bullet u_{i}=a^{i}u_{i}, which induces a GrG_{r}-action on EE. For any open subset YXDY\subset X\setminus D, let HarmGr((E,θ)|Y,hdet(E))\mathop{\rm Harm}\nolimits^{G_{r}}((E,\theta)_{|Y},h_{\det(E)}) denote the set of GrG_{r}-invariant harmonic metrics hh of (E,θ)|Y(E,\theta)_{|Y} such that det(h)=hdet(E)|Y\det(h)=h_{\det(E)|Y}.

Proposition 7.5.

Let hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}).

  • For any PD>0P\in D_{>0}, there exists 𝒃P(h)𝒫(q,P,𝒗P){\boldsymbol{b}}_{P}(h)\in\mathcal{P}(q,P,{\boldsymbol{v}}_{P}) determined by the following condition.

    bP,i(h)=inf{b||zP|b|vP,i|h is bounded}.b_{P,i}(h)=\inf\Bigl{\{}b\in{\mathbb{R}}\,\Big{|}\,\mbox{\rm$|z_{P}|^{b}|v_{P,i}|_{h}$ is bounded}\Bigr{\}}.
  • For any PDessP\in D_{\mathop{\rm ess}\nolimits} and any I𝒮(q,P)I\in\mathcal{S}(q,P), there exist 𝒂I(h)𝒫{\boldsymbol{a}}_{I}(h)\in\mathcal{P} and ϵ>0\epsilon>0 such that the following estimates hold as |zP|0|z_{P}|\to 0 on {|arg(αIzPρ(I))π|<(1δ)π/2}\bigl{\{}|\arg(\alpha_{I}z_{P}^{-\rho(I)})-\pi|<(1-\delta)\pi/2\bigr{\}} for any δ>0\delta>0:

    log|vi|h+aI,i(h)Re(αIzPρ(I))=O(|zP|ρ(I)+ϵ)\log|v_{i}|_{h}+a_{I,i}(h)\mathop{\rm Re}\nolimits(\alpha_{I}z_{P}^{-\rho(I)})=O\bigl{(}|z_{P}|^{-\rho(I)+\epsilon}\bigr{)}

    Here, αI\alpha_{I} and ρ(I)\rho(I) are determined for II and q=(1)r1det(θ)q=(-1)^{r-1}\det(\theta) as in §6.1.1.

Proof   It is enough to study the case D={P}D=\{P\} and X=XPX=X_{P}. There exists a holomorphic line bundle det(E)1/r\det(E)^{1/r} on XPX_{P}^{\ast} with an isomorphism

(det(E)1/r)rdet(E).(\det(E)^{1/r})^{\otimes r}\simeq\det(E). (131)

There exists a flat metric hdet(E)1/rh_{\det(E)^{1/r}} of det(E)1/r\det(E)^{1/r} such that hdet(E)1/rr=hdet(E)h_{\det(E)^{1/r}}^{\otimes\,r}=h_{\det(E)} under the isomorphism (131). By using the frames 𝒗P{\boldsymbol{v}}_{P} and zPi(dzP)(r+12i)/2z_{P}^{i}(dz_{P})^{(r+1-2i)/2}, we obtain an isomorphism (E,θ)det(E)1/r(𝕂XP,r,θ(q))(E,\theta)\otimes\det(E)^{-1/r}\simeq(\mathbb{K}_{X_{P}^{\ast},r},\theta(q)). Then, the claims of the proposition are reduced to Proposition 3.21 and Theorem 6.1. ∎

Proposition 7.6.

Suppose that h1,h2HarmGr(E,θ,hdet(E))h_{1},h_{2}\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) satisfy 𝐛P(h1)=𝐛P(h2){\boldsymbol{b}}_{P}(h_{1})={\boldsymbol{b}}_{P}(h_{2}) for any PD>0P\in D_{>0} and 𝐚I(h1)=𝐚I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) for any I𝒮(q)I\in\mathcal{S}(q). Then, for any relatively compact neighbourhood NN of DD in XX, h1h_{1} and h2h_{2} are mutually bounded on NDN\setminus D.

Proof   By using the local isomorphisms in the proof of Proposition 7.5, the claim is reduced to Proposition 3.19, Proposition 3.21 and Theorem 6.1. ∎

7.2.3. Completeness at infinity

Let Z(qr1)Z(q_{\leq r-1}) denote the zero set of qr1q_{\leq r-1}, which is assumed to be finite. We set D~=DZ(qr1)\widetilde{D}=D\cup Z(q_{\leq r-1}). Note that ψi\psi_{i} (i=1,,r1)(i=1,\ldots,r-1) induce isomorphisms (Li+1/Li)|XD~KXD~1(L_{i+1}/L_{i})_{|X\setminus\widetilde{D}}\simeq K_{X\setminus\widetilde{D}}^{-1}. Hence, for any hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}), the Hermitian metric h|Li+1h|Li1h_{|L_{i+1}}\otimes h_{|L_{i}}^{-1} (i=1,,r1)(i=1,\ldots,r-1) on Li+1/LiL_{i+1}/L_{i} induce Kähler metrics g(h)ig(h)_{i} on XD~X\setminus\widetilde{D}.

Definition 7.7.

hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) is called complete at infinity of XX if there exists a relatively compact open subset NN of D~\widetilde{D} such that g(h)i|XNg(h)_{i|X\setminus N} are complete. Let HarmGr(E,θ,hdet(E);D,c)\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits) denote the set of hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) which are complete at infinity of XX.

Because XD~X\setminus\widetilde{D} is an open Riemann surface, there exists an isomorphism det(E)|XD~𝒪XD~\det(E)_{|X\setminus\widetilde{D}}\simeq\mathcal{O}_{X\setminus\widetilde{D}}. Hence, there exists a holomorphic line bundle det(E)1/r\det(E)^{1/r} on XD~X\setminus\widetilde{D} with an isomorphism ρ:(det(E)1/r)rdet(E)|XD~\rho:(\det(E)^{1/r})^{\otimes r}\simeq\det(E)_{|X\setminus\widetilde{D}}. We say that such (det(E)1/r1,ρ1)(\det(E)^{1/r}_{1},\rho_{1}) and (det(E)1/r2,ρ2)(\det(E)^{1/r}_{2},\rho_{2}) are isomorphic if there exists an isomorphism κ:det(E)1/r1det(E)1/r2\kappa:\det(E)^{1/r}_{1}\simeq\det(E)^{1/r}_{2} such that ρ2κr=ρ1\rho_{2}\circ\kappa^{\otimes\,r}=\rho_{1}.

Lemma 7.8.

If we choose (det(E)1/r,ρ)(\det(E)^{1/r},\rho) appropriately, there exist isomorphisms φi:Li|XD~KXD~(r+12i)/2det(E)1/r\varphi_{i}:L_{i|X\setminus\widetilde{D}}\simeq K_{X\setminus\widetilde{D}}^{(r+1-2i)/2}\otimes\det(E)^{1/r} (i=1,,r)(i=1,\ldots,r) such that the following conditions are satisfied:

  • The following diagram is commutative.

    E|XD~φi𝕂XD~,rdet(E)1/rθθ(q)E|XD~KXD~φi𝕂XD~,rdet(E)1/rKXD~.\begin{CD}E_{|X\setminus\widetilde{D}}@>{\bigoplus\varphi_{i}}>{}>\mathbb{K}_{X\setminus\widetilde{D},r}\otimes\det(E)^{1/r}\\ @V{\theta}V{}V@V{\theta(q)}V{}V\\ E_{|X\setminus\widetilde{D}}\otimes K_{X\setminus\widetilde{D}}@>{\bigoplus\varphi_{i}}>{}>\mathbb{K}_{X\setminus\widetilde{D},r}\otimes\det(E)^{1/r}\otimes K_{X\setminus\widetilde{D}}.\end{CD}
  • The induced morphism

    det(φi):det(E|XD~)det(𝕂XD~,rdet(E)1/r)=(det(E)1/r)r\det\Bigl{(}\bigoplus\varphi_{i}\Bigr{)}:\det(E_{|X\setminus\widetilde{D}})\simeq\det\bigl{(}\mathbb{K}_{X\setminus\widetilde{D},r}\otimes\det(E)^{1/r}\bigr{)}=(\det(E)^{1/r})^{\otimes\,r}

    is equal to ρ1\rho^{-1}.

Such (det(E)1/r,ρ)(\det(E)^{1/r},\rho) is unique up to isomorphisms. Such an isomorphism φi\bigoplus\varphi_{i} is unique up to the multiplication of an rr-th root of 11.

Proof   We take a holomorphic line bundle det(E)01/r\det(E)_{0}^{1/r} with an isomorphism ρ0:(det(E)01/r)rdet(E)\rho_{0}:(\det(E)_{0}^{1/r})^{\otimes r}\simeq\det(E). There exists an isomorphism of holomorphic line bundles ν1:L1|XD~KXD~(r1)/2det(E)1/r0|XD~\nu_{1}:L_{1|X\setminus\widetilde{D}}\simeq K_{X\setminus\widetilde{D}}^{\otimes(r-1)/2}\otimes\det(E)^{1/r}_{0|X\setminus\widetilde{D}}. We obtain the holomorphic isomorphisms νi:Li|XD~KXD~(r+12i)/2det(E)1/r0|XD~\nu_{i}:L_{i|X\setminus\widetilde{D}}\simeq K_{X\setminus\widetilde{D}}^{\otimes(r+1-2i)/2}\otimes\det(E)^{1/r}_{0|X\setminus\widetilde{D}} (i=2,,r)(i=2,\ldots,r) such that the following diagram is commutative:

i=1rLi|XD~νi𝕂XD~,rdet(E)1/r0θθ(q)i=1rLi|XD~KXD~νi𝕂XD~,rdet(E)01/rKXD~.\begin{CD}\bigoplus_{i=1}^{r}L_{i|X\setminus\widetilde{D}}@>{\bigoplus\nu_{i}}>{}>\mathbb{K}_{X\setminus\widetilde{D},r}\otimes\det(E)^{1/r}_{0}\\ @V{\theta}V{}V@V{\theta(q)}V{}V\\ \bigoplus_{i=1}^{r}L_{i|X\setminus\widetilde{D}}\otimes K_{X\setminus\widetilde{D}}@>{\bigoplus\nu_{i}}>{}>\mathbb{K}_{X\setminus\widetilde{D},r}\otimes\det(E)_{0}^{1/r}\otimes K_{X\setminus\widetilde{D}}.\end{CD}

The induced isomorphism det(νi):det(E)|XD~(det(E)01/r)r\det\bigl{(}\bigoplus\nu_{i}\bigr{)}:\det(E)_{|X\setminus\widetilde{D}}\simeq(\det(E)_{0}^{1/r})^{\otimes r} is equal to βρ01\beta\rho_{0}^{-1}, where β\beta is a nowhere vanishing holomorphic function.

