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

Harmonic metrics of generically regular semisimple Higgs bundles on non-compact Riemann surfaces

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

We prove that a generically regular semisimple Higgs bundle equipped with a non-degenerate symmetric pairing on any Riemann surface always has a harmonic metric compatible with the pairing. We also study the classification of such compatible harmonic metrics in the case where the Riemann surface is the complement of a finite set DD in a compact Riemann surface. In particular, we prove the uniqueness of a compatible harmonic metric if the Higgs bundle is wild and regular semisimple at each point of DD.

MSC: 53C07, 58E15, 14D21, 81T13.
Keywords: harmonic bundle, non-degenerate symmetric product, real structure

1 Introduction

1.1 Harmonic bundles

Let XX be a Riemann surface. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on XX. Let hh be a Hermitian metric of EE. We obtain the Chern connection h=¯E+E,h\nabla_{h}=\overline{\partial}_{E}+\partial_{E,h} and the adjoint θh\theta^{\dagger}_{h} of θ\theta. The metric hh is called a harmonic metric of the Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) if h+θ+θh\nabla_{h}+\theta+\theta^{\dagger}_{h} is flat, i.e., hh+[θ,θh]=0\nabla_{h}\circ\nabla_{h}+[\theta,\theta^{\dagger}_{h}]=0, and (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) is called a harmonic bundle. It was introduced by Hitchin [5], and it has been one of the most important and interesting mathematical objects.

A starting point is the study of the existence and the classification of harmonic metrics. If XX is compact, the following definitive theorem is most fundamental, which is due to Hitchin [5] and Simpson [17].

Theorem 1.1 ([5, 17])

If XX is compact, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) has a harmonic metric if and only if (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is polystable of degree 0. If h1h_{1} and h2h_{2} are two harmonic metrics of (E,¯E,θ)(E,\overline{\partial}_{E},\theta), there exists a decomposition (E,¯E,θ)=i=1m(Ei,¯Ei,θi)(E,\overline{\partial}_{E},\theta)=\bigoplus_{i=1}^{m}(E_{i},\overline{\partial}_{E_{i}},\theta_{i}) such that (i) the decomposition is orthogonal with respect to both h1h_{1} and h2h_{2}, (ii) h2|Ei=aih1|Eih_{2|E_{i}}=a_{i}h_{1|E_{i}} for some ai>0a_{i}>0.  

The study in the non-compact case was pioneered by Simpson [17, 18], and pursued by Biquard-Boalch [2] and the second author [14]. Let XX be the complement of a finite subset DD in a compact Riemann surface X¯\overline{X}. A Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) on XX induces a coherent sheaf on the cotangent bundle TXT^{\ast}X whose support ΣE,θ\Sigma_{E,\theta} is called the spectral curve of the Higgs bundle. The natural morphism ΣE,θX\Sigma_{E,\theta}\to X is finite and flat. The Higgs bundle is called tame if the closure of ΣE,θ\Sigma_{E,\theta} in TX¯(logD)T^{\ast}\overline{X}(\log D) is proper over X¯\overline{X}. The Higgs bundle is called wild if the closure of ΣE,θ\Sigma_{E,\theta} in the projective completion of TX¯T^{\ast}\overline{X} is complex analytic. If a tame (resp. wild) Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is equipped with a harmonic metric hh, (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) is called a tame (resp. wild) harmonic bundle. From a tame (resp. wild) harmonic bundle (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h), we obtain a regular (resp. good) filtered Higgs bundle (𝒫h(E),θ)(\mathcal{P}^{h}_{\ast}(E),\theta) on (X¯,D)(\overline{X},D). (See §4.2.2 for the notions of regular and good filtered Higgs bundles. See [14, §2.5] for the notation 𝒫h(E)\mathcal{P}^{h}_{\ast}(E).) The following theorem was proved by Simpson [18] in the tame case and generalized to the wild case in [2, 14].

Theorem 1.2

For a wild harmonic bundle (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) on (X¯,D)(\overline{X},D), (𝒫h(E),θ)(\mathcal{P}^{h}_{\ast}(E),\theta) is polystable of degree 0. Conversely, for any polystable good filtered Higgs bundle (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) of degree 0 on (X,D)(X,D) such that (𝒱,θ)|XD=(E,¯E,θ)(\mathcal{V},\theta)_{|X\setminus D}=(E,\overline{\partial}_{E},\theta), there exists a harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that 𝒫h(E)=𝒫𝒱\mathcal{P}^{h}_{\ast}(E)=\mathcal{P}_{\ast}\mathcal{V}. Moreover, if hih_{i} (i=1,2)(i=1,2) be two harmonic metrics of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that 𝒫h1(E)=𝒫h2(E)\mathcal{P}^{h_{1}}_{\ast}(E)=\mathcal{P}^{h_{2}}_{\ast}(E), then there exists a decomposition (E,¯E,θ)=i=1m(Ei,¯Ei,θi)(E,\overline{\partial}_{E},\theta)=\bigoplus_{i=1}^{m}(E_{i},\overline{\partial}_{E_{i}},\theta_{i}) as in Theorem 1.1.  

More recently, in [9, 10], we studied a different type of existence and classification results. Let XX be any Riemann surface, which is not necessarily the complement of a finite subset in a compact Riemann surface. Let KXK_{X} denote the canonical bundle of XX. Let r2r\geq 2 be an integer. We set 𝕂X,r:=i=1rKX(r+12i)/2\mathbb{K}_{X,r}:=\bigoplus_{i=1}^{r}K_{X}^{(r+1-2i)/2}. For any rr-differential qrq_{r} on XX, we obtain the cyclic Higgs field θ(qr)\theta(q_{r}) of 𝕂X,r\mathbb{K}_{X,r} from qr:KX(r1)/2KX(r1)/2KXq_{r}:K_{X}^{-(r-1)/2}\to K_{X}^{(r-1)/2}\otimes K_{X} and the identity morphisms KX(r+12i)/2KX(r+12(i+1)/2)KXK_{X}^{(r+1-2i)/2}\simeq K_{X}^{(r+1-2(i+1)/2)}\otimes K_{X}.

Theorem 1.3 ([9])

If qr0q_{r}\neq 0, (𝕂X,r,θ(qr))(\mathbb{K}_{X,r},\theta(q_{r})) has a harmonic metric. More precisely, there exists a unique harmonic metric hch^{c} such that (i) det(hc)=1\det(h^{c})=1, (ii) 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, (iii) the metrics h|KX(r+12i)/2c(h|KX(r+12(i1))/2c)1h^{c}_{|K_{X}^{(r+1-2i)/2}}\otimes(h^{c}_{|K_{X}^{(r+1-2(i-1))/2}})^{-1} of TX=KX1TX=K_{X}^{-1} are complete.  

Note that in general there are many other harmonic metrics hh of (𝕂X,r,θ(qr))(\mathbb{K}_{X,r},\theta(q_{r})) satisfying the conditions (i) and (ii). In [10], we studied a classification of such harmonic metrics under the additional assumption that XX is the complement of a finite subset DD in a compact Riemann surface X¯\overline{X}, and that qrq_{r} has at worst multi-growth orders at each point of DD. See the introductions of [9, 10] for a brief review of previous studies on this type of harmonic bundles and related subjects.

After [9, 10], it seems reasonable to study whether a Higgs bundle with an appropriate additional symmetry has a harmonic metric compatible with the symmetry. Note that such a harmonic metric is interesting in relation with the theory of minimal surfaces. If the base space is compact, it should be a consequence of the existence and uniqueness theorem due to Hitchin and Simpson (see Theorem 1.1). Namely, we know the existence of a harmonic metric of a polystable Higgs bundle of degree 0, and the uniqueness should imply the compatibility of the harmonic metric with the symmetry. (For example, see §2.3.2.) If the base space is the complement of a finite subset in a compact Riemann surface, and if the Higgs bundle is wild, we may still apply a similar argument to obtain the classification of harmonic metrics. However, in a more general situation, we do not know a useful general result for neither existence nor uniqueness. We are motivated to find many examples. Even in the wild case, it would be helpful if we could simplify the assumption on Higgs bundles. We also note that the analysis of the non-compact case would be useful for the compact case. In this paper, we shall study Higgs bundles equipped with a non-degenerate symmetric pairing under the assumption that the Higgs field is generically regular semisimple.

1.2 Higgs bundles with non-degenerate symmetric product

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on any Riemann surface XX. A non-degenerate symmetric pairing of the Higgs bundle is a holomorphic symmetric bilinear form C:EE𝒪XC:E\otimes E\to\mathcal{O}_{X} such that C(θid)=C(idθ)C(\theta\otimes\mathop{\rm id}\nolimits)=C(\mathop{\rm id}\nolimits\otimes\theta). When (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is equipped with a non-degenerate symmetric pairing CC, we say that a harmonic metric hh is compatible with CC if the induced morphism ΨC:EE\Psi_{C}:E\to E^{\lor} is isometric with respect to hh and hh^{\lor}. In that case, the harmonic map (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is induced by a GL(r,)\mathop{\rm GL}\nolimits(r,{\mathbb{R}})-harmonic bundle, where r=rankEr=\mathop{\rm rank}\nolimits E.

We shall also impose the following generic regular semisimplicity condition.

  • There exists QXQ\in X such that the fiber of ΣE,θX\Sigma_{E,\theta}\to X over QQ consists of exactly rr points.