Let XD~=λΛUλX\setminus\widetilde{D}=\bigcup_{\lambda\in\Lambda}U_{\lambda} be a locally finite covering by simply connected open subsets UλU_{\lambda}. There exists an rr-th root β1/rλ\beta^{1/r}_{\lambda} of β|Uλ\beta_{|U_{\lambda}}. If UλUμU_{\lambda}\cap U_{\mu}\neq\emptyset, then γλ,μ:=β1/rλ(β1/rμ)1\gamma_{\lambda,\mu}:=\beta^{1/r}_{\lambda}\cdot(\beta^{1/r}_{\mu})^{-1} induces a locally constant function UλUμGrU_{\lambda}\cap U_{\mu}\longrightarrow G_{r}\subset{\mathbb{C}}^{\ast}. We obtain γλ,μγμ,λ=1\gamma_{\lambda,\mu}\cdot\gamma_{\mu,\lambda}=1. If UλUμUνU_{\lambda}\cap U_{\mu}\cap U_{\nu}\neq\emptyset, then γλ,μγμ,νγν,λ=1\gamma_{\lambda,\mu}\cdot\gamma_{\mu,\nu}\cdot\gamma_{\nu,\lambda}=1. We obtain a holomorphic line bundle \mathcal{L} on XD~X\setminus\widetilde{D} by gluing holomorphic line bundles 𝒪Uλeλ\mathcal{O}_{U_{\lambda}}\cdot e_{\lambda} (λΛ)(\lambda\in\Lambda) on UλU_{\lambda} via the relation eλ=γλ,μeμe_{\lambda}=\gamma_{\lambda,\mu}e_{\mu} on UλUμU_{\lambda}\cap U_{\mu}. Because γλ,μr=1\gamma_{\lambda,\mu}^{r}=1, we obtain an isomorphism ρ:r𝒪XD~\rho_{\mathcal{L}}:\mathcal{L}^{\otimes\,r}\simeq\mathcal{O}_{X\setminus\widetilde{D}} by eλr1e_{\lambda}^{\otimes r}\longmapsto 1.

For each λ\lambda, we obtain an isomorphism φi,λ:=βλ1/rνi|Uλeλ:Li|Uλ(KXD~(r+12i)/2det(E)1/r0)|Uλ|Uλ\varphi_{i,\lambda}:=\beta_{\lambda}^{-1/r}\nu_{i|U_{\lambda}}\otimes e_{\lambda}:L_{i|U_{\lambda}}\simeq\bigl{(}K_{X\setminus\widetilde{D}}^{(r+1-2i)/2}\otimes\det(E)^{1/r}_{0}\bigr{)}_{|U_{\lambda}}\otimes\mathcal{L}_{|U_{\lambda}}. By the construction, the morphisms φi,λ\varphi_{i,\lambda} (λΛ)(\lambda\in\Lambda) determine an isomorphism φi:Li|XD~KXD~(r+12i)/2det(E)01/r\varphi_{i}:L_{i|X\setminus\widetilde{D}}\simeq K_{X\setminus\widetilde{D}}^{(r+1-2i)/2}\otimes\det(E)_{0}^{1/r}\otimes\mathcal{L}. By setting (det(E)1/r,ρ)=(det(E)01/r,ρ0ρ)(\det(E)^{1/r},\rho)=(\det(E)_{0}^{1/r}\otimes\mathcal{L},\rho_{0}\otimes\rho_{\mathcal{L}}), we obtain the claim of the lemma. ∎

In the following, we choose (det(E)1/r,ρ)(\det(E)^{1/r},\rho) as in Lemma 7.8. Let hdet(E)1/rh_{\det(E)^{1/r}} be the flat metric of det(E)1/r\det(E)^{1/r} determined by hdet(E)1/rr=hdet(E)|XD~h_{\det(E)^{1/r}}^{\otimes r}=h_{\det(E)|X\setminus\widetilde{D}}. We obtain the bijection

Υ:Harm(q|XD~)HarmGr((E,θ)|XD~,hdet(E)|XD~)\Upsilon:\mathop{\rm Harm}\nolimits(q_{|X\setminus\widetilde{D}})\simeq\mathop{\rm Harm}\nolimits^{G_{r}}\bigl{(}(E,\theta)_{|X\setminus\widetilde{D}},h_{\det(E)|X\setminus\widetilde{D}}\bigr{)}

determined by Υ(h)=hhdet(E)1/r\Upsilon(h)=h\otimes h_{\det(E)^{1/r}} under the isomorphism in Lemma 7.8.

Proposition 7.9.

Let NN be any relatively compact open neighbourhood of D~\widetilde{D} in XX.

  • For any h1,h2HarmGr(E,θ,hdet(E);D,c)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits), the restrictions h1|XNh_{1|X\setminus N} and h2|XNh_{2|X\setminus N} are mutually bounded.

  • For any hHarmGr(E,θ,hdet(E);D,c)h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits), the metrics g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded, and |q|XN|g(h)i|q_{|X\setminus N}|_{g(h)_{i}} are bounded.

Proof   We may assume that the boundary of NN in XX is smooth and compact. We obtain the first claim by applying [24, Proposition 3.29] to Υ1(hi)|XN\Upsilon^{-1}(h_{i})_{|X\setminus N}. For any hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}), we obtain g(h)i|XD~=g(Υ1(h|XD~))ig(h)_{i|X\setminus\widetilde{D}}=g(\Upsilon^{-1}(h_{|X\setminus\widetilde{D}}))_{i} (i=1,,r1)(i=1,\ldots,r-1). Hence, we obtain the second claim from [24, Proposition 3.27]. ∎

Note that we may naturally regard |ψi|h2|\psi_{i}|_{h}^{2} as real sections of KXKX¯K_{X}\otimes\overline{K_{X}}. Note that g(h)i=|ψi|h2g(h)_{i}=|\psi_{i}|_{h}^{2} (i=1,,r1)(i=1,\ldots,r-1). The second claim of Proposition 7.9 is reworded as follows.

Corollary 7.10.

Let hHarmGr(E,θ,hdet(E);D,c)h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits). For any relatively compact open neighbourhood NN of D~\widetilde{D} in XX, the metrics (|ψi|h2)|XN(|\psi_{i}|_{h}^{2})_{|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded, and (|ψr|h2/|ψi|h2)|XN\bigl{(}|\psi_{r}|_{h}^{2}/|\psi_{i}|_{h}^{2}\bigr{)}_{|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are bounded. ∎

7.2.4. Existence and uniqueness

By Proposition 7.5, we obtain the following map:

HarmGr(E,θ,hdet(E))PD>0𝒫(q,P,𝒗P)×I𝒮(q)𝒫.\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)})\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}(q,P,{\boldsymbol{v}}_{P})\times\prod_{I\in\mathcal{S}(q)}\mathcal{P}. (132)

We obtain the following map as the restriction of (132):

HarmGr(E,θ,hdet(E);D,c)PD>0𝒫(q,P,𝒗P)×I𝒮(q)𝒫.\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits)\longrightarrow\prod_{P\in D_{>0}}\mathcal{P}(q,P,{\boldsymbol{v}}_{P})\times\prod_{I\in\mathcal{S}(q)}\mathcal{P}. (133)
Theorem 7.11.

The map (133) is a bijection. In particular, if XX is compact, the map (132) is a bijection.

Proof   Suppose that h1,h2HarmGr(E,θ,hdet(E);D,c)h_{1},h_{2}\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)};D,\mathop{\rm c}\nolimits) satisfy 𝒃P(h1)=𝒃P(h2){\boldsymbol{b}}_{P}(h_{1})={\boldsymbol{b}}_{P}(h_{2}) for any PD>0P\in D_{>0} and 𝒂I(h1)=𝒂I(h2){\boldsymbol{a}}_{I}(h_{1})={\boldsymbol{a}}_{I}(h_{2}) for any I𝒮(q)I\in\mathcal{S}(q). By Proposition 7.6 and Proposition 7.9, we obtain that h1h_{1} and h2h_{2} are mutually bounded. By applying the argument in the proof of Theorem 6.6, we obtain h1=h2h_{1}=h_{2}.

Let 𝒃P{\boldsymbol{b}}_{P} (PD>0)(P\in D_{>0}) and 𝒂I{\boldsymbol{a}}_{I} (I𝒮(q))(I\in\mathcal{S}(q)). Let NN be a relatively compact open neighbourhood of DD in XZ(qr1)X\setminus Z(q_{\leq r-1}). By using the argument in the proof of Proposition 7.5, we can construct

hNHarm((E,θ,hdet(E))|XZ(qr1))h_{N}\in\mathop{\rm Harm}\nolimits((E,\theta,h_{\det(E)})_{|X\setminus Z(q_{\leq r-1})})

such that 𝒃P(hN)=𝒃P{\boldsymbol{b}}_{P}(h_{N})={\boldsymbol{b}}_{P} for PD>0P\in D_{>0} and 𝒂I(hN)=𝒂I{\boldsymbol{a}}_{I}(h_{N})={\boldsymbol{a}}_{I} for I𝒮(q)I\in\mathcal{S}(q). Let N1N_{1} be a relatively compact neighbourhood of DD in NN. Let N2N_{2} be a relatively compact neighbourhood of Z(qr1)Z(q_{\leq r-1}). We set hcE,θ:=Υ(hc)h^{\mathop{\rm c}\nolimits}_{E,\theta}:=\Upsilon(h^{\mathop{\rm c}\nolimits}), where hch^{\mathop{\rm c}\nolimits} denotes the unique complete metric in Harm(q|XD~)\mathop{\rm Harm}\nolimits(q_{|X\setminus\widetilde{D}}). (See §7.2.3 for Υ\Upsilon.) Then, there exists a GrG_{r}-invariant Hermitian metric h0h_{0} of EE such that (i) h0|N1D=hN|N1Dh_{0|N_{1}\setminus D}=h_{N|\setminus N_{1}\setminus D}, (ii) h0|X(NN2)=hcE,θ|X(NN2)h_{0|X\setminus(N\cup N_{2})}=h^{\mathop{\rm c}\nolimits}_{E,\theta|X\setminus(N\cup N_{2})}, (iii) det(h0)=hdet(E)\det(h_{0})=h_{\det(E)}. By Proposition 3.18, we obtain hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) which is mutually bounded with h0h_{0}, which satisfies 𝒃P(h)=𝒃P{\boldsymbol{b}}_{P}(h)={\boldsymbol{b}}_{P} (PD>0)(P\in D_{>0}) and 𝒂I(h)=𝒂I{\boldsymbol{a}}_{I}(h)={\boldsymbol{a}}_{I} (I𝒮(q))(I\in\mathcal{S}(q)). ∎

Remark 7.12.