Remark 1.4

From the definition, it is clear that a Higgs bundle (E,θ)(E,\theta) is generically regular semisimple if one of the following holds: (i) tr(θr)0\mathop{\rm tr}\nolimits(\theta^{r})\neq 0 and tr(θk)=0\mathop{\rm tr}\nolimits(\theta^{k})=0 (k=1,,r1)(k=1,\ldots,r-1); (ii) tr(θr1)0\mathop{\rm tr}\nolimits(\theta^{r-1})\neq 0 and tr(θk)=0\mathop{\rm tr}\nolimits(\theta^{k})=0 (k=1,,r2,r)(k=1,\ldots,r-2,r), where r=rankEr=\mathop{\rm rank}\nolimits E. We also remark that most Higgs bundles are generically regular semisimple. For example, a Higgs bundle (E,θ)(E,\theta) is generically regular semisimple if one of the following holds: (i) the characteristic polynomial of (E,θ)(E,\theta) is irreducible; (ii) EE has no proper subbundle preserved by θ\theta. See §2.4 (Proposition 2.42 and Proposition 2.43) for more details.  

We shall prove the following general theorem.

Theorem 1.5 (Theorem 2.34)

Suppose that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple and equipped with a non-degenerate symmetric pairing CC. Then, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) has a harmonic metric compatible with CC.

In the compact case, it follows from Theorem 1.1, but note that we do not have to assume that the Higgs bundle is polystable of degree 0. For the proof of Theorem 1.5 in the non-compact case, we use a method developed in [10] together with a key estimate for harmonic metrics compatible with a non-degenerate symmetric product (Proposition 2.37). It follows from Simpson’s main estimate ([18, 19]) on the norm of the Higgs field with the aid of a linear algebraic argument (Proposition 2.15).

Remark 1.6

We can obtain another a priori estimate for harmonic metrics compatible with a non-degenerate symmetric pairing from a variant of Simpson’s main estimate [18, 13] and a linear algebraic argument. It is useful even in the study of Higgs bundles on compact Riemann surfaces which are not necessarily equipped with a global non-degenerate symmetric pairing. In [16], the estimate is applied to study large-scale solutions of Hitchin equations. Non-degenerate symmetric pairings are also useful in the study of Hitchin metric of the moduli space of Higgs bundles [15].  

In particular, suppose (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is cyclic, that is, (E,¯E)(E,\overline{\partial}_{E}) is a direct sum of holomorphic line bundles LkL_{k} (k=1,,r)(k=1,\ldots,r) and θ\theta takes LkL_{k} to Lk+1KXL_{k+1}\otimes K_{X} (k=1,,r)(k=1,\ldots,r) where Lr+1:=L1L_{r+1}:=L_{1}. Then, if det(θ)0\det(\theta)\neq 0, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is obviously generically regular semisimple. We obtain the following corollary.

Corollary 1.7

Suppose (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is a cyclic Higgs bundle on XX satisfying detθ0\det\theta\neq 0 and equipped with a non-degenerate symmetric pairing CC. Then, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) has a harmonic metric compatible with CC.

If XX is compact, the assumptions imply that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is polystable of degree 0, and we can easily deduce Theorem 1.5 from Theorem 1.1. (See §2.3.2.) We can also obtain the uniqueness of such compatible harmonic metrics in the compact case. In the non-compact case, the uniqueness does not hold, in general.

We also study more detailed classification in the case where XX is the complement of a finite subset DD in a compact Riemann surface X¯\overline{X}. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a generically regular semisimple Higgs bundle on XX which is wild at each point of DD. Suppose that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is equipped with a non-degenerate symmetric pairing CC.

Theorem 1.8 (Theorem 4.15)

If hh is a harmonic metric of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC, then the induced symmetric pairing CC of 𝒫h(E)\mathcal{P}^{h}_{\ast}(E) is perfect. Conversely, for any good filtered Higgs bundle (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) on (X¯,D)(\overline{X},D) equipped with a perfect pairing C𝒱C_{\mathcal{V}} such that (𝒱,θ,C𝒱)|X=(E,¯E,θ,C)(\mathcal{V},\theta,C_{\mathcal{V}})_{|X}=(E,\overline{\partial}_{E},\theta,C), there exists a unique harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC satisfying the boundary condition 𝒫h(E)=𝒫𝒱\mathcal{P}^{h}_{\ast}(E)=\mathcal{P}_{\ast}\mathcal{V}.

We remark that in Theorem 1.8, we do not have to assume that (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) is polystable of degree 0 because it follows from the existence of a perfect pairing.

Let us mention a condition under which we can prove the uniqueness of a compatible harmonic metric without the boundary condition. For each PDP\in D, let (UP,z)(U_{P},z) be a holomorphic coordinate neighbourhood around PP such that z(P)=0z(P)=0. We can regard UPU_{P} as an open subset of {\mathbb{C}}. For e>0e\in{\mathbb{Z}}_{>0}, let φe:\varphi_{e}:{\mathbb{C}}\to{\mathbb{C}} be the map defined by φe(ζ)=ζe\varphi_{e}(\zeta)=\zeta^{e}. We set UP(e):=φe1(UP)U^{(e)}_{P}:=\varphi_{e}^{-1}(U_{P}). We obtain the endomorphism fP(e)f^{(e)}_{P} of φe(E)|UP(e){0}\varphi_{e}^{\ast}(E)_{|U^{(e)}_{P}\setminus\{0\}} defined by fP(e)dζ/ζ=φe(θ)f^{(e)}_{P}\,d\zeta/\zeta=\varphi_{e}^{\ast}(\theta). If UU is sufficiently small, and if e=r!e=r!, there exist holomorphic functions α1,,αr\alpha_{1},\ldots,\alpha_{r} on U(e){0}U^{(e)}\setminus\{0\} and the eigen decomposition

φe(E)=i=1rEi\varphi_{e}^{\ast}(E)=\bigoplus_{i=1}^{r}E_{i}

such that (i) αiαj\alpha_{i}-\alpha_{j} (ij)(i\neq j) are nowhere vanishing on U(e){0}U^{(e)}\setminus\{0\}, (ii) f(e)αiidEif^{(e)}-\alpha_{i}\mathop{\rm id}\nolimits_{E_{i}} is 0 on EαiE_{\alpha_{i}}, (iii) rankEi=1\mathop{\rm rank}\nolimits E_{i}=1. We say that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at PP if (αiαj)1(\alpha_{i}-\alpha_{j})^{-1} (ij)(i\neq j) are holomorphic at 0.

Theorem 1.9 (Theorem 4.16)

Suppose (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple on XX which is wild at each point of DD and equipped with a non-degenerate symmetric pairing CC. If (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at each point of DD, there exists a unique harmonic metric on (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC.  

1.3 Hitchin section

Our main examples are Higgs bundles contained in the Hitchin section [7], which we recall by following [8]. Let qjq_{j} (j=2,,r)(j=2,\ldots,r) be holomorphic jj-differentials on XX. The multiplication of qjq_{j} induces the following morphisms:

KX(r2i+1)/2KX(r2i+2(j1)+1)/2KX(jir).K_{X}^{(r-2i+1)/2}\to K_{X}^{(r-2i+2(j-1)+1)/2}\otimes K_{X}\quad(j\leq i\leq r).

We also have the multiplications of i(ri)/2i(r-i)/2 for i=1,,r1i=1,\ldots,r-1:

KX(r2i+1)/2KX(r2(i+1)+1)/2KX.K_{X}^{(r-2i+1)/2}\to K_{X}^{(r-2(i+1)+1)/2}\otimes K_{X}.

They define a Higgs field θ(𝒒)\theta(\boldsymbol{q}) of 𝕂X,r\mathbb{K}_{X,r}. The natural pairings KX(r2i+1)/2KX(r2i+1)/2𝒪XK_{X}^{(r-2i+1)/2}\otimes K_{X}^{-(r-2i+1)/2}\to\mathcal{O}_{X} induce a non-degenerate symmetric bilinear form C𝕂,X,rC_{\mathbb{K},X,r} of 𝕂X,r\mathbb{K}_{X,r}. It is a non-degenerate symmetric pairing of (𝕂X,r,θ(𝒒))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})). The following result is a partial affirmative answer to [20, Question A].

Corollary 1.10 (Corollary 4.18)

If the Higgs bundle (𝕂X,r,θ(𝐪))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})) is generically regular semisimple, then there exists a harmonic metric of (𝕂X,r,θ(𝐪))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})) which is compatible with C𝕂,X,rC_{\mathbb{K},X,r}. It is induced by an SL(r,)\mathop{\rm SL}\nolimits(r,{\mathbb{R}})-harmonic bundle.  

In the other extreme case qi=0q_{i}=0 (i=2,,r)(i=2,\ldots,r), (𝕂X,r,θ(𝒒))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})) has a harmonic metric if and only if XX is hyperbolic. For example, (𝕂,r,θ(0,,0))(\mathbb{K}_{{\mathbb{C}},r},\theta(0,\ldots,0)) does not have a harmonic metric. (See [9, Lemma 3.13].)

See §4.3.1–§4.3.3 for more concrete examples. In §4.3.1, we explain an example for which Theorem 1.9 ensures the uniqueness of harmonic metrics. In §4.3.2, we explain an example for which the uniqueness does not hold.

Remark 1.11

In [11], by using different techniques, we shall establish that the Higgs bundle (𝕂X,r,θ(𝐪))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})) has a harmonic metric for any 𝐪\boldsymbol{q} if XX is hyperbolic. We shall also study the existence of harmonic metrics of Higgs bundles in the Collier section and Gothen section by using the techniques in both [11] and this paper.  

1.4 Higher dimensional case

Though we do not study the higher dimensional case in this paper, some part of the theory can be easily generalized. We give only a brief sketch of the statements without proof. More details will be explained elsewhere.