This type of classification is well known in the case where XX is compact and DD is empty. Indeed, under the assumption that q0q\neq 0, (E,θ)(E,\theta) is stable with respect to the GrG_{r}-action. Hence, for a prescribed Hermitian metric hdet(E)h_{\det(E)} of det(E)\det(E), there uniquely exists a GrG_{r}-invariant Hermitian metric hh of EE such that (i) hh is Hermitian-Einstein in the sense that the trace-free part of F(h,θ)F(h,\theta)^{\bot} is 0, (ii) det(h)=hdet(E)\det(h)=h_{\det(E)}. If moreover c1(E)=0c_{1}(E)=0, we may choose a flat metric hdet(E)h_{\det(E)} of det(E)\det(E), then such hh is a harmonic metric.

We obtain the following theorem as a consequence of Theorem 7.11 and Corollary 7.10.

Theorem 7.13.

Suppose DD is empty. Let NXN\subset X be a relatively compact open set containing all zeros of qr1q_{\leq r-1}. There uniquely exists a metric hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) such that the metrics |ψi|h2|\psi_{i}|_{h}^{2} (i=1,,r1)(i=1,\cdots,r-1) are complete on XNX\setminus N. Moreover, on XNX\setminus N, |ψi|2h|\psi_{i}|^{2}_{h} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded, and |ψr|2h|ψi|2h\frac{|\psi_{r}|^{2}_{h}}{|\psi_{i}|^{2}_{h}} (i=1,,r1)(i=1,\cdots,r-1) are bounded. ∎

7.2.5. Boundedness at infinity

Let gg be a complete Kähler metric of XX. Let NN be a relatively compact open neighbourhood of D~\widetilde{D}. We say that hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) is bounded at infinity of XX with respect to gg if there exists 0<δ=δ(N)<10<\delta=\delta(N)<1 such that on XNX\setminus N

δ|ψi|h,g2δ1(i=1,,r1).\delta\leq|\psi_{i}|_{h,g}^{2}\leq\delta^{-1}\quad(i=1,\ldots,r-1). (134)

In other words, g(h)ig(h)_{i} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded with gg on XNX\setminus N.

Lemma 7.14.

Suppose that there exists hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) which is bounded at infinity of XX. Then, hh is complete at infinity of XX. Moreover, |q|g|q|_{g} is bounded on XNX\setminus N, and hence |ψr|h,g|\psi_{r}|_{h,g} is bounded on XNX\setminus N.

Proof   The first claim is clear by definition. The boundedness of |q|g|q|_{g} on XNX\setminus N follows from Proposition 7.9. ∎

We obtain the following uniqueness up to boundedness.

Corollary 7.15.

Suppose that hiHarmGr(E,θ,hdet(E))h_{i}\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) (i=1,2)(i=1,2) are bounded at infinity of XX. Then, h1h_{1} and h2h_{2} are mutually bounded on XNX\setminus N. If moreover DD is empty, we obtain h1=h2h_{1}=h_{2} on XX.

Proof   The first claim follows from Proposition 7.9 and Lemma 7.14. The second claim follows from Theorem 7.13 and Lemma 7.14. ∎

A bounded solution does not necessarily exist for a prescribed complete Kähler metric.

Corollary 7.16.

Let gg^{\prime} be another complete Kähler metric of XX. If there exist h,hHarmGr(E,θ,hdet(E))h,h^{\prime}\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) such that hh (resp. hh^{\prime}) is bounded at infinity of XX with respect to gg (resp. gg^{\prime}), then gg and gg^{\prime} are mutually bounded on XX.

Proof   By Proposition 7.9 and Lemma 7.14, hh and hh^{\prime} are mutually bounded on XNX\setminus N. It implies that |g(h)i|2|g(h)_{i}|^{2} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded with gg and gg^{\prime} on XNX\setminus N. Hence, we obtain that gg and gg^{\prime} are mutually bounded on XX. ∎

Proposition 7.17.

Let gg be a complete hyperbolic metric of XX. Assume that |q|g|q|_{g} is bounded on XNX\setminus N. Then, hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) is bounded at infinity of XX with respect to gg if and only if hh is complete at infinity.

Proof   The “only if” part of the claim is clear by definition. Let us prove that the “if” part of the claim. Let KK be a compact neighbourhood of D~\widetilde{D} in NN with smooth boundary. Note that XKX\setminus K is hyperbolic.

Lemma 7.18.

Let gXKg_{X\setminus K} be a complete hyperbolic metric of XKX\setminus K.

  • ggXKg\leq g_{X\setminus K} on XKX\setminus K.

  • gg and gXKg_{X\setminus K} are mutually bounded on XNX\setminus N.

Proof   We may assume that the Gaussian curvature of gXKg_{X\setminus K} and gg are constantly 2-2. Recall that they correspond to complete solutions of the Toda equation (1) with r=2r=2 and q=0q=0 on XKX\setminus K and XX, respectively. (See also §1.1.5.) The first claim follows from [24, Theorem 1.7]. The second claim of the lemma follows from [24, Proposition 3.29]. ∎

Because |q|g|q|_{g} is bounded on XX, |q|XK|gXK|q_{|X\setminus K}|_{g_{X\setminus K}} is bounded. Let hcXKHarm(q|XK)h^{\mathop{\rm c}\nolimits}_{X\setminus K}\in\mathop{\rm Harm}\nolimits(q_{|X\setminus K}) be the complete solution on XKX\setminus K. According to [24, Theorem 1.8], g(hcXK)ig(h^{\mathop{\rm c}\nolimits}_{X\setminus K})_{i} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded with gXKg_{X\setminus K}. Hence, g(hcXK)i|XNg(h^{\mathop{\rm c}\nolimits}_{X\setminus K})_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded with g|XNg_{|X\setminus N}.

Suppose that hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}) is complete at infinity. By [24, Proposition 3.29], hcXK|XNh^{\mathop{\rm c}\nolimits}_{X\setminus K|X\setminus N} and Υ1(h|XN)\Upsilon^{-1}(h_{|X\setminus N}) are mutually bounded. (See §7.2.3 for Υ\Upsilon.) Hence, g(h)i|XNg(h)_{i|X\setminus N} (i=1,,r1)(i=1,\ldots,r-1) are mutually bounded with g|XNg_{|X\setminus N}, i.e., hh is bounded at infinity with respect to gg. ∎

We obtain the following theorem as a consequence of Theorem 7.13, Lemma 7.14, and Proposition 7.17.

Theorem 7.19.

Suppose that DD is empty. Let gg be a complete hyperbolic metric of XX. If qq is bounded with respect to gg, there uniquely exists a bounded solution hHarmGr(E,θ,hdet(E))h\in\mathop{\rm Harm}\nolimits^{G_{r}}(E,\theta,h_{\det(E)}). Conversely, if there exists such a bounded solution hh, qq is bounded with respect to gg. ∎

Remark 7.20.

According to Proposition 7.17, if gg is complete hyperbolic and if |q|g|q|_{g} is bounded on XNX\setminus N, we may replace the condition “completeness at infinity” in Theorem 7.11 with the condition “boundedness at infinity with respect to gg”.

8. Appendix

8.1. The case of trivial rr-differentials

8.1.1. Rank 22 case

Let XX be a Riemann surface with a finite subset DD. Suppose that XDX\setminus D has a Kähler metric gXDg_{X\setminus D} with infinite volume such that the Gaussian curvature of the Riemannian metric Re(gXD)\mathop{\rm Re}\nolimits(g_{X\setminus D}) is constantly k-k (k>0)(k>0).

Proposition 8.1.

For any 𝐚=(aP)PDD0{\boldsymbol{a}}=(a_{P})_{P\in D}\in{\mathbb{R}}^{D}_{\geq 0}, there exists a Kähler metric g𝐚g_{{\boldsymbol{a}}} of XDX\setminus D whose Gaussian curvature is constantly k-k, such that the following holds.

  • For any neighbourhood NN of DD, g𝒂|XNg_{{\boldsymbol{a}}|X\setminus N} and gXD|XNg_{X\setminus D|X\setminus N} are mutually bounded.

  • Let (XP,zP)(X_{P},z_{P}) be a coordinate neighbourhood around PP such that zP(P)=0z_{P}(P)=0. If aP>0a_{P}>0, then g𝒂|XP{P}|zP|2(aP1)g_{{\boldsymbol{a}}|X_{P}\setminus\{P\}}|z_{P}|^{-2(a_{P}-1)} is mutually bounded with dzPdz¯Pdz_{P}\,d\overline{z}_{P} around PP. If aP=0a_{P}=0, then g𝒂|XP{P}g_{{\boldsymbol{a}}|X_{P}\setminus\{P\}} is mutually bounded with |zP|2(log|zP|)2dzPdz¯P|z_{P}|^{-2}(\log|z_{P}|)^{-2}dz_{P}\,d\overline{z}_{P} around PP.

Proof   It is enough to study the case where k=2k=2. Recall that in this paper, for a Kähler metric gg on XDX\setminus D, we use the Hermitian metric on KX/2K_{X}^{\ell/2} denoted as g/2g^{-\ell/2}, determined as follows.

  • If g=g0dzdz¯g=g_{0}\,dz\otimes d\overline{z} for a local holomorphic coordinate zz, then |(dz)/2|2g/2=(g02)/2|(dz)^{\ell/2}|^{2}_{g^{-\ell/2}}=(\frac{g_{0}}{2})^{-\ell/2}.

Let ωg\omega_{g} denote the associated Kähler form, which is locally 12g0dzdz¯\frac{\sqrt{-1}}{2}g_{0}dz\wedge d\overline{z}.

We set 𝕂XD,2:=KXD1/2KXD1/2\mathbb{K}_{X\setminus D,2}:=K_{X\setminus D}^{1/2}\oplus K_{X\setminus D}^{-1/2}. Let 0XD,20_{X\setminus D,2} denote the quadratic differential on XDX\setminus D which is constantly 0. It induces the Higgs field θ(0XD,2)\theta(0_{X\setminus D,2}) of 𝕂XD,2\mathbb{K}_{X\setminus D,2}. We consider the action of G2={±1}G_{2}=\{\pm 1\} on KXD±1/2K_{X\setminus D}^{\pm 1/2} given by ax=a±1xa\bullet x=a^{\pm 1}x, which induces a G2G_{2}-action on 𝕂XD,2\mathbb{K}_{X\setminus D,2}. A Kähler metric gg of XDX\setminus D induces a Hermitian metric h(1)(g)=g1/2g1/2h^{(1)}(g)=g^{-1/2}\oplus g^{1/2}.

Lemma 8.2.

h(1)(g)h^{(1)}(g) is a harmonic metric if and only if the Gaussian curvature of Re(g)\mathop{\rm Re}\nolimits(g) is constantly 2-2.

Proof   Let R(g)R(g) denote the curvature of the Chern connection of (K1XD,g)(K^{-1}_{X\setminus D},g). Then, by a direct calculation, we can check that h(1)h^{(1)} is harmonic if and only if 1R(g)+2ωg=0\sqrt{-1}R(g)+2\omega_{g}=0, which implies that the Gaussian curvature of gg is 2-2. ∎

Let Harm(0XD,2)\mathop{\rm Harm}\nolimits(0_{X\setminus D,2}) denote the set of G2G_{2}-invariant harmonic metrics hh of (𝕂XD,2,θ(0XD,2))(\mathbb{K}_{X\setminus D,2},\theta(0_{X\setminus D,2})) such that det(h)=1\det(h)=1. Note that hh is decomposed as h=h|KXD1/2h|KXD1/2h=h_{|K_{X\setminus D}^{1/2}}\oplus h_{|K_{X\setminus D}^{-1/2}}, and it is equal to h(1)(g)h^{(1)}(g) for a Kähler metric gg of XDX\setminus D whose Gaussian curvature is 2-2.

Let us give some preliminaries on hyperbolic metrics on XP:=XP{P}X_{P}^{\ast}:=X_{P}\setminus\{P\} (PD)(P\in D). We may assume that XP={|zP|<1/2}X_{P}=\{|z_{P}|<1/2\}. It is well known and easy to check that gP,1:=2(1|zP|2)2dzPdz¯Pg_{P,1}:=2(1-|z_{P}|^{2})^{-2}dz_{P}\,d\overline{z}_{P} is a hyperbolic metric of XPX_{P} normalized as 1R(gP,1)+2ωgP,1=0\sqrt{-1}R(g_{P,1})+2\omega_{g_{P,1}}=0. For a>0a>0, we set as follows on XPX_{P}^{\ast}:

gP,a=2dzPadz¯Pa(1|zP|2a)2=2a2|zP|2(a1)dzPdz¯P(1|zP|2a)2.g_{P,a}=\frac{2dz_{P}^{a}\,d\overline{z}_{P}^{a}}{(1-|z_{P}|^{2a})^{2}}=\frac{2a^{2}|z_{P}|^{2(a-1)}\,dz_{P}\,d\overline{z}_{P}}{(1-|z_{P}|^{2a})^{2}}.

Because gP,ag_{P,a} is locally obtained as the pull back of gP,1g_{P,1} by the map zPzPaz_{P}\longmapsto z_{P}^{a}, it is a hyperbolic metric on XPX_{P}^{\ast} normalized as 1R(gP,a)+2ωgP,a=0\sqrt{-1}R(g_{P,a})+2\omega_{g_{P,a}}=0. We also set

gP,0:=2dzPdz¯P|zP|2(log|zP|2)2.g_{P,0}:=\frac{2dz_{P}\,d\overline{z}_{P}}{|z_{P}|^{2}(\log|z_{P}|^{2})^{2}}.

It is easy to check 1R(gP,0)+2ωgP,0=0\sqrt{-1}R(g_{P,0})+2\omega_{g_{P,0}}=0.

Lemma 8.3.

Let gPg_{P} be any hyperbolic metric of XPX_{P}^{\ast} normalized as 1R(gP)+2ωgP=0\sqrt{-1}R(g_{P})+2\omega_{g_{P}}=0.

  • There exists a0a\geq 0 such that gPg_{P} and gP,ag_{P,a} are mutually bounded around PP.

  • Let XP,1X_{P,1} be a relatively compact neighbourhood of PP in XPX_{P}. Then, the volume of XP,1{P}X_{P,1}\setminus\{P\} is finite with respect to gPg_{P}.

Proof   Let hHarm(0XP,2)h\in\mathop{\rm Harm}\nolimits(0_{X_{P}^{\ast},2}). Because θ(0XP,2)\theta(0_{X_{P}^{\ast},2}) is nilpotent, the harmonic bundle (𝕂XP,2,θ(0XP,2),h)(\mathbb{K}_{X_{P}^{\ast},2},\theta(0_{X_{P}^{\ast},2}),h) is tame. As in the proof of Proposition 3.21, we obtain the locally free 𝒪XP\mathcal{O}_{X_{P}}-modules

𝒫ha𝕂XP,2=𝒫haK1/2XP𝒫haK1/2XP(a),\mathcal{P}^{h}_{a}\mathbb{K}_{X_{P}^{\ast},2}=\mathcal{P}^{h}_{a}K^{1/2}_{X_{P}^{\ast}}\oplus\mathcal{P}^{h}_{a}K^{-1/2}_{X_{P}^{\ast}}\quad(a\in{\mathbb{R}}),

and we can prove that there exists bb\in{\mathbb{R}} such that zP(dzP)1/2z_{P}(dz_{P})^{1/2} and zP2(dzP)1/2z_{P}^{2}(dz_{P})^{-1/2} are sections of 𝒫hb𝕂XP,2\mathcal{P}^{h}_{b}\mathbb{K}_{X_{P}^{\ast},2}. We obtain the numbers

bi(h):=inf{b||zP|b|zPi(dzP)(32i)/2|h is bounded}(i=1,2).b_{i}(h):=\inf\bigl{\{}b\in{\mathbb{R}}\,\big{|}\,\mbox{$|z_{P}|^{b}\cdot|z_{P}^{i}(dz_{P})^{(3-2i)/2}|_{h}$ is bounded}\bigr{\}}\quad(i=1,2).

Because θ(0XP,2)\theta(0_{X_{P}^{\ast},2}) is nilpotent, we obtain

θ(0XP,2)=O(|log|zP||1)dzP/zP\theta(0_{X_{P}^{\ast},2})=O\bigl{(}\bigl{|}\log|z_{P}|\bigr{|}^{-1}\bigr{)}dz_{P}/z_{P}

with respect to hh, according to [39]. Because θ(0XP,2)(zP(dzP)1/2)=zP2(dzP)1/2(dzP/zP)\theta(0_{X_{P}^{\ast},2})(z_{P}(dz_{P})^{1/2})=z_{P}^{2}(dz_{P})^{-1/2}\cdot(dz_{P}/z_{P}), we obtain b1(h)b2(h)b_{1}(h)\geq b_{2}(h). Because det(h)=1\det(h)=1, we obtain b1(h)+b2(h)+3=0b_{1}(h)+b_{2}(h)+3=0. Moreover, according to the norm estimate of Simpson [39], if hjHarm(0XP,2)h_{j}\in\mathop{\rm Harm}\nolimits(0_{X_{P}^{\ast},2}) (j=1,2)(j=1,2) satisfy bi(h1)=bi(h2)b_{i}(h_{1})=b_{i}(h_{2}) (i=1,2)(i=1,2) then h1h_{1} and h2h_{2} are mutually bounded.

For a hyperbolic metric gPg_{P} of XPX_{P}^{\ast} normalized as 1R(gP)+2ωgP=0\sqrt{-1}R(g_{P})+2\omega_{g_{P}}=0, we obtain h(1)(gP)Harm(0XP,2)h^{(1)}(g_{P})\in\mathop{\rm Harm}\nolimits(0_{X_{P}^{\ast},2}) induced by gPg_{P}. For any (b1,b2)2(b_{1},b_{2})\in{\mathbb{R}}^{2} such that b1b2b_{1}\geq b_{2} and b1+b2+3=0b_{1}+b_{2}+3=0, there exists a0a\in{\mathbb{R}}_{\geq 0} such that b1=a232b_{1}=\frac{a}{2}-\frac{3}{2} and b2=a232b_{2}=-\frac{a}{2}-\frac{3}{2}. Then, we can directly check that bi(h(1)(gP,a))=bib_{i}\bigl{(}h^{(1)}(g_{P,a})\bigr{)}=b_{i}. Then, we obtain the first claim. The second claim can be checked in the case of gP,ag_{P,a} (a0)(a\in{\mathbb{R}}_{\geq 0}). ∎

Take any 𝒂D0{\boldsymbol{a}}\in{\mathbb{R}}^{D}_{\geq 0}. We obtain hP,aP=h(1)(gP,aP)Harm(0XP,2)h_{P,a_{P}}=h^{(1)}(g_{P,a_{P}})\in\mathop{\rm Harm}\nolimits(0_{X_{P}^{\ast},2}) corresponding to gP,aPg_{P,a_{P}}. Let hXDHarm(0XD,2)h_{X\setminus D}\in\mathop{\rm Harm}\nolimits(0_{X\setminus D,2}) on XDX\setminus D corresponding to the hyperbolic metric gXDg_{X\setminus D}. Let X1,PX_{1,P} (PD)(P\in D) be relatively compact neighbourhoods of PP in XPX_{P}. There exists a G2G_{2}-invariant Hermitian metric h0h_{0} such that (i) h0=hP,aPh_{0}=h_{P,a_{P}} on X1,PX_{1,P}, (ii) h0=hXDh_{0}=h_{X\setminus D} on XPDXPX\setminus\coprod_{P\in D}X_{P}, (iii) det(h0)=1\det(h_{0})=1. Note that the support of F(h0)=R(h0)+[θ(0XD,2),θ(0XD,2)h0]F(h_{0})=R(h_{0})+[\theta(0_{X\setminus D,2}),\theta(0_{X\setminus D,2})^{\dagger}_{h_{0}}] is compact. We also note that K1/2XDK^{-1/2}_{X\setminus D} is the unique proper Higgs subbundle of (𝕂XD,2,θ(0XD,2))\bigl{(}\mathbb{K}_{X\setminus D,2},\theta(0_{X\setminus D,2})\bigr{)}.