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle of rank rr on XX. It naturally induces a coherent 𝒪TX\mathcal{O}_{T^{\ast}X}-module whose support ΣE,θ\Sigma_{E,\theta} is called the spectral variety. The Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is called generically regular semisimple if there exists QXQ\in X such that the fiber of ΣE,θX\Sigma_{E,\theta}\to X over QQ consists of exactly rr points. A non-degenerate symmetric pairing CC of EE is called a non-degenerate symmetric pairing of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) if θ\theta is self-adjoint with respect to CC. Note that the existence of CC implies that c1(E)=0c_{1}(E)=0 in H2(X,)H^{2}(X,{\mathbb{Q}}).

Let us consider a generically regular semisimple Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) on XX equipped with a non-degenerate symmetric pairing CC. If XX is compact Kähler, we can prove that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is polystable. By the theory of Simpson [17], we can prove that there exists a unique Hermitian-Einstein metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) which is compatible with CC. If moreover ch2(E)=0\mathop{\rm ch}\nolimits_{2}(E)=0, hh is pluri-harmonic. In this sense, Theorem 1.5 can be generalized if XX is compact Kähler.

Suppose that there exists a complex projective manifold X¯\overline{X} with a simple normal crossing hypersurface DD such that X=X¯DX=\overline{X}\setminus D. If (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is good wild along DD and equipped with a pluri-harmonic metric compatible with CC, then CC induces a perfect pairing of the associated good filtered Higgs bundle (𝒫h(E),θ)(\mathcal{P}^{h}_{\ast}(E),\theta). Conversely, suppose that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) and CC extends to a good filtered Higgs bundle (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) with a perfect pairing C𝒱C_{\mathcal{V}} such that ch2(𝒫𝒱)=0\mathop{\rm ch}\nolimits_{2}(\mathcal{P}_{\ast}\mathcal{V})=0, then there exists a unique pluri-harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC such that 𝒫hE=𝒫𝒱\mathcal{P}^{h}_{\ast}E=\mathcal{P}_{\ast}\mathcal{V}. In this sense, Theorem 1.8 can be generalized.

We can generalize the regular semisimplicity condition for a good wild Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) on (X,D)(X,D) at PDP\in D. If it is satisfied at any point of DD, there exists a unique good filtered extension (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) such that CC induces a perfect pairing of (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta). If ch2(𝒫𝒱)=0\mathop{\rm ch}\nolimits_{2}(\mathcal{P}_{\ast}\mathcal{V})=0, then (E,¯E,θ)(E,\overline{\partial}_{E},\theta) has a unique pluri-harmonic metric compatible with CC. If ch2(𝒫𝒱)0\mathop{\rm ch}\nolimits_{2}(\mathcal{P}_{\ast}\mathcal{V})\neq 0, then (E,¯E,θ)(E,\overline{\partial}_{E},\theta) does not have a pluri-harmonic metric compatible with CC. In this sense, Theorem 1.9 can be generalized.

2 Harmonic bundles with real structure

2.1 Real structures and symmetric pairings of vector spaces

2.1.1 Real structures of vector spaces

For complex vector spaces ViV_{i} (i=1,2)(i=1,2), an {\mathbb{R}}-linear morphism f:V1V2f:V_{1}\to V_{2} is called sesqui-linear if f(αv)=α¯f(v)f(\alpha v)=\overline{\alpha}f(v) for any α\alpha\in{\mathbb{C}} and vV1v\in V_{1}.

Let VV be a finite dimensional complex vector space. A real structure of VV is a sesqui-linear morphism κ:VV\kappa:V\to V such that κκ=idV\kappa\circ\kappa=\mathop{\rm id}\nolimits_{V}. When a real structure κ\kappa is provided with VV, we set V,κ:={vV|κ(v)=v}V_{{\mathbb{R}},\kappa}:=\{v\in V\,|\,\kappa(v)=v\}. It is naturally an {\mathbb{R}}-vector space. The naturally induced morphism V,κV{\mathbb{C}}\otimes V_{{\mathbb{R}},\kappa}\to V is an isomorphism of complex vector spaces. We have κ(αv)=α¯v\kappa(\alpha\otimes v_{{\mathbb{R}}})=\overline{\alpha}\otimes v_{{\mathbb{R}}} under the identification.

Let C,κC_{{\mathbb{R}},\kappa} be a positive definite symmetric bilinear form on V,κV_{{\mathbb{R}},\kappa}. It extends to a symmetric bilinear form CC and a Hermitian metric hh on VV in the standard ways:

C(αu,βv)=αβC,κ(u,v),h(αu,βv)=αβ¯C,κ(u,v).C(\alpha\otimes u_{{\mathbb{R}}},\beta\otimes v_{{\mathbb{R}}})=\alpha\beta C_{{\mathbb{R}},\kappa}(u_{{\mathbb{R}}},v_{{\mathbb{R}}}),\quad\quad h(\alpha\otimes u_{{\mathbb{R}}},\beta\otimes v_{{\mathbb{R}}})=\alpha\overline{\beta}C_{{\mathbb{R}},\kappa}(u_{{\mathbb{R}}},v_{{\mathbb{R}}}). (1)

Here, α,β\alpha,\beta\in{\mathbb{C}} and u,vV,κu_{{\mathbb{R}}},v_{{\mathbb{R}}}\in V_{{\mathbb{R}},\kappa}. We have C(κ(u),κ(v))=C(u,v)¯C(\kappa(u),\kappa(v))=\overline{C(u,v)} and h(κ(u),κ(v))=h(u,v)¯=h(v,u)h(\kappa(u),\kappa(v))=\overline{h(u,v)}=h(v,u).

2.1.2 Compatibility of symmetric pairings and Hermitian metrics

Let us recall the notion of compatibility of a non-degenerate symmetric pairing and a Hermitian metric on a complex vector space VV. Let VV^{\lor} denote the dual space of VV. Let ,:V×V\langle\cdot,\cdot\rangle:V^{\lor}\times V\to{\mathbb{C}} denote the canonical pairing.

Let C:V×VC:V\times V\to{\mathbb{C}} be a non-degenerate symmetric bilinear form. We obtain the linear isomorphism ΨC:VV\Psi_{C}:V\simeq V^{\lor} by ΨC(u),v=C(u,v)\langle\Psi_{C}(u),v\rangle=C(u,v). We obtain the symmetric bilinear form C:V×VC^{\lor}:V^{\lor}\times V^{\lor}\to{\mathbb{C}} by

C(u,v)=C(ΨC1(u),ΨC1(v)).C^{\lor}(u^{\lor},v^{\lor})=C(\Psi_{C}^{-1}(u^{\lor}),\Psi_{C}^{-1}(v^{\lor})).

We have ΨCΨC=idV\Psi_{C^{\lor}}\circ\Psi_{C}=\mathop{\rm id}\nolimits_{V}.

Let hh be a Hermitian metric of VV. We obtain the sesqui-linear isomorphism Ψh:VV\Psi_{h}:V\simeq V^{\lor} by Ψh(u),v=h(v,u)\langle\Psi_{h}(u),v\rangle=h(v,u). We obtain the Hermitian metric hh^{\lor} of VV^{\lor} by

h(u,v)=h(Ψh1(v),Ψh1(u)).h^{\lor}(u^{\lor},v^{\lor})=h\bigl{(}\Psi_{h}^{-1}(v^{\lor}),\Psi_{h}^{-1}(u^{\lor})\bigr{)}.

It is easy to see that ΨhΨh=idV\Psi_{h^{\lor}}\circ\Psi_{h}=\mathop{\rm id}\nolimits_{V}.

Definition 2.1

We say that hh is compatible with CC if ΨC\Psi_{C} is isometric with respect to hh and hh^{\lor}. Let Herm(V,C)\mathop{\rm Herm}\nolimits(V,C) denote the space of Hermitian metrics of VV which are compatible with CC.  

Lemma 2.2

The following conditions are equivalent.

  • hh is compatible with CC.

  • C(u,v)=C(Ψh(u),Ψh(v))¯C(u,v)=\overline{C^{\lor}(\Psi_{h}(u),\Psi_{h}(v))} holds for any u,vVu,v\in V.

  • ΨCΨh=ΨhΨC\Psi_{C^{\lor}}\circ\Psi_{h}=\Psi_{h^{\lor}}\circ\Psi_{C} holds. It is also equivalent to ΨCΨh=ΨhΨC\Psi_{C}\circ\Psi_{h^{\lor}}=\Psi_{h}\circ\Psi_{C^{\lor}}.

Proof   Let fVf\in V^{\lor} and uVu\in V. We have

f,ΨCΨh(u)=C(f,Ψh(u))=Ψh(u),ΨC(f)=h(ΨC1(f),u).\langle f,\Psi_{C^{\lor}}\circ\Psi_{h}(u)\rangle=C^{\lor}(f,\Psi_{h}(u))=\langle\Psi_{h}(u),\Psi_{C}^{\lor}(f)\rangle=h(\Psi_{C}^{-1}(f),u).

We also have

f,ΨhΨC(u)=h(f,ΨC(u))=h(ΨC(u),f)¯=Ψh(f),ΨC(u)¯=C(u,Ψh1(f))¯.\langle f,\Psi_{h^{\lor}}\circ\Psi_{C}(u)\rangle=h^{\lor}(f,\Psi_{C}(u))=\overline{h^{\lor}(\Psi_{C}(u),f)}=\overline{\langle\Psi_{h^{\lor}}(f),\Psi_{C}(u)\rangle}=\overline{C(u,\Psi_{h}^{-1}(f))}.

Hence, we obtain the claim of Lemma 2.2.  

From a non-degenerate symmetric bilinear form CC and a Hermitian metric hh, we obtain a sesqui-linear isomorphism κ=ΨCΨh\kappa=\Psi_{C^{\lor}}\circ\Psi_{h}.

Lemma 2.3

hh is compatible with CC if and only if κ\kappa is a real structure of VV.

Proof   We have κ1=Ψh1ΨC1=ΨhΨC\kappa^{-1}=\Psi_{h}^{-1}\circ\Psi_{C^{\lor}}^{-1}=\Psi_{h^{\lor}}\circ\Psi_{C}. Hence, the claim follows from Lemma 2.2.  

Lemma 2.4

Suppose that hh is compatible with CC.

  • We have h(κ(u),κ(v))=h(u,v)¯=h(v,u)h(\kappa(u),\kappa(v))=\overline{h(u,v)}=h(v,u) and h(u,v)=C(u,κ(v))h(u,v)=C(u,\kappa(v)) for any u,vVu,v\in V.

  • For any u,vV,κu,v\in V_{{\mathbb{R}},\kappa}, we have C(u,v)=h(u,v)C(u,v)=h(u,v)\in{\mathbb{R}}.

  • Let C,κC_{{\mathbb{R}},\kappa} denote the {\mathbb{R}}-valued symmetric bilinear form on V,κV_{{\mathbb{R}},\kappa} obtained as the restriction of CC. Then, it is also equal to the restriction of hh to V,κV_{{\mathbb{R}},\kappa}. Moreover, hh and CC are related with C,κC_{{\mathbb{R}},\kappa} as in (1).

Proof   By the constructions, we have

h(κ(u),κ(v))=h(ΨC1Ψh(u),ΨC1Ψh(v))=h(Ψh(u),Ψh(v))=h(v,u).h(\kappa(u),\kappa(v))=h\bigl{(}\Psi_{C}^{-1}\Psi_{h}(u),\Psi_{C}^{-1}\Psi_{h}(v)\bigr{)}=h^{\lor}(\Psi_{h}(u),\Psi_{h}(v))=h(v,u).

We also have

C(u,κ(v))=C(κ(v),u)=ΨC(ΨC1Ψh(v)),u=Ψh(v),u=h(u,v).C(u,\kappa(v))=C(\kappa(v),u)=\langle\Psi_{C}\bigl{(}\Psi_{C}^{-1}\circ\Psi_{h}(v)\bigr{)},u\rangle=\langle\Psi_{h}(v),u\rangle=h(u,v).

If u,vV,κu,v\in V_{{\mathbb{R}},\kappa}, we obtain h(u,v)=h(κ(u),κ(v))=h(v,u)=h(u,v)¯h(u,v)=h(\kappa(u),\kappa(v))=h(v,u)=\overline{h(u,v)}, and hence h(u,v)h(u,v)\in{\mathbb{R}}. We also have h(u,v)=C(u,κ(v))=C(u,v)h(u,v)=C(u,\kappa(v))=C(u,v). Thus, we obtain the second claim. The third claim follows easily.  

Corollary 2.5

hh is compatible with CC if and only if there exists a base of VV which is orthonormal with respect to both hh and CC.

Proof   The “only if” part follows from Lemma 2.4. The “if” part is easy to see.  

2.1.3 Hermitian automorphisms

Let hh be a Hermitian metric of VV. Let (V,h)\mathcal{H}(V,h) be the set of automorphisms ff of VV which are Hermitian with respect to hh, i.e., h(fu,v)=h(u,fv)h(fu,v)=h(u,fv) for any u,vVu,v\in V. Let +(V,h)(V,h)\mathcal{H}_{+}(V,h)\subset\mathcal{H}(V,h) be the subset of f(V,h)f\in\mathcal{H}(V,h) such that any eigenvalues of ff are positive. There exists the exponential map exp:(V,h)+(V,h)\exp:\mathcal{H}(V,h)\to\mathcal{H}_{+}(V,h) defined by exp(H)=m=01m!Hm\exp(H)=\sum_{m=0}^{\infty}\frac{1}{m!}H^{m}. We also have the well defined logarithm log:+(V,h)(V,h)\log:\mathcal{H}_{+}(V,h)\to\mathcal{H}(V,h) such that logexp=id(V,h)\log\circ\exp=\mathop{\rm id}\nolimits_{\mathcal{H}(V,h)} and logexp=id+(V,h)\log\circ\exp=\mathop{\rm id}\nolimits_{\mathcal{H}_{+}(V,h)}. For H𝒫+(V,h)H\in\mathcal{P}_{+}(V,h), there exist a unitary matrix PP and a diagonal matrix GG such that H=PGP1H=PGP^{-1}. The (i,i)(i,i)-entries Gi,iG_{i,i} of GG are positive. We have log(H)=Plog(G)P1\log(H)=P\log(G)P^{-1}, where log(G)\log(G) is the diagonal matrix whose (i,i)(i,i)-entries are log(Gi,i)\log(G_{i,i}). We recall the following well known lemma.

Lemma 2.6

The exponential map exp:(V,h)+(V,h)\exp:\mathcal{H}(V,h)\to\mathcal{H}_{+}(V,h) is a diffeomorphism.

Proof   It is easy to see that the exponential function is CC^{\infty}. It is enough to check that the logarithm is also CC^{\infty}. For H+(V,h)H\in\mathcal{H}_{+}(V,h), we have

log(H)=12π1Γlog(z)(zidVH)1𝑑z.\log(H)=\frac{1}{2\pi\sqrt{-1}}\int_{\Gamma}\log(z)\cdot(z\mathop{\rm id}\nolimits_{V}-H)^{-1}dz. (2)

Here, Γ\Gamma is the union of small circles with the counter clock-wise direction around the eigenvalues of HH in {\mathbb{C}}. Hence, log:+(V,h)(V,h)\log:\mathcal{H}_{+}(V,h)\to\mathcal{H}(V,h) is also CC^{\infty}.  

Let CC be a non-degenerate symmetric bilinear form on VV. Suppose that hh is compatible with CC. Let (V,C,h)\mathcal{H}(V,C,h) be the set of f(V,h)f\in\mathcal{H}(V,h) which are anti-symmetric with respect to CC, i.e., C(fu,v)+C(u,fv)=0C(fu,v)+C(u,fv)=0 for any u,vVu,v\in V. Let +(V,C,h)\mathcal{H}_{+}(V,C,h) be the set of f+(V,h)f\in\mathcal{H}_{+}(V,h) which are isometric with respect to CC, i.e., C(fu,fv)=C(u,v)C(fu,fv)=C(u,v) for any u,vVu,v\in V.

Lemma 2.7

The exponential map induces a diffeomorphism exp:(V,C,h)+(V,C,h)\exp:\mathcal{H}(V,C,h)\to\mathcal{H}_{+}(V,C,h).

Proof   We set n:=dimVn:=\dim_{{\mathbb{C}}}V. Let InI_{n} denote the identity (n×n)(n\times n)-matrix. Let 0\mathcal{H}_{0} denote the space of Hermitian (n×n)(n\times n)-matrices. Let 0,+0\mathcal{H}_{0,+}\subset\mathcal{H}_{0} denote the space of positive definite Hermitian (n×n)(n\times n)-matrices. Let 10\mathcal{H}_{1}\subset\mathcal{H}_{0} denote the subspace of H0H\in\mathcal{H}_{0} such that H+Ht=0H+{}^{t}\!H=0. Let 1,+0,+\mathcal{H}_{1,+}\subset\mathcal{H}_{0,+} denote the subspace of H0H\in\mathcal{H}_{0} such that HHt=InH\cdot{}^{t}\!H=I_{n}.

The exponential map induces a diffeomorphism exp:00,+\exp:\mathcal{H}_{0}\to\mathcal{H}_{0,+}. It induces a CC^{\infty}-map 11,+\mathcal{H}_{1}\to\mathcal{H}_{1,+}. Let H1,+H\in\mathcal{H}_{1,+}. We note that H¯=Ht\overline{H}={}^{t}\!H. There exist a diagonal matrix GG and a unitary matrix PP such that H=PGP1H=PGP^{-1}. The diagonal entries of GG are positive numbers. We have

GP¯1PG=P¯1P.G\overline{P}^{-1}\cdot PG=\overline{P}^{-1}\cdot P.

It implies that (P¯1P)i,j=0(\overline{P}^{-1}\cdot P)_{i,j}=0 unless Gi,iGj,j=1G_{i,i}\cdot G_{j,j}=1. Hence, we obtain

log(G)(P¯1P)+(P¯1P)logG=0.\log(G)\cdot(\overline{P}^{-1}\cdot P)+(\overline{P}^{-1}\cdot P)\cdot\log G=0.

We obtain P¯(logG)P¯1+P(logG)P1=0\overline{P}(\log G)\overline{P}^{-1}+P\cdot(\log G)P^{-1}=0, i.e., (logH)t+logH=(logH)¯+logH=0{}^{t}\!(\log H)+\log H=\overline{(\log H)}+\log H=0. Hence, the exponential map induces a diffeomorphism 11,+\mathcal{H}_{1}\longrightarrow\mathcal{H}_{1,+}.

Let 𝒗{\boldsymbol{v}} be a base of V,κV_{{\mathbb{R}},\kappa} which is orthonormal with respect to C,κC_{{\mathbb{R}},\kappa}. Any f+(V,C,h)f\in\mathcal{H}_{+}(V,C,h) (resp. f(V,C,h)f\in\mathcal{H}(V,C,h)) is represented by a matrix in 1,+\mathcal{H}_{1,+} (resp. 1\mathcal{H}_{1}) with respect to 𝒗{\boldsymbol{v}}. Hence, the exponential map induces a diffeomorphism (V,C,h)+(V,C,h)\mathcal{H}(V,C,h)\simeq\mathcal{H}_{+}(V,C,h).  

We recall that any f+(V,h)f\in\mathcal{H}_{+}(V,h) has natural ss-powers fs=exp(slogf)f^{s}=\exp(s\log f) (s)(s\in{\mathbb{R}}). We obtain the following lemma from Lemma 2.7.