Let h1h_{1} denote the Hermitian metric of K1/2XDK^{-1/2}_{X\setminus D} induced by h0h_{0}. Because rankK1/2XD=1\mathop{\rm rank}\nolimits K^{-1/2}_{X\setminus D}=1, we obtain F(h1)=R(h1)F(h_{1})=R(h_{1}). Let us prove that

XD1R(h1)=.\int_{X\setminus D}\sqrt{-1}R(h_{1})=-\infty. (135)

Let h2h_{2} denote the Hermitian metric of K1/2XDK^{-1/2}_{X\setminus D} induced by the hyperbolic metric gXDg_{X\setminus D}. By the assumption, we have

XD1R(h2)=2XDωgXD=.\int_{X\setminus D}\sqrt{-1}R(h_{2})=-2\int_{X\setminus D}\omega_{g_{X\setminus D}}=-\infty.

By the construction, h2=h1h_{2}=h_{1} holds on XXPX\setminus\coprod X_{P}. By Lemma 8.3, the volume of PD(XPD)\coprod_{P\in D}(X_{P}\setminus D) is finite with respect to gXDg_{X\setminus D}. We obtain

XXP1R(h1)=XXP1R(h2)=.\int_{X\setminus\coprod X_{P}}\sqrt{-1}R(h_{1})=\int_{X\setminus\coprod X_{P}}\sqrt{-1}R(h_{2})=-\infty.

Because h1|X1,Ph_{1|X_{1,P}} (PD)(P\in D) are induced by gP,aPg_{P,a_{P}}, we obtain

X1,PD1R(h1)=X1,PD2ωgP,𝒂<0.\int_{X_{1,P}\setminus D}\sqrt{-1}R(h_{1})=\int_{X_{1,P}\setminus D}-2\omega_{g_{P,{\boldsymbol{a}}}}<0.

Because PD(XPX1,P)\coprod_{P\in D}(X_{P}\setminus X_{1,P}) is compact, we obtain

PDXPX1,P1R(h1)<.\sum_{P\in D}\int_{X_{P}\setminus X_{1,P}}\sqrt{-1}R(h_{1})<\infty.

In all, we obtain (135). As a result, (𝕂XD,2,θ(0XD,2),h0)(\mathbb{K}_{X\setminus D,2},\theta(0_{X\setminus D,2}),h_{0}) is analytically stable. By Proposition 2.11, there exists h𝒂Harm(0XD,2)h_{{\boldsymbol{a}}}\in\mathop{\rm Harm}\nolimits(0_{X\setminus D,2}) such that h𝒂h_{{\boldsymbol{a}}} and h0h_{0} are mutually bounded on XDX\setminus D. The Kähler metric g𝒂g_{{\boldsymbol{a}}} on KXD1K_{X\setminus D}^{-1} induced by h𝒂h_{{\boldsymbol{a}}} satisfies the desired conditions. ∎

Remark 8.4.

If there exists a complete Kähler metric gXDg_{X\setminus D} of XDX\setminus D with finite volume, then there exist a compact Riemann surface X¯\overline{X} and an open embedding XDX¯X\setminus D\longrightarrow\overline{X} whose complement is a finite subset. (For example, see [11, 40], where much more general results are proved.) Therefore, we may apply the theory of tame harmonic bundles [39] due to Simpson to study a problem as in Proposition 8.1. The stability condition provides us with a constraint on 𝐚{\boldsymbol{a}}.

8.1.2. Rank rr case

Let rr be a positive integer larger than 22. Let 0XD,r0_{X\setminus D,r} denote the rr-differential on XDX\setminus D which is constantly 0. Let hHarm(0XD,r)h\in\mathop{\rm Harm}\nolimits(0_{X\setminus D,r}). Because θ(0XD,r)\theta(0_{X\setminus D,r}) is nilpotent, the harmonic bundle (𝕂XD,r,θ(0XD,r),h)(\mathbb{K}_{X\setminus D,r},\theta(0_{X\setminus D,r}),h) is tame. As in the proof of Proposition 3.21, from (𝕂XP,r,θ(0XP,r),h|XP)(\mathbb{K}_{X_{P}^{\ast},r},\theta(0_{X_{P}^{\ast},r}),h_{|X_{P}^{\ast}}), we obtain the locally free 𝒪XP\mathcal{O}_{X_{P}}-modules 𝒫ha𝕂XP,r=i=1r𝒫ahK(r+12i)/2XP\mathcal{P}^{h}_{a}\mathbb{K}_{X_{P}^{\ast},r}=\bigoplus_{i=1}^{r}\mathcal{P}_{a}^{h}K^{(r+1-2i)/2}_{X_{P}^{\ast}} (a)(a\in{\mathbb{R}}), and we can prove that there exists bb\in{\mathbb{R}} such that zPi(dzP)(r+12i)/2z_{P}^{i}(dz_{P})^{(r+1-2i)/2} (i=1,,r)(i=1,\ldots,r) are sections of 𝒫hb𝕂XP,r\mathcal{P}^{h}_{b}\mathbb{K}_{X_{P}^{\ast},r}. For i=1,,ri=1,\ldots,r, we obtain the numbers

bP,i(h):=inf{b||zP|b|zPi(dz)P(r+1i)/2|h is bounded}.b_{P,i}(h):=\inf\bigl{\{}b\in{\mathbb{R}}\,\big{|}\,\mbox{$|z_{P}|^{b}\cdot|z_{P}^{i}(dz)_{P}^{(r+1-i)/2}|_{h}$ is bounded}\bigr{\}}.

Let 𝒃P(h){\boldsymbol{b}}_{P}(h) denote the tuple (bP,1(h),,bP,r(h))(b_{P,1}(h),\ldots,b_{P,r}(h)). Because θ(0XD,r)\theta(0_{X\setminus D,r}) is nilpotent, we obtain |θ(0XD,r)|h=O((log|zP|)1)dzP/zP|\theta(0_{X\setminus D,r})|_{h}=O\bigl{(}(-\log|z_{P}|)^{-1}\bigr{)}dz_{P}/z_{P} with respect to hh. Because

θ(0XP,r)(zPi(dzP)(r+12i)/2)=zPi+1(dzP)(r+12(i+1))/2(dzP/zP)\theta(0_{X_{P}^{\ast},r})(z_{P}^{i}(dz_{P})^{(r+1-2i)/2})=z_{P}^{i+1}(dz_{P})^{(r+1-2(i+1))/2}\cdot(dz_{P}/z_{P})

for i=1,,r1i=1,\ldots,r-1, we obtain bP,i(h)bP,i+1(h)b_{P,i}(h)\geq b_{P,i+1}(h) (i=1,,r1)(i=1,\ldots,r-1). Because det(h)=1\det(h)=1, we obtain bP,i(h)=r(r+1)/2\sum b_{P,i}(h)=-r(r+1)/2.

For simplicity, we assume that the Gaussian curvature of gXDg_{X\setminus D} is 2-2. Note that (𝕂XD,r,θ(0XD,r))(\mathbb{K}_{X\setminus D,r},\theta(0_{X\setminus D,r})) is isomorphic to the (r1)(r-1)-th symmetric product of (𝕂XD,2,θ(0XD,2))(\mathbb{K}_{X\setminus D,2},\theta(0_{X\setminus D,2})). Let hr(1)(gXD)Harm(0XD,r)h_{r}^{(1)}(g_{X\setminus D})\in\mathop{\rm Harm}\nolimits(0_{X\setminus D,r}) denote the harmonic metric induced by h(1)(gXD)h^{(1)}(g_{X\setminus D}).

Proposition 8.5.

Suppose that 𝐛P=(bP,1,,bP,r)r{\boldsymbol{b}}_{P}=(b_{P,1},\ldots,b_{P,r})\in{\mathbb{R}}^{r} (PD)(P\in D) satisfy

bP,1bP,r,bP,j=r(r+1)2.b_{P,1}\geq\cdots\geq b_{P,r},\quad\sum b_{P,j}=-\frac{r(r+1)}{2}.

Then, there exists hHarm(0XD,r)h\in\mathop{\rm Harm}\nolimits(0_{X\setminus D,r}) such that (i) 𝐛P(h)=𝐛P{\boldsymbol{b}}_{P}(h)={\boldsymbol{b}}_{P} (PD)(P\in D), (ii) For any neighbourhood NN of DD, hh and hr(1)(gXD)h_{r}^{(1)}(g_{X\setminus D}) are mutually bounded on XNX\setminus N.

Proof   Let us explain an outline of the proof.

Lemma 8.6.

There exists hP,𝐛PHarm(0XP,r)h_{P,{\boldsymbol{b}}_{P}}\in\mathop{\rm Harm}\nolimits(0_{X_{P}^{\ast},r}) such that 𝐛(hP,𝐛)=𝐛P{\boldsymbol{b}}(h_{P,{\boldsymbol{b}}})={\boldsymbol{b}}_{P}.

Proof   Let YY be a compact Riemann surface whose genus g(Y)g(Y) is larger than 10i|bP,i|+10r210\sum_{i}|b_{P,i}|+10r^{2}. Let QQ be a point of YY. Let (YQ,zQ)(Y_{Q},z_{Q}) be a holomorphic coordinate neighbourhood of QQ in YY such that zQ(Q)=0z_{Q}(Q)=0. For any aa\in{\mathbb{R}}, 𝒫a(𝕂YQ,r(Q))\mathcal{P}_{a}\bigl{(}\mathbb{K}_{Y_{Q},r}(\ast Q)\bigr{)} denote the locally free 𝒪YQ\mathcal{O}_{Y_{Q}}-module obtained as follows:

𝒫a(𝕂YQ,r(Q))=i=1r𝒪YQ([abP,i])zQi(dzQ)(r+12i)/2.\mathcal{P}_{a}\bigl{(}\mathbb{K}_{Y_{Q},r}(\ast Q)\bigr{)}=\bigoplus_{i=1}^{r}\mathcal{O}_{Y_{Q}}([a-b_{P,i}])z_{Q}^{i}(dz_{Q})^{(r+1-2i)/2}.

Here, for cc\in{\mathbb{R}}, [c][c] denotes max{n|nc}\max\{n\in{\mathbb{Z}}\,|\,n\leq c\}. From 𝒫a(𝕂YQ,r(Q))\mathcal{P}_{a}\bigl{(}\mathbb{K}_{Y_{Q},r}(\ast Q)\bigr{)} and 𝕂YQ,r\mathbb{K}_{Y\setminus Q,r}, we obtain locally free 𝒪Y\mathcal{O}_{Y}-modules 𝒫a(𝕂Y,r(Q))\mathcal{P}_{a}\bigl{(}\mathbb{K}_{Y,r}(\ast Q)\bigr{)}. Thus, we obtain a filtered bundle 𝒫(𝕂Y,r(Q))\mathcal{P}_{\ast}(\mathbb{K}_{Y,r}(\ast Q)) on (Y,Q)(Y,Q). We can easily check θ(0YQ,r)𝒫a(𝕂Y,r(Q))𝒫a(𝕂Y,r(Q))KY(Q)\theta(0_{Y\setminus Q,r})\mathcal{P}_{a}(\mathbb{K}_{Y,r}(\ast Q))\subset\mathcal{P}_{a}(\mathbb{K}_{Y,r}(\ast Q))\otimes K_{Y}(Q). We obtain

deg(𝒫(K(r+12i)/2Y(Q)))=r+12i2(2g(Y)2)(bP,i+i).\deg\Bigl{(}\mathcal{P}_{\ast}\bigl{(}K^{(r+1-2i)/2}_{Y}(\ast Q)\bigr{)}\Bigr{)}=\frac{r+1-2i}{2}(2g(Y)-2)-(b_{P,i}+i).

For any 1jr1\leq j\leq r, we obtain the following:

i=jrdeg(𝒫(K(r+12i)/2Y(Q)))=(j1)(r+1j)2(2g(Y)2)i=jr(bP,i+i).\sum_{i=j}^{r}\deg\Bigl{(}\mathcal{P}_{\ast}\bigl{(}K^{(r+1-2i)/2}_{Y}(\ast Q)\bigr{)}\Bigr{)}\\ =-\frac{(j-1)(r+1-j)}{2}(2g(Y)-2)-\sum_{i=j}^{r}(b_{P,i}+i). (136)

Hence, for any 1<jr1<j\leq r, we obtain

i=jrdeg(𝒫(K(r+12i)/2Y(Q)))<0=deg(𝒫(𝕂Y,r(Q))).\sum_{i=j}^{r}\deg\Bigl{(}\mathcal{P}_{\ast}\bigl{(}K^{(r+1-2i)/2}_{Y}(\ast Q)\bigr{)}\Bigr{)}<0=\deg\Bigl{(}\mathcal{P}_{\ast}\bigl{(}\mathbb{K}_{Y,r}(\ast Q)\bigr{)}\Bigr{)}.

If a non-zero subbundle E𝕂YQ,rE\subset\mathbb{K}_{Y\setminus Q,r} satisfies θ(0YQ,r)EEKYQ\theta(0_{Y\setminus Q,r})E\subset E\otimes K_{Y\setminus Q}, then there exists 1jr1\leq j\leq r such that E=i=jrKYQ(r+12i)/2E=\bigoplus_{i=j}^{r}K_{Y\setminus Q}^{(r+1-2i)/2}. Therefore, the regular filtered Higgs bundle (𝒫𝕂Y,r(Q),θ(0YQ,r))\bigl{(}\mathcal{P}_{\ast}\mathbb{K}_{Y,r}(\ast Q),\theta(0_{Y\setminus Q},r)\bigr{)} is stable. According to [39], there uniquely exists a harmonic metric hY,𝒃Ph_{Y,{\boldsymbol{b}}_{P}} of (𝕂YQ,r,θ(0YQ,r))(\mathbb{K}_{Y\setminus Q,r},\theta(0_{Y\setminus Q,r})) such that det(hY,𝒃P)=1\det(h_{Y,{\boldsymbol{b}}_{P}})=1 and that 𝕂YQ,r\mathbb{K}_{Y\setminus Q,r} with hY,𝒃Ph_{Y,{\boldsymbol{b}}_{P}} induces 𝒫(𝕂Y,r(Q))\mathcal{P}_{\ast}\bigl{(}\mathbb{K}_{Y,r}(\ast Q)\bigr{)}. By the uniqueness, we obtain that hY,𝒃Ph_{Y,{\boldsymbol{b}}_{P}} is GrG_{r}-invariant.

We embed XPX_{P} into YY by using zPz_{P} and (YQ,zQ)(Y_{Q},z_{Q}). Then, we can construct hP,𝒃Ph_{P,{\boldsymbol{b}}_{P}} as the pull back of hY,𝒃Ph_{Y,{\boldsymbol{b}}_{P}}. ∎

Let h0h_{0} be a GrG_{r}-invariant Hermitian metric of 𝕂XD,r\mathbb{K}_{X\setminus D,r} satisfying the following conditions.

  • h0=hr(1)(gXD)h_{0}=h_{r}^{(1)}(g_{X\setminus D}) on XPDXPX\setminus\coprod_{P\in D}X_{P}.

  • There exist relatively compact neighbourhoods X1,PX_{1,P} (PD)(P\in D) in XPX_{P} such that h0=hP,𝒃Ph_{0}=h_{P,{\boldsymbol{b}}_{P}} on X1,P{P}X_{1,P}\setminus\{P\}.

  • det(h0)=1\det(h_{0})=1.

Note that the support of F(h0)F(h_{0}) is compact.

Lemma 8.7.

(𝕂XD,r,θ(0XD,r),h0)(\mathbb{K}_{X\setminus D,r},\theta(0_{X\setminus D,r}),h_{0}) is analytically stable with respect to the GrG_{r}-action.

Proof   For a GrG_{r}-invariant Hermitian metric hh of 𝕂XD,r\mathbb{K}_{X\setminus D,r}, let hjh_{\geq j} (1jr)(1\leq j\leq r) denote the induced Hermitian metric of i=jrK(r+12i)/2XD\bigoplus_{i=j}^{r}K^{(r+1-2i)/2}_{X\setminus D}. Because the volume of XDX\setminus D with respect to gXDg_{X\setminus D} is infinite, we obtain the following for 1<jr1<j\leq r:

XD1ΛTrF(hr(1)(gXD)j)=XD1ΛTrR(hr(1)(gXD)j)=.\int_{X\setminus D}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits F\bigl{(}h_{r}^{(1)}(g_{X\setminus D})_{\geq j}\bigr{)}=\\ \int_{X\setminus D}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits R(h_{r}^{(1)}\bigl{(}g_{X\setminus D})_{\geq j}\bigr{)}=-\infty. (137)

By Lemma 8.3, and by h0|XXP=hr(1)(gXD)|XXPh_{0|X\setminus\coprod X_{P}}=h_{r}^{(1)}(g_{X\setminus D})_{|X\setminus\coprod X_{P}}, we obtain

XXP1ΛTrF((h0)j)=XXP1ΛTrR((h0)j)=.\int_{X\setminus\coprod X_{P}}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits F\bigl{(}(h_{0})_{\geq j}\bigr{)}=\int_{X\setminus\coprod X_{P}}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits R\bigl{(}(h_{0})_{\geq j}\bigr{)}=-\infty. (138)

On X1,PX_{1,P}, h0=hP,𝒃Ph_{0}=h_{P,{\boldsymbol{b}}_{P}} holds by the construction. Because hP,𝒃Ph_{P,{\boldsymbol{b}}_{P}} is harmonic, we obtain F(hP,𝒃P)=0F(h_{P,{\boldsymbol{b}}_{P}})=0. Let πj\pi_{j} denote the projection of 𝕂X1,P,r\mathbb{K}_{X_{1,P},r} onto i=jrK(r+12i)/2X1,P\bigoplus_{i=j}^{r}K^{(r+1-2i)/2}_{X_{1,P}^{\ast}}. It is the orthogonal projection with respect to h0h_{0}. By the Chern-Weil formula [38, Lemma 3.2], we obtain

X1,P1ΛTrF((h0)j)=X1,P|(¯E+θ)(πj)|2h00.\int_{X_{1,P}}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits F\bigl{(}(h_{0})_{\geq j}\bigr{)}=-\int_{X_{1,P}}\bigl{|}(\overline{\partial}_{E}+\theta)(\pi_{j})\bigr{|}^{2}_{h_{0}}\leq 0. (139)

Because XPX1,PX_{P}\setminus X_{1,P} is relatively compact, we obtain

XPX1,P1ΛTrF((h0)j)<.\int_{X_{P}\setminus X_{1,P}}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits F\bigl{(}(h_{0})_{\geq j}\bigr{)}<\infty. (140)

By (138), (139) and (140), we obtain

XD1ΛTrF((h0)j)=<0.\int_{X\setminus D}\sqrt{-1}\Lambda\mathop{\rm Tr}\nolimits F\bigl{(}(h_{0})_{\geq j}\bigr{)}=-\infty<0.

If a non-zero subbundle EE of 𝕂XD,r\mathbb{K}_{X\setminus D,r} satisfies θ(0XD,r)EEKXD\theta(0_{X\setminus D,r})E\subset E\otimes K_{X\setminus D}, there exists 1jr1\leq j\leq r such that E=i=jrK(r+12i)/2XDE=\bigoplus_{i=j}^{r}K^{(r+1-2i)/2}_{X\setminus D}. Therefore, (𝕂XD,r,θ(0XD,r),h0)(\mathbb{K}_{X\setminus D,r},\theta(0_{X\setminus D,r}),h_{0}) is analytically stable with respect to the GrG_{r}-action.

By Proposition 2.11, there exists hHarm(0XD,r)h\in\mathop{\rm Harm}\nolimits(0_{X\setminus D,r}) such that hh and h0h_{0} are mutually bounded. Then, hh satisfies the desired conditions. ∎

8.2. Existence of harmonic metrics in the potential theoretically hyperbolic case

8.2.1. Solvability of the Poisson equation and the existence of harmonic metrics

Let GG be a compact Lie group with a character κ:GS1\kappa:G\longrightarrow S^{1}. Let XX be an open Riemann surface equipped with a GG-action. Let gXg_{X} be any GG-invariant Kähler metric of XX. Let Λ\Lambda denote the adjoint of the multiplication of the Kähler form associated with gXg_{X}. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a (G,κ)(G,\kappa)-homogeneous Higgs bundle on XX. Let h0h_{0} be a GG-invariant Hermitian metric of EE.

A function on XX is called locally bounded if its restriction to any compact set is bounded.

Condition 8.8.

Assume that there exists a GG-invariant 0{\mathbb{R}}_{\geq 0}-valued locally bounded function α\alpha on XX such that

1Λ¯α|ΛF(h0)|h0\sqrt{-1}\Lambda\overline{\partial}\partial\alpha\geq\bigl{|}\Lambda F(h_{0})\bigr{|}_{h_{0}}

in the sense of distributions. Note that this condition is independent of the choice of gXg_{X}.