Lemma 2.8

If f+(V,C,h)f\in\mathcal{H}_{+}(V,C,h), then fs+(V,C,h)f^{s}\in\mathcal{H}_{+}(V,C,h) for any ss\in{\mathbb{R}}.  

We also obtain the following lemma.

Lemma 2.9

For any f+(V,C,h)f\in\mathcal{H}_{+}(V,C,h), we have det(f)=1\det(f)=1.

Proof   Let 𝒗{\boldsymbol{v}} be a base of VV as in the proof of Lemma 2.7. Then, ff is represented by a positive definite Hermitian matrix HH such that H¯H=In\overline{H}\cdot H=I_{n}. It implies det(H)2=det(H¯)det(H)=1\det(H)^{2}=\det(\overline{H})\det(H)=1, and hence det(H)=1\det(H)=1.  

Let f+(V,C,h)f\in\mathcal{H}_{+}(V,C,h). Let V(f,a)VV(f,a)\subset V denote the eigen space of ff corresponding to the eigenvalue aa\in{\mathbb{R}}. We obtain the decomposition V=a>0V(f,a)V=\bigoplus_{a>0}V(f,a). We set V(f,a,a1)=V(f,a)V(f,a1)V(f,a,a^{-1})=V(f,a)\oplus V(f,a^{-1}) for a>1a>1.

Lemma 2.10

The decomposition

V=V(f,1)a>1V(f,a,a1)V=V(f,1)\oplus\bigoplus_{a>1}V(f,a,a^{-1}) (3)

is orthogonal with respect to both hh and CC. The real structure κ\kappa preserves the decomposition (3). Moreover, κ\kappa exchanges V(f,a)V(f,a) and V(f,a1)V(f,a^{-1}).

Proof   Let 𝒗{\boldsymbol{v}} be a base of V,κV_{{\mathbb{R}},\kappa} which is orthonormal with respect to C,κC_{{\mathbb{R}},\kappa}. We obtain the Hermitian matrix HH representing ff with respect to 𝒗{\boldsymbol{v}}. We have H¯=H1\overline{H}=H^{-1}, i.e., κfκ=f1\kappa\circ f\circ\kappa=f^{-1}. Hence, we obtain that κ\kappa preserves (3), and that it exchanges V(f,a)V(f,a) and V(f,a1)V(f,a^{-1}). It implies that (3) is induced by a decomposition

V,κ=V(f,1),κa>1V(f,a,a1),κ.V_{{\mathbb{R}},\kappa}=V(f,1)_{{\mathbb{R}},\kappa}\oplus\bigoplus_{a>1}V(f,a,a^{-1})_{{\mathbb{R}},\kappa}. (4)

We note that (3) is orthogonal with respect to hh because it is induced by the eigen decomposition of the Hermitian automorphism ff. It implies that the decomposition (4) is orthogonal with respect to C,κC_{{\mathbb{R}},\kappa}. Hence, the decomposition (3) is orthogonal with respect to CC.  

2.1.4 Difference between two compatible Hermitian metrics

Let hh and hh^{\prime} be Hermitian metrics of VV. There exists the unique automorphism s(h,h)s(h,h^{\prime}) such that h(u,v)=h(s(h,h)u,v)h^{\prime}(u,v)=h(s(h,h^{\prime})u,v) for any u,vVu,v\in V. Note that s(h,h)s(h,h^{\prime}) is self-adjoint with respect to both hh and hh^{\prime}. Let s(h,h)s(h,h^{\prime})^{\lor} denote the automorphism of VV^{\lor} obtained as the dual of s(h,h)s(h,h^{\prime}). We have Ψh=Ψhs(h,h)=s(h,h)Ψh\Psi_{h^{\prime}}=\Psi_{h}\circ s(h,h^{\prime})=s(h,h^{\prime})^{\lor}\circ\Psi_{h}. Suppose that hh is compatible with a non-degenerate symmetric pairing CC.

Lemma 2.11

The following conditions are equivalent.

  • hh^{\prime} is compatible with CC.

  • s(h,h)s(h,h^{\prime}) is an isometry with respect to CC, i.e., C(s(h,h)u,s(h,h)v)=C(u,v)C(s(h,h^{\prime})u,s(h,h^{\prime})v)=C(u,v) for any u,vVu,v\in V.

  • κs(h,h)κ=s(h,h)1\kappa\circ s(h,h^{\prime})\circ\kappa=s(h,h^{\prime})^{-1}.

  • Let 𝒗{\boldsymbol{v}} be an orthonormal frame of V,κV_{{\mathbb{R}},\kappa} with respect to C,κC_{{\mathbb{R}},\kappa}. Let AA be the Hermitian matrix representing s(h,h)s(h,h^{\prime}) with respect to 𝒗{\boldsymbol{v}}. Then, AAtA\cdot{}^{t}\!A equals the identity matrix, which is equivalent to A1=A¯A^{-1}=\overline{A}.

Proof   To simplify the description, we set f=s(h,h)f=s(h,h^{\prime}). Because ΨC1Ψh=ΨC1Ψhf=Ψh1ΨCf\Psi_{C}^{-1}\circ\Psi_{h^{\prime}}=\Psi_{C}^{-1}\circ\Psi_{h}\circ f=\Psi_{h}^{-1}\circ\Psi_{C}\circ f and Ψh1ΨC=Ψh1(f)1ΨC\Psi_{h^{\prime}}^{-1}\circ\Psi_{C}=\Psi_{h}^{-1}\circ(f^{\lor})^{-1}\circ\Psi_{C}, hh^{\prime} is compatible with CC if and only if ΨCf=(f)1ΨC\Psi_{C}\circ f=(f^{\lor})^{-1}\circ\Psi_{C}. The latter condition is equivalent to that ff is isometry with respect to CC. Thus, the first condition is equivalent to the second. Because ΨC1Ψh=κf\Psi_{C}^{-1}\circ\Psi_{h^{\prime}}=\kappa\circ f and Ψh1ΨC=f1κ\Psi_{h^{\prime}}^{-1}\circ\Psi_{C}=f^{-1}\circ\kappa, the first condition is equivalent to the third. It is easy to see that the fourth condition is equivalent to both the second and third.  

We note that s(h,h)1/2s(h,h^{\prime})^{1/2} induces an isometry (V,h)(V,h)(V,h^{\prime})\simeq(V,h), i.e., h(s(h,h)1/2u,s(h,h)1/2v)=h(u,v)h(s(h,h^{\prime})^{1/2}u,s(h,h^{\prime})^{1/2}v)=h^{\prime}(u,v) for any u,vVu,v\in V.

Lemma 2.12

Suppose that hh^{\prime} is compatible with CC. We set κ=ΨC1Ψh\kappa^{\prime}=\Psi_{C}^{-1}\circ\Psi_{h^{\prime}}, which is a real structure of VV. Then, the following holds.

  • s(h,h)1/2s(h,h^{\prime})^{1/2} is an isometry with respect to CC.

  • s(h,h)1/2s(h,h^{\prime})^{1/2} induces an isomorphism (V,κ)(V,κ)(V,\kappa^{\prime})\simeq(V,\kappa), i.e., s(h,h)1/2κ=κs(h,h)1/2s(h,h^{\prime})^{1/2}\circ\kappa^{\prime}=\kappa\circ s(h,h^{\prime})^{1/2}.

  • There exists an {\mathbb{R}}-isomorphism s(h,h)1/2:V,κV,κs(h,h^{\prime})_{{\mathbb{R}}}^{1/2}:V_{{\mathbb{R}},\kappa^{\prime}}\simeq V_{{\mathbb{R}},\kappa} whose complexification equals s(h,h)1/2s(h,h^{\prime})^{1/2}. Moreover, s(h,h)1/2s(h,h^{\prime})^{1/2}_{{\mathbb{R}}} is an isometry (V,κ,C,κ)(V,κ,C,κ)(V_{{\mathbb{R}},\kappa^{\prime}},C_{{\mathbb{R}},\kappa^{\prime}})\simeq(V_{{\mathbb{R}},\kappa},C_{{\mathbb{R}},\kappa}).

Proof   We obtain the first claim from Lemma 2.8. By Lemma 2.11, we obtain κs(h,h)1/2κ=s(h,h)1/2\kappa\circ s(h,h^{\prime})^{1/2}\circ\kappa=s(h,h^{\prime})^{-1/2}. Because κ=s(h,h)1κ\kappa^{\prime}=s(h,h^{\prime})^{-1}\circ\kappa, we obtain κs(h,h)1/2=s(h,h)1/2κ=s(h,h)1/2κ\kappa\circ s(h,h^{\prime})^{1/2}=s(h,h^{\prime})^{-1/2}\circ\kappa=s(h,h^{\prime})^{1/2}\circ\kappa^{\prime}. The third claim follows from the first and second.  

Let hHerm(V,C)h^{\prime}\in\mathop{\rm Herm}\nolimits(V,C). There exists the decomposition V=a>0VaV=\bigoplus_{a>0}V_{a} such that (i) it is orthogonal with respect to both hh^{\prime} and hh, (ii) h|Va=ah|Vah^{\prime}_{|V_{a}}=a\cdot h_{|V_{a}}. For a>1a>1, we set V~a=VaVa1\widetilde{V}_{a}=V_{a}\oplus V_{a^{-1}}. We obtain the decomposition

V=V1a>1V~a.V=V_{1}\oplus\bigoplus_{a>1}\widetilde{V}_{a}. (5)

We obtain the following lemma from Lemma 2.10.

Lemma 2.13

The decomposition (5) is orthogonal with respect to hh, hh^{\prime} and CC. It is preserved by κ\kappa. Moreover, κ\kappa exchanges VaV_{a} and Va1V_{a^{-1}}.  

2.1.5 Regular semisimple automorphisms

Let CC be a non-degenerate symmetric bilinear form of VV. Let FF be an endomorphism of VV satisfying the following conditions.

  • FF is symmetric with respect to CC, i.e., C(FidV)=C(idVF)C(F\otimes\mathop{\rm id}\nolimits_{V})=C(\mathop{\rm id}\nolimits_{V}\otimes F).

  • FF is regular semisimple, i.e., the multiplicity of each eigenvalue of FF is 11.

We have the eigen decomposition of FF:

V=α𝒮p(F)Vα.V=\bigoplus_{\alpha\in{\mathcal{S}p}(F)}V_{\alpha}. (6)

Here, 𝒮p(F){\mathcal{S}p}(F) denotes the set of eigenvalues of FF. The following lemma is obvious but useful in this study.

Lemma 2.14

  • The decomposition (6) is orthogonal with respect to CC, and the restriction of CC to each VαV_{\alpha} is non-degenerate.

  • There uniquely exists a Hermitian metric hcanh^{\mathop{\rm can}\nolimits} of VV such that (i) hcanh^{\mathop{\rm can}\nolimits} is compatible with CC, (ii) the decomposition (6) is orthogonal. The metric hcanh^{\mathop{\rm can}\nolimits} is called the canonical compatible metric of (V,C,F)(V,C,F).  

We set c0(F):=max{|α||α𝒮p(F)}c_{0}(F):=\max\bigl{\{}|\alpha|\,\big{|}\,\alpha\in{\mathcal{S}p}(F)\bigr{\}} and c1(F):=min{|αβ||α,β𝒮p(F),αβ}c_{1}(F):=\min\bigl{\{}|\alpha-\beta|\,\big{|}\,\alpha,\beta\in{\mathcal{S}p}(F),\,\,\alpha\neq\beta\bigr{\}}.

Proposition 2.15

There exists a constant B>0B>0 depending only on ci(F)c_{i}(F) (i=0,1)(i=0,1) and dimV\dim V such that the following holds for any hHerm(V,C)h\in\mathop{\rm Herm}\nolimits(V,C):

|s(hcan,h)|hcan+|s(hcan,h)1|hcanB(1+|F|h)dimV.|s(h^{\mathop{\rm can}\nolimits},h)|_{h^{\mathop{\rm can}\nolimits}}+|s(h^{\mathop{\rm can}\nolimits},h)^{-1}|_{h^{\mathop{\rm can}\nolimits}}\leq B\cdot(1+|F|_{h})^{\dim V}.

Proof   We set n:=dimVn:=\dim V. We take an ordering {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\} of 𝒮p(F){\mathcal{S}p}(F). Let eie_{i} be a base of VαiV_{\alpha_{i}} such that hcan(ei,ei)=1h^{\mathop{\rm can}\nolimits}(e_{i},e_{i})=1 and C(ei,ei)=1C(e_{i},e_{i})=1. By the construction, 𝒆=(e1,,en){\boldsymbol{e}}=(e_{1},\ldots,e_{n}) is an orthonormal base of VV with respect to both hcanh^{\mathop{\rm can}\nolimits} and CC. Let 𝒗=(v1,,vn){\boldsymbol{v}}=(v_{1},\ldots,v_{n}) be a base of VV which is orthonormal with respect to both hh and CC. (See Corollary 2.5.)

Let KK be a matrix determined by 𝒗=𝒆K{\boldsymbol{v}}={\boldsymbol{e}}\,K. Because both 𝒗{\boldsymbol{v}} and 𝒆{\boldsymbol{e}} are orthonormal with respect to CC, we have Kt=K1{}^{t}\!K=K^{-1}.

For any 0\ell\in{\mathbb{Z}}_{\geq 0}, let A(F,𝒗)A(F^{\ell},{\boldsymbol{v}}) be the matrix representing FF^{\ell} with respect to 𝒗{\boldsymbol{v}}. Let Γ\Gamma be the diagonal (n×n)(n\times n)-matrix whose (i,i)(i,i)-th entries are αi\alpha_{i}. We have

A(F,𝒗)=K1ΓK=KtΓK.A(F^{\ell},{\boldsymbol{v}})=K^{-1}\Gamma^{\ell}K={}^{t}\!K\Gamma^{\ell}K.

Let (A(F,𝒗))i,j(A(F^{\ell},{\boldsymbol{v}}))_{i,j} denote the (i,j)(i,j)-entry of A(F,𝒗)A(F^{\ell},{\boldsymbol{v}}). We have

A(F,𝒗)i,j=k=1nαkKk,iKk,j.A(F^{\ell},{\boldsymbol{v}})_{i,j}=\sum_{k=1}^{n}\alpha_{k}^{\ell}K_{k,i}K_{k,j}.

Because 𝒗{\boldsymbol{v}} is orthonormal with respect to hh, there exists B1B_{1} depending only on nn such that

|A(F,𝒗)i,j|B1(1+|F|h).|A(F^{\ell},{\boldsymbol{v}})_{i,j}|\leq B_{1}(1+\bigl{|}F\bigr{|}_{h})^{\ell}.

Let W(α1,,αn)W(\alpha_{1},\ldots,\alpha_{n}) be the (n×n)(n\times n)-matrix whose (i,j)(i,j)-th entry is αji1\alpha_{j}^{i-1}. It is invertible, and we obtain

Kk,i2==0n1(W(α1,,αn)1)k,A(F,𝒗)i,i.K_{k,i}^{2}=\sum_{\ell=0}^{n-1}\Bigl{(}W(\alpha_{1},\ldots,\alpha_{n})^{-1}\Bigr{)}_{k,\ell}\cdot A(F^{\ell},{\boldsymbol{v}})_{i,i}.

There exists B2>0B_{2}>0 depending only on nn and ci(F)c_{i}(F) (i=0,1)(i=0,1) such that

|Kk,i|2B2(1+|F|h)n.|K_{k,i}|^{2}\leq B_{2}(1+|F|_{h})^{n}.

We set H(h,𝒆)i,j=h(ei,ej)H(h,{\boldsymbol{e}})_{i,j}=h(e_{i},e_{j}). We obtain the Hermitian matrix H(h,𝒆)=(H(h,𝒆)i,j)H(h,{\boldsymbol{e}})=(H(h,{\boldsymbol{e}})_{i,j}). We obtain s(hcan,h)𝒆=𝒆Ht(h,𝒆)s(h^{\mathop{\rm can}\nolimits},h){\boldsymbol{e}}={\boldsymbol{e}}\cdot{}^{t}\!H(h,{\boldsymbol{e}}). Because 𝒗{\boldsymbol{v}} is orthonormal with respect to hh, we obtain H(h,𝒆)=K1tK1¯=KK¯tH(h,{\boldsymbol{e}})={}^{t}\!K^{-1}\cdot\overline{K^{-1}}=K\cdot{}^{t}\!\overline{K}. Therefore, there exists B3>0B_{3}>0 depending only on nn and ci(F)c_{i}(F) (i=0,1)(i=0,1) such that

|H(h,𝒆)i,j|B3(1+|F|h)n.|H(h,{\boldsymbol{e}})_{i,j}|\leq B_{3}(1+|F|_{h})^{n}.

Thus, we obtain the claim for |s(hcan,h)|hcan|s(h^{\mathop{\rm can}\nolimits},h)|_{h^{\mathop{\rm can}\nolimits}}. Because Ht(h,𝒆)1=H(h,𝒆){}^{t}\!H(h,{\boldsymbol{e}})^{-1}=H(h,{\boldsymbol{e}}), we also obtain the claim for |s(hcan,h)1|hcan|s(h^{\mathop{\rm can}\nolimits},h)^{-1}|_{h^{\mathop{\rm can}\nolimits}}. Thus, we obtain Proposition 2.15.  

2.1.6 Vector bundles

Let MM be a paracompact CC^{\infty}-manifold. Let VV be a complex vector bundle on MM with a non-degenerate symmetric pairing CC. A Hermitian metric hh of VV is called compatible with CC if h|Ph_{|P} is compatible with C|PC_{|P} for any PMP\in M.

Lemma 2.16

There exists a Hermitian metric hh of VV compatible with CC.

Proof   For any PMP\in M, there exist a neighbourhood MPM_{P} of PP around MM and a frame 𝒗P{\boldsymbol{v}}_{P} of V|MPV_{|M_{P}} which is orthonormal with respect to C|MPC_{|M_{P}}. There exists a Hermitian metric of V|MPV_{|M_{P}} which is compatible with C|MPC_{|M_{P}}.

Let UiMU_{i}\subset M (i=1,2)(i=1,2) be open subsets. Suppose that there exist Hermitian metrics hih_{i} of V|UiV_{|U_{i}} which are compatible with C|UiC_{|U_{i}}. By using Lemma 2.7 and a partition of unity on U1U2U_{1}\cup U_{2} subordinated to {U1,U2}\{U_{1},U_{2}\}, we can construct a Hermitian metric hh of V|U1U2V_{|U_{1}\cup U_{2}} which is compatible with C|U1U2C_{|U_{1}\cup U_{2}}.

There exists a locally finite open covering {Ui|i=1,2,,}\{U_{i}\,|\,i=1,2,\ldots,\} of XX such that V|UiV_{|U_{i}} has a Hermitian metric which is compatible with C|UiC_{|U_{i}}. Then, we can inductively prove the existence of a Hermitian metric of V|i=1mUiV_{|\bigcup_{i=1}^{m}U_{i}} which is a compatible with C|i=1mUiC_{|\bigcup_{i=1}^{m}U_{i}}. We can obtain a compatible Hermitian metric of VV as a limit.  