Proposition 8.9.

There exists a GG-invariant harmonic metric hh of the Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that

logTr(hh01)α+logrank(E),logTr(h0h1)α+logrank(E).\log\mathop{\rm Tr}\nolimits(h\cdot h_{0}^{-1})\leq\alpha+\log\mathop{\rm rank}\nolimits(E),\quad\log\mathop{\rm Tr}\nolimits(h_{0}\cdot h^{-1})\leq\alpha+\log\mathop{\rm rank}\nolimits(E). (141)

Proof   Let X1X2X_{1}\subset X_{2}\subset\cdots be a smooth exhaustive sequence of XX. Let hih_{i} be a harmonic metric of (E,¯E,θ)|Xi(E,\overline{\partial}_{E},\theta)_{|X_{i}} such that hi|Xi=h0|Xih_{i|\partial X_{i}}=h_{0|\partial X_{i}}. Let sis_{i} be the automorphism of E|XiE_{|X_{i}} determined by hi=h0|Xisih_{i}=h_{0|X_{i}}s_{i}. By (21), we obtain

1Λ¯logTr(si)(|F(h0)|h0)|Xi.\sqrt{-1}\Lambda\overline{\partial}\partial\log\mathop{\rm Tr}\nolimits(s_{i})\leq\Bigl{(}|F(h_{0})|_{h_{0}}\Bigr{)}_{|X_{i}}.

By the condition, we obtain the following on XiX_{i}:

1Λ¯(logTr(si)α)0.\sqrt{-1}\Lambda\overline{\partial}\partial\Bigl{(}\log\mathop{\rm Tr}\nolimits(s_{i})-\alpha\Bigr{)}\leq 0.

Note that si|Xi=idEi|Xis_{i|\partial X_{i}}=\mathop{\rm id}\nolimits_{E_{i}|\partial X_{i}}. Hence, we obtain

logTr(si)α|Xi+log(rankE)minXi(α)α|Xi+log(rankE).\log\mathop{\rm Tr}\nolimits(s_{i})\leq\alpha_{|X_{i}}+\log(\mathop{\rm rank}\nolimits E)-\min_{\partial X_{i}}(\alpha)\leq\alpha_{|X_{i}}+\log(\mathop{\rm rank}\nolimits E).

Similarly, we obtain

logTr(si1)α|Xi+log(rankE).\log\mathop{\rm Tr}\nolimits(s_{i}^{-1})\leq\alpha_{|X_{i}}+\log(\mathop{\rm rank}\nolimits E).

Then, we obtain the claim of the proposition. ∎

Corollary 8.10.

If det(h0)\det(h_{0}) is flat, there exists a GG-invariant harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) satisfying (141) and det(h)=det(h0)\det(h)=\det(h_{0}). Moreover, if α\alpha is bounded, and if X|F(h0)|<\int_{X}|F(h_{0})|<\infty, then (¯E+θ)(hh01)(\overline{\partial}_{E}+\theta)(h\cdot h_{0}^{-1}) is L2L^{2}. ∎

Remark 8.11.

In [34], the solvability of the Poisson equation is applied to the solvability of the Hermitian-Einstein equation. Proposition 8.9 is an analogue in the context of harmonic metrics on Riemann surfaces.

8.2.2. Potential theoretically hyperbolic Riemann surfaces

Let XX be a potential theoretically hyperbolic Riemann surface, i.e., there exists a non-constant non-positive subharmonic function on XX. It is well known that there exists a Green function G(p,q):X<0{\texttt{G}}(p,q):X\longrightarrow{\mathbb{R}}_{<0} characterized as follows (see [43, §7]):

  • G(p,){\texttt{G}}(p,\bullet) is harmonic on X{p}X\setminus\{p\}.

  • Let (Xp,zp)(X_{p},z_{p}) be a holomorphic coordinate neighbourhood such that zp(p)=0z_{p}(p)=0. Then, G(p,)log|zp|2{\texttt{G}}(p,\bullet)-\log|z_{p}|^{2} is harmonic on XpX_{p}.

  • If HH is any other subharmonic function satisfying the above two conditions, then G(p,)H{\texttt{G}}(p,\bullet)\geq H.

Recall that for any compact support 22-form ψ\psi on XX we obtain

1¯(12πG(p,q)ψ(q))=ψ.\sqrt{-1}\partial\overline{\partial}\Bigl{(}\frac{1}{2\pi}\int{\texttt{G}}(p,q)\psi(q)\Bigr{)}=\psi.
Lemma 8.12.

Let gg be any Kähler metric of XX. The volume form is denoted by dvol\mathop{\rm dvol}\nolimits. For a function v:X0v:X\longrightarrow{\mathbb{R}}_{\geq 0}, suppose that |G(p,q)(vdvol)(q)||{\texttt{G}}(p,q)(v\cdot\mathop{\rm dvol}\nolimits)(q)| is integrable on XX for any pp. Then, by setting α(v)(p):=12πG(p,q)(vdvol)(q)\alpha(v)(p):=-\frac{1}{2\pi}\int{\texttt{G}}(p,q)(v\mathop{\rm dvol}\nolimits)(q), we obtain α(v)0\alpha(v)\geq 0 and 1Λ¯α(v)=v\sqrt{-1}\Lambda\overline{\partial}\partial\alpha(v)=v. ∎

The following is clear by the construction of the Green function (see [43, §7]).

Lemma 8.13.

Let KK be any compact subset of XX. Let NN be any relatively compact neighbourhood of KK in XX. Then, G(p,q){\texttt{G}}(p,q) is bounded on K×(XN)K\times(X\setminus N). ∎

Corollary 8.14.

Let ψ\psi be a compact support 22-form on XX. Then, there exists a bounded function α\alpha on XX such that ¯α=ψ\partial\overline{\partial}\alpha=\psi. ∎

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a (G,κ)(G,\kappa)-homogeneous Higgs bundle on XX. Let h0h_{0} be a GG-invariant Hermitian metric of EE. Suppose that

G(p,q)(|F(h0)|h0)dvol(q)<\int{\texttt{G}}(p,q)\bigl{(}|F(h_{0})|_{h_{0}}\bigr{)}\mathop{\rm dvol}\nolimits(q)<\infty

for any pp.

Proposition 8.15.

There exists a GG-invariant harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that

max{log|hh01|h0,log|h0h1|h0}α(|F(h0)|h0)+logrank(E).\max\Bigl{\{}\log\bigl{|}h\cdot h_{0}^{-1}\bigr{|}_{h_{0}},\,\,\log\bigl{|}h_{0}\cdot h^{-1}\bigr{|}_{h_{0}}\Bigr{\}}\leq\alpha(|F(h_{0})|_{h_{0}})+\log\mathop{\rm rank}\nolimits(E).

If det(h0)\det(h_{0}) is flat, hh also satisfies det(h)=det(h0)\det(h)=\det(h_{0}). ∎

Corollary 8.16.

Suppose that the support of F(h0)F(h_{0}) is compact. Then, there exists a GG-invariant harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that hh and h0h_{0} are mutually bounded. If det(h0)\det(h_{0}) is flat, it also satisfies (i) det(h)=det(h0)\det(h)=\det(h_{0}), (ii) (¯E+θ)(hh01)(\overline{\partial}_{E}+\theta)(h\cdot h_{0}^{-1}) is L2L^{2}. ∎

8.2.3. Upper half plane

Let us state a consequence in the case X=={z|Im(z)>0}X=\mathbb{H}=\{z\in{\mathbb{C}}\,|\,\mathop{\rm Im}\nolimits(z)>0\}. Let (x,y)(x,y) be the real coordinate system determined by z=x+1yz=x+\sqrt{-1}y.

Proposition 8.17.

Let h0h_{0} be a GG-invariant Hermitian metric of EE such that |F(h0)|=O((1+y)2)|F(h_{0})|=O((1+y)^{-2}). Then, there exists a GG-invariant harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that

max{log|hh01|h0,log|h0h1|h0}=O(log(2+y)).\max\Bigl{\{}\log\bigl{|}h\cdot h_{0}^{-1}\bigr{|}_{h_{0}},\,\log\bigl{|}h_{0}\cdot h^{-1}\bigr{|}_{h_{0}}\Bigr{\}}=O\Bigl{(}\log(2+y)\Bigr{)}.

If det(h0)\det(h_{0}) is flat, we obtain det(h)=det(h0)\det(h)=\det(h_{0}). If |F(h0)|=O((1+y)2ϵ)|F(h_{0})|=O((1+y)^{-2-\epsilon}) for a positive constant ϵ\epsilon, then hh and h0h_{0} are mutually bounded.

Proof   For a function ff on \mathbb{H} such that f=O((1+y)2)f=O\bigl{(}(1+y)^{-2}\bigr{)}, we set

u(x,y)=14πlog((xξ)2+(yη)2(xξ)2+(y+η)2)f(ξ,η)dξdη.u(x,y)=-\frac{1}{4\pi}\int_{\mathbb{H}}\log\Bigl{(}\frac{(x-\xi)^{2}+(y-\eta)^{2}}{(x-\xi)^{2}+(y+\eta)^{2}}\Bigr{)}f(\xi,\eta)\,d\xi\,d\eta.

Then, (x2+y2)u=f-(\partial_{x}^{2}+\partial_{y}^{2})u=f. If ff is positive, then uu is also positive on \mathbb{H}. The rest follows from Lemma 4.5 and Proposition 8.9. ∎

8.3. Uniqueness of harmonic metrics in the potential theoretically parabolic case

Let XX be a potential theoretically parabolic Riemann surface, i.e., XX is non-compact, and any non-positive subharmonic function on XX is constant. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on XX.

Proposition 8.18.

Let hih_{i} (i=1,2)(i=1,2) be harmonic metrics of the Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) which are mutually bounded. Then, there exists a decomposition

(E,¯E,θ)=j=1m(Ej,¯Ej,θj)(E,\overline{\partial}_{E},\theta)=\bigoplus_{j=1}^{m}(E_{j},\overline{\partial}_{E_{j}},\theta_{j})

such that (i) the decomposition is orthogonal with respect to both h1h_{1} and h2h_{2}, (ii) there exist positive numbers aja_{j} such that h1|Ej=ajh2|Ejh_{1|E_{j}}=a_{j}h_{2|E_{j}}.