2.1.7 Appendix: Compatibility of Hermitian metric and skew-symmetric pairing

Let ω\omega be a symplectic form of a finite dimensional complex vector space VV. We obtain the isomorphism Ψω:VV\Psi_{\omega}:V\longrightarrow V^{\lor} by Ψω(u),v=ω(u,v)\langle\Psi_{\omega}(u),v\rangle=\omega(u,v). We obtain the induced symplectic form ω\omega^{\lor} of VV^{\lor} by

ω(u,v)=ω(Ψω1(u),Ψω1(v)).\omega^{\lor}(u^{\lor},v^{\lor})=\omega\bigl{(}\Psi_{\omega}^{-1}(u^{\lor}),\Psi_{\omega}^{-1}(v^{\lor})\bigr{)}.

We have ΨωΨω=idV\Psi_{\omega^{\lor}}\circ\Psi_{\omega}=-\mathop{\rm id}\nolimits_{V}. We also have (ω)=ω(\omega^{\lor})^{\lor}=\omega. We say that a Hermitian metric hh is compatible with ω\omega if Ψω\Psi_{\omega} is isometric with respect to hh and hh^{\lor}. The following lemma is similar to Lemma 2.2.

Lemma 2.17

The following conditions are equivalent.

  • hh is compatible with ω\omega.

  • ΨhΨω=ΨωΨh\Psi_{h^{\lor}}\circ\Psi_{\omega}=\Psi_{\omega^{\lor}}\circ\Psi_{h}.

  • ω(u,v)=ω(Ψh(u),Ψh(v))¯\omega(u,v)=\overline{\omega^{\lor}(\Psi_{h}(u),\Psi_{h}(v))} holds for any u,vVu,v\in V.  

From a symplectic form ω\omega and a Hermitian metric hh, we obtain a sesqui-linear isomorphism κ=ΨhΨω\kappa=\Psi_{h^{\lor}}\circ\Psi_{\omega}.

Lemma 2.18

hh is compatible with ω\omega if and only if κκ=idV\kappa\circ\kappa=-\mathop{\rm id}\nolimits_{V}. If hh is compatible with ω\omega, the following holds.

  • h(κ(u),κ(v))=h(u,v)¯h(\kappa(u),\kappa(v))=\overline{h(u,v)} and ω(u,κ(v))=h(u,v)\omega(u,\kappa(v))=h(u,v) for any u,vVu,v\in V.  

Suppose that hh is compatible with ω\omega. We have κ(1u)=1κ(u)\kappa(\sqrt{-1}u)=-\sqrt{-1}\kappa(u) and κ(1κ(1u))=u\kappa(\sqrt{-1}\kappa(\sqrt{-1}u))=-u for any uVu\in V. Hence, VV is naturally a left quaternionic vector space. For any vVv\in V such that h(v,v)=1h(v,v)=1, we obtain the 22-dimensional {\mathbb{C}}-vector space vV\mathbb{H}\cdot v\subset V generated by vv and κ(v)\kappa(v). We have ω(v,κ(v))=h(v,v)=1\omega(v,\kappa(v))=h(v,v)=1. In particular, the restriction of ω\omega to v\mathbb{H}\cdot v is a symplectic form. We also have h(v,κ(v))=ω(v,κ(v))=0h(v,\kappa(v))=-\omega(v,\kappa(v))=0. For any uVu\in V, we have h(u,v)=ω(u,κ(v))h(u,v)=\omega(u,\kappa(v)) and h(u,κ(v))=ω(u,v)h(u,\kappa(v))=-\omega(u,v). The orthogonal complement of ω\mathbb{H}\cdot\omega with respect to ω\omega is equal to the orthogonal complement of ω\mathbb{H}\cdot\omega with respect to hh. There exists a decomposition V=i=1mviV=\bigoplus_{i=1}^{m}\mathbb{H}\cdot v_{i} such that (i) the decomposition is orthogonal with respect to both ω\omega and hh, (ii) each vi\mathbb{H}\cdot v_{i} has a base vi,κ(vi)v_{i},\kappa(v_{i}), (iii) h(vi,vi)=ω(vi,κ(vi))=1h(v_{i},v_{i})=\omega(v_{i},\kappa(v_{i}))=1 and h(vi,κ(vi))=0h(v_{i},\kappa(v_{i}))=0.

Lemma 2.19

hh is compatible with ω\omega if and only if there exists a base 𝐯{\boldsymbol{v}} of VV such that (i) 𝐯{\boldsymbol{v}} is orthonormal with respect to hh, (ii) 𝐯{\boldsymbol{v}} is symplectic with respect to ω\omega.  

Suppose that ω\omega and hh are compatible. Let f+(V,h)f\in\mathcal{H}_{+}(V,h) such that ff preserves ω\omega, i.e., ω(fu,fv)=ω(u,v)\omega(fu,fv)=\omega(u,v) for any u,vVu,v\in V. We obtain the eigen decomposition V=a>0V(f,a)V=\bigoplus_{a>0}V(f,a) of ff. By setting V(f,a,a1)=V(f,a)V(f,a1)V(f,a,a^{-1})=V(f,a)\oplus V(f,a^{-1}), we obtain a decomposition (3). The following lemma is similar to Lemma 2.10.

Lemma 2.20

We have κf=f1κ\kappa\circ f=f^{-1}\circ\kappa and κ(V(f,a))=V(f,a1)\kappa\bigl{(}V(f,a)\bigr{)}=V(f,a^{-1}). The decomposition (3) is orthogonal with respect to both ω\omega and hh. Moreover, V(f,a)V(f,a) and V(f,a1)V(f,a^{-1}) are Lagrangian with respect to ω\omega, and they are orthogonal with respect to hh.  

We also have the converse. Let V=a1UaV=\bigoplus_{a\geq 1}U_{a} be a decomposition which is orthogonal with respect to both ω\omega and hh. For a>1a>1, let Ua,1UaU_{a,1}\subset U_{a} be a Lagrangian subspace with respect to ω|Ua\omega_{|U_{a}}, and we set Ua,2=κ(Ua,1)U_{a,2}=\kappa(U_{a,1}), which is the orthogonal complement of Ua,1U_{a,1} in UaU_{a} with respect to hh. Let gg be the automorphism of VV defined by

g=idU1a>1(aidUa,1a1idUa,2).g=\mathop{\rm id}\nolimits_{U_{1}}\oplus\bigoplus_{a>1}\bigl{(}a\mathop{\rm id}\nolimits_{U_{a,1}}\oplus a^{-1}\mathop{\rm id}\nolimits_{U_{a,2}}\bigr{)}.

Then, gg is Hermitian with respect to hh, and gg preserves ω\omega.

Let hh^{\prime} be another Hermitian metric of VV compatible with ω\omega. Note that the automorphism s(h,h)s(h,h^{\prime}) is Hermitian with respect to hh and that s(h,h)s(h,h^{\prime}) preserves ω\omega. We obtain the decomposition V=a>0VaV=\bigoplus_{a>0}V_{a} such that (i) the decomposition is orthogonal with respect to both hh and hh^{\prime}, (ii) h=ahh^{\prime}=ah on VaV_{a}. We set V~a=VaVa1\widetilde{V}_{a}=V_{a}\oplus V_{a^{-1}} for a>1a>1.

Lemma 2.21

The decomposition V=V1a>1V~aV=V_{1}\oplus\bigoplus_{a>1}\widetilde{V}_{a} is orthogonal with respect to hh, hh^{\prime} and ω\omega. It is preserved by κ\kappa. The subspaces VaV_{a} and Va1V_{a^{-1}} of V~a\widetilde{V}_{a} are Lagrangian with respect to ω\omega, and κ\kappa exchanges VaV_{a} and Va1V_{a^{-1}}.  

2.2 Harmonic bundles with real structure

2.2.1 GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundles

Let us recall the notion of GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle. Let YY be a Riemann surface. Let π:Y~Y\pi:\widetilde{Y}\to Y be a universal covering. We fix QYQ\in Y and Q~π1(Q)\widetilde{Q}\in\pi^{-1}(Q).

Let VV_{{\mathbb{R}}} be an {\mathbb{R}}-vector bundle of rank nn on YY equipped with a flat connection \nabla_{{\mathbb{R}}}. Let CC_{{\mathbb{R}}} be a positive definite symmetric bilinear form of VV_{{\mathbb{R}}}. There exists a \nabla_{{\mathbb{R}}}-flat trivialization π1(V)Y~×n\pi^{-1}(V_{{\mathbb{R}}})\simeq\widetilde{Y}\times{\mathbb{R}}^{n}. From π1C\pi^{-1}C_{{\mathbb{R}}}, we obtain a π1(Y,Q)\pi_{1}(Y,Q)-equivariant map FC:Y~GL(n,)/O(n)F_{C_{{\mathbb{R}}}}:\widetilde{Y}\to\mathop{\rm GL}\nolimits(n,{\mathbb{R}})/O(n), where we naturally identify GL(n,)/O(n)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})/O(n) with the space of positive definite symmetric bilinear forms. If FCF_{C_{{\mathbb{R}}}} is harmonic with respect to a Kähler metric gY~g_{\widetilde{Y}} of Y~\widetilde{Y} and the natural Riemannian metric of GL(n,)/O(n)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})/O(n), CC_{{\mathbb{R}}} is called a harmonic metric of (V,)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}}). The condition is independent of the choice of gY~g_{\widetilde{Y}}. Such a tuple (V,,C)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) is called a GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle.

Lemma 2.22

(V,,C)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) is a GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle on YY if and only if (V,,h)(V_{{\mathbb{R}}}\otimes{\mathbb{C}},\nabla,h) is a harmonic bundle, i.e., the induced map Y~GL(n,)/U(n)\widetilde{Y}\to\mathop{\rm GL}\nolimits(n,{\mathbb{C}})/U(n) is harmonic. Here, hh denotes the Hermitian metric induced by CC_{{\mathbb{R}}}.