Proof   Let ss be the automorphism of EE determined by h2=h1sh_{2}=h_{1}s. Then, logTr(s)\log\mathop{\rm Tr}\nolimits(s) is a bounded subharmonic function on XX. Because XX is assumed to be potential theoretically parabolic, logTr(s)\log\mathop{\rm Tr}\nolimits(s) is constant. According to (19), it implies that (¯E+θ)s=0(\overline{\partial}_{E}+\theta)s=0, and hence we obtain the desired decomposition. ∎

The following corollary of the proposition has been well known after [38, 39].

Corollary 8.19.

Let XX be the complement of a finite subset DD of a compact Riemann surface X¯\overline{X}. Then, the claim of Proposition 8.18 holds.

Proof   Let ff be an 0{\mathbb{R}}_{\leq 0}-valued subharmonic function on XX. According to [36, Theorem 3.3.25], it uniquely extends to a subharmonic function on X¯\overline{X}. By the maximum principle, we obtain that ff is constant, i.e., XX is potential theoretically parabolic. ∎

Remark 8.20.

It is also easy and standard to prove the claim of the corollary directly. Indeed, in the proof of Proposition 8.18, without any assumption on XX, we obtain the boundedness of logTr(s)\log\mathop{\rm Tr}\nolimits(s) and 1Λ¯logTr(s)0\sqrt{-1}\Lambda\overline{\partial}\partial\log\mathop{\rm Tr}\nolimits(s)\leq 0 on XX. If X=X¯DX=\overline{X}\setminus D, 1Λ¯logTr(s)0\sqrt{-1}\Lambda\overline{\partial}\partial\log\mathop{\rm Tr}\nolimits(s)\leq 0 holds on X¯\overline{X} according to [39, Lemma 2.2]. Hence, logTr(s)\log\mathop{\rm Tr}\nolimits(s) is constant, and we obtain (¯E+θ)s=0(\overline{\partial}_{E}+\theta)s=0.

References

  • [1] D. Baraglia, Cyclic Higgs bundles and the affine Toda equations, Geom. Dedicata 174 (2015), 25–42.
  • [2] Y. Benoist and D. Hulin, Cubic differentials and finite volume convex projective surfaces, Geom. Topol. 17 (2013), 595–620.
  • [3] Y. Benoist and D. Hulin, Cubic differentials and hyperbolic convex sets, J. Differential Geom. 98 (2014), 1–19.
  • [4] O. Biquard and P. Boalch, Wild nonabelian Hodge theory on curves, Compositio Math. 140 (2004) 179–204.
  • [5] J. Bolton, F. Pedit and L. Woodward, Minimal surfaces and the affine Toda field model, J. Reine Angew. Math. 459 (1995), 119–150.
  • [6] F. E. Burstall and F. Pedit, Harmonic maps via Adler-Kostant-Symes theory, Harmonic Maps and Integrable Systems, eds. A. P. Fordy and J. C. Wood, Aspects of Math. E23, Vieweg, 1994, 221–272.
  • [7] S. Cecotti and C. Vafa, Topological—anti-topological fusion, Nuclear Phys. B 367 (1991), 359–461.
  • [8] B. Collier, N. Tholozan and J. Toulisse, The geometry of maximal representations of surface groups in SO(2,n)SO(2,n), Duke Math. J., 168, 2873–2949, 2019.
  • [9] S. K. Donaldson, Boundary value problems for Yang-Mills fields, J. Geom. Phys. 8 (1992), 89–122.
  • [10] D. Dumas and M. Wolf, Polynomial cubic differentials and convex polygons in the projective plane, Geom. Funct. Anal. 25 (2015), 1734–1798.
  • [11] P. Eberlein, Lattices in spaces of nonpositive curvature, Ann. of Math. (2) 111 (1980), 435–476.
  • [12] O. Forster, Lectures on Riemann surfaces, Graduate Texts in Mathematics, 81. Springer-Verlag, New York-Berlin, 1981.
  • [13] P. Griffiths, J. Harris, Principles of algebraic geometry, Wiley-Interscience [John Wiley & Sons], New York, 1978.
  • [14] M. A. Guest, A. Its and C. S. Lin, Isomonodromy aspects of the tttt^{*} equations of Cecotti and Vafa I. Stokes data, Int. Math. Res. Notices 2015 (2015), 11745–11784.
  • [15] M. A. Guest and C. S. Lin, Nonlinear PDE aspects of the tttt^{*} equations of Cecotti and Vafa, J. reine angew. Math. 689 (2014) 1–32.
  • [16] R. C. Gunning and R. Narasimhan, Immersion of open Riemann surfaces, Math. Ann. 174 (1967), 103–108.
  • [17] S. Gupta, Harmonic maps and wild Teichmüller spaces, J. Topol. Anal.13 (2021), 349–393.
  • [18] Z. Han, L. T. Tam, A. Treibergs, and T. A. Wan, Harmonic maps from the complex plane into surfaces with nonpositive curvature, Comm. Anal. Geom. 3 (1995), 85–114.
  • [19] N. J. Hitchin, The self-duality equations on a Riemann surface, Proc. London Math. Soc. (3) 55 (1987), 59–126.
  • [20] F. Labourie, Flat projective structures on surfaces and cubic holomorphic differentials, Pure Appl. Math. Q. 3 (2007), 1057–1099.
  • [21] R. E. Langer, On the zeros of exponential sums and integrals, Bull. Amer. Math. Soc. 37 (1931), 213–239.
  • [22] Y. B. Levin, Yu. Lyubarskii, M. Sodin and V. Tkachenko, Lectures on entire functions, Translations of Mathematical Monographs, 150, American Mathematical Society, Providence, RI, 1996.
  • [23] Q. Li, On the uniqueness of vortex equations and its geometric applications, J. Geom. Anal. 29 (2019), 105–120.
  • [24] Q. Li and T. Mochizuki, Complete solutions of Toda equations and cyclic Higgs bundles over non-compact surfaces, arXiv:2010.05401, 2020.
  • [25] J. Loftin, Affine spheres and convex RPnRP^{n}-manifolds, Amer. J. Math. 123 (2001), 255–274.
  • [26] J. Loftin, The compactification of the moduli space of convex 2\mathbb{R}\mathbb{P}^{2} surfaces. I, J. Differential Geom. 68 (2004), 223–276.
  • [27] J. Loftin, Convex 2\mathbb{R}\mathbb{P}^{2} structures and cubic differentials under neck separation, J. Differential Geom. 113 (2019), 315–383.
  • [28] H. Majima, Asymptotic analysis for integrable connections with irregular singular points, Lecture Notes in Mathematics, 1075, Springer-Verlag, Berlin, 1984.
  • [29] T. Mochizuki, Wild harmonic bundles and wild pure twistor DD-modules, Astérisque 340, Société Mathématique de France, Paris, 2011.
  • [30] T. Mochizuki, Harmonic bundles and Toda lattices with opposite sign I, to appear in RIMS Kôkyûroku Bessatsu (essentially the part I of arXiv:1301.1718)
  • [31] T. Mochizuki, Harmonic bundles and Toda lattices with opposite sign II, Comm. Math. Phys. 328 (2014), 1159–1198.
  • [32] T. Mochizuki, Asymptotic behaviour of certain families of harmonic bundles on Riemann surfaces, J. Topol. 9 (2016), 1021–1073.
  • [33] T. Mochizuki, Kobayashi-Hitchin correspondence for analytically stable bundles. Trans. Amer. Math. Soc. 373 (2020), 551–596.
  • [34] L. Ni, The Poisson equation and Hermitian-Einstein metrics on holomorphic vector bundles over complete noncompact Kähler manifolds, Indiana Univ. Math. J. 51 (2002), 679–704.
  • [35] X. Nie, Poles of cubic differentials and ends of convex RP2RP^{2} surfaces, J. Differential Geom. 123 (2023), 67–140.
  • [36] J. Noguchi and T. Ochiai, Geometric function theory in several complex variables, Translated from the Japanese by Noguchi. Translations of Mathematical Monographs, 80. American Mathematical Society, Providence, RI, 1990.
  • [37] R. Schoen and S. T. Yau, Existence of incompressible minimal surfaces and the topology of three-dimensional manifolds with nonnegative scalar curvature, Ann. of Math. (2) 110 (1979), 127–142.
  • [38] C. Simpson, Constructing variations of Hodge structure using Yang-Mills theory and application to uniformization, J. Amer. Math. Soc. 1 (1988), 867–918.
  • [39] C. Simpson, Harmonic bundles on non-compact curves, J. Amer. Math. Soc. 3 (1990), 713–770.
  • [40] Y. T. Siu and S. T. Yau, Compactification of negatively curved complete Kähler manifolds of finite volume, Seminar on Differential Geometry, pp. 363–380, Ann. of Math. Stud., 102, Princeton Univ. Press, Princeton, N.J., 1982.
  • [41] J. Tamarkin, Some general problems of the theory of ordinary linear differential equations and expansion of an arbitrary function in series of fundamental functions, Math. Z. 27 (1928), 1–54.
  • [42] A. Tamburelli and M. Wolf, Planar minimal surfaces with polynomial growth in the Sp(4,)Sp(4,\mathbb{R})-symmetric space, arXiv:2002.07295, 2020.
  • [43] D. Varolin, Riemann surfaces by way of complex analytic geometry, Graduate Studies in Mathematics, 125. American Mathematical Society, Providence, RI, 2011. xviii+236.
  • [44] T. Y.-H. Wan, Constant mean curvature surface, harmonic maps, and universal Teichmüller space, J. Differential Geom. 35 (1992), 643–657.
  • [45] T. Y.-H. Wan and T. K.-H. Au, Parabolic constant mean curvature spacelike surfaces, Proc. Amer. Math. Soc. 120 (1994), 559–564.
  • [46] C. P. Wang, Some examples of complete hyperbolic affine 2-spheres in 3\mathbb{R}^{3}, Global differential geometry and global analysis (Berlin, 1990), Lecture Notes in Math.,1481, Springer, Berlin, 1991, 271–280.
  • [47] W. Wasow, Asymptotic expansions for ordinary differential equations, Reprint of the 1976 edition. Dover Publications, Inc., New York, 1987.
  • [48] C. E. Wilder, Expansion problems of ordinary linear differential equations with auxiliary conditions at more than two points, Trans. Amer. Math. Soc. 18 (1917), 415–442.
  • [49] M. Wolf, The Teichmüller theory of harmonic maps, J. Differential Geom. 29 (1989) 449–479.
  • [50] M. Wolf, Infinite energy harmonic maps and degeneration of hyperbolic surfaces in moduli space, J. Differential Geom. 33 (1991), 487–539.