Proof   Because GL(n,)/O(n)GL(n,)/U(n)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})/O(n)\to\mathop{\rm GL}\nolimits(n,{\mathbb{C}})/U(n) is totally geodesic, we obtain the claim of the lemma.  

2.2.2 Real structures of harmonic bundles

Let (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) be a harmonic bundle on the Riemann surface YY.

Definition 2.23

A real structure of the harmonic bundle (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) is a holomorphic symmetric non-degenerate pairing CC of EE such that the following conditions are satisfied.

  • h|Qh_{|Q} is compatible with C|QC_{|Q} for any QYQ\in Y.

  • θ\theta is self-adjoint with respect to CC, i.e., C(θu,v)=C(u,θv)C(\theta u,v)=C(u,\theta v) for any local sections uu and vv of EE.  

Let CC be a real structure of (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h). We obtain the holomorphic isomorphism ΨC:EE\Psi_{C}:E\simeq E^{\lor} defined by ΨC(u)(v)=C(u,v)\Psi_{C}(u)(v)=C(u,v). We have the sesqui-linear isomorphism Ψh:EE\Psi_{h}:E\simeq E^{\lor} defined by Ψh(u)(v)=h(v,u)\Psi_{h}(u)(v)=h(v,u). Let κ:EE\kappa:E\simeq E be the sesqui-linear isomorphism defined by κ=Ψh1ΨC\kappa=\Psi_{h}^{-1}\circ\Psi_{C}. It is a real structure of a complex vector bundle EE, i.e., κκ=idE\kappa\circ\kappa=\mathop{\rm id}\nolimits_{E}. Let E,κE_{{\mathbb{R}},\kappa} be the κ\kappa-invariant part of EE. There exists the natural isomorphism E,κE{\mathbb{C}}\otimes_{{\mathbb{R}}}E_{{\mathbb{R}},\kappa}\simeq E. There exists a positive definite symmetric bilinear form C,κC_{{\mathbb{R}},\kappa} of EE_{{\mathbb{R}}} which induces both hh and CC.

Let h=¯E+E,h\nabla_{h}=\overline{\partial}_{E}+\partial_{E,h} denote the Chern connection of (E,¯E)(E,\overline{\partial}_{E}) with hh. Let θh\theta_{h}^{\dagger} denote the adjoint of θ\theta with respect to hh. We obtain the flat connection 𝔻h1=h+θ+θh\mathbb{D}^{1}_{h}=\nabla_{h}+\theta+\theta^{\dagger}_{h} of EE.

Lemma 2.24

κ\kappa is 𝔻h1\mathbb{D}^{1}_{h}-flat. As a result, there exists a flat connection ,κ\nabla_{{\mathbb{R}},\kappa} of E,κE_{{\mathbb{R}},\kappa} which induces 𝔻h1\mathbb{D}^{1}_{h} under the isomorphism EE,κE\simeq{\mathbb{C}}\otimes_{{\mathbb{R}}}E_{{\mathbb{R}},\kappa}.

Proof   Because CC is holomorphic, ΨC¯E=¯EΨC\Psi_{C}\circ\overline{\partial}_{E}=\overline{\partial}_{E^{\lor}}\circ\Psi_{C} holds. Because θ\theta is self-adjoint with respect to CC, we have ΨCθ=θΨC\Psi_{C}\circ\theta=\theta^{\lor}\circ\Psi_{C}. Because ΨC\Psi_{C} is an isometry with respect to hh and hh^{\lor}, we obtain ΨCE,h=E,hΨC\Psi_{C}\circ\partial_{E,h}=\partial_{E^{\lor},h^{\lor}}\circ\Psi_{C} and ΨCθh=(θ)hΨC\Psi_{C}\circ\theta^{\dagger}_{h}=(\theta^{\lor})^{\dagger}_{h^{\lor}}\circ\Psi_{C}. We have Ψhθ=(θh)Ψh\Psi_{h}\circ\theta=(\theta_{h}^{\dagger})^{\lor}\circ\Psi_{h} and Ψhθh=θΨh\Psi_{h}\circ\theta^{\dagger}_{h}=\theta^{\lor}\circ\Psi_{h}. We have

Ψh(¯E,hu),v=h(v,u)h(E,hv,u)=Ψh(u),vΨh(u),E,h(v)=E,hΨh(u),v.\langle\Psi_{h}(\overline{\partial}_{E,h}u),v\rangle=\partial h(v,u)-h(\partial_{E,h}v,u)=\partial\langle\Psi_{h}(u),v\rangle-\langle\Psi_{h}(u),\partial_{E,h}(v)\rangle=\langle\partial_{E^{\lor},h^{\lor}}\Psi_{h}(u),v\rangle.

Hence, we obtain Ψh¯E=E,hΨh\Psi_{h}\circ\overline{\partial}_{E}=\partial_{E^{\lor},h^{\lor}}\circ\Psi_{h}. Similarly, we obtain ΨhE,h=¯EΨh\Psi_{h}\circ\partial_{E,h}=\overline{\partial}_{E^{\lor}}\circ\Psi_{h}. Because κ=ΨC1Ψh\kappa=\Psi_{C}^{-1}\circ\Psi_{h}, we obtain the claim of Lemma 2.24.  

Lemma 2.25

(E,κ,,κ,C,κ)(E_{{\mathbb{R}},\kappa},\nabla_{{\mathbb{R}},\kappa},C_{{\mathbb{R}},\kappa}) is a GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle.

Proof   It follows from Lemma 2.22.  

Let (V,,C)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) be a GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle. We set V=VV_{{\mathbb{C}}}={\mathbb{C}}\otimes_{{\mathbb{R}}}V_{{\mathbb{R}}}. Let hh be the Hermitian metric of VV_{{\mathbb{C}}} induced by CC_{{\mathbb{R}}}, and let \nabla_{{\mathbb{C}}} denote the induced flat connection. Then, (V,,h)(V_{{\mathbb{C}}},\nabla_{{\mathbb{C}}},h) is a harmonic bundle. Let (V,¯V,θ)(V_{{\mathbb{C}}},\overline{\partial}_{V_{{\mathbb{C}}}},\theta) denote the Higgs bundle underlying (V,,h)(V_{{\mathbb{C}}},\nabla,h). Let CC denote the holomorphic perfect symmetric pairing of VV_{{\mathbb{C}}} induced by CC_{{\mathbb{R}}}. Then, CC is a real structure of the harmonic bundle (V,¯V,θ,h)(V_{{\mathbb{C}}},\overline{\partial}_{V_{{\mathbb{C}}}},\theta,h). It is easy to observe the following.

Proposition 2.26

By the constructions, GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundles are equivalent to harmonic bundles equipped with a real structure.  

Remark 2.27

Let UU\subset{\mathbb{C}} be an open neighbourhood of 0. Let (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) be a harmonic bundle on U=U{0}U^{\ast}=U\setminus\{0\} equipped with a real structure CC. Let (V,,C)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) be the corresponding GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle on UU^{\ast}. We obtain the {\mathbb{R}}-local system LL_{{\mathbb{R}}} on UU^{\ast} obtained as the sheaf of flat sections of (V,)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}}). If (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) is wild at 0, the {\mathbb{C}}-local system L=LL_{{\mathbb{C}}}=L_{{\mathbb{R}}}\otimes{\mathbb{C}} is equipped with the three level of filtrations along each ray {te1θ| 0<t<ϵ}\{te^{\sqrt{-1}\theta}\,|\,0<t<\epsilon\}, the Stokes filtrations, the parabolic filtrations, and the weight filtrations. They are important for our understanding of the associated meromorphic flat bundle. Because the filtrations are described by the growth orders of flat sections with respect to the harmonic metric, we can easily observe that they are induced by the filtrations of LL_{{\mathbb{R}}} along the ray.  

2.2.3 SL(n,)\mathop{\rm SL}\nolimits(n,{\mathbb{R}})-harmonic bundles

Let ¯Y\underline{{\mathbb{R}}}_{Y} denote the product line bundle Y×Y\times{\mathbb{R}}. It has a naturally defined flat connection \nabla_{{\mathbb{R}}} and a positive definite symmetric pairing CC_{{\mathbb{R}}}. The tuple (¯Y,,C)(\underline{{\mathbb{R}}}_{Y},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) is a GL(1,)\mathop{\rm GL}\nolimits(1,{\mathbb{R}})-harmonic bundle. A GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundle (V,,C)(V_{{\mathbb{R}}},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) is called an SL(n,)\mathop{\rm SL}\nolimits(n,{\mathbb{R}})-harmonic bundle when an isomorphism (det(V),,C)(¯Y,,C)(\det(V_{{\mathbb{R}}}),\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}})\simeq(\underline{{\mathbb{R}}}_{Y},\nabla_{{\mathbb{R}}},C_{{\mathbb{R}}}) is equipped.

Let h0,Yh_{0,Y} denote the Hermitian metric of 𝒪Y\mathcal{O}_{Y} defined by h0,Y(1,1)=1h_{0,Y}(1,1)=1. Let C0,YC_{0,Y} be the holomorphic symmetric pairing of 𝒪Y\mathcal{O}_{Y} defined by C0,Y(1,1)=1C_{0,Y}(1,1)=1. Then, (𝒪Y,0,h0,Y)(\mathcal{O}_{Y},0,h_{0,Y}) with C0,YC_{0,Y} is a harmonic bundle with a real structure on YY.

Proposition 2.28

SL(n,)\mathop{\rm SL}\nolimits(n,{\mathbb{R}})-harmonic bundles are equivalent to harmonic bundles (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) with real structure CC equipped with an isomorphism (det(E),¯det(E),trθ,det(h),det(C))(𝒪Y,0,h0,Y,C0,Y)(\det(E),\overline{\partial}_{\det(E)},\mathop{\rm tr}\nolimits\theta,\det(h),\det(C))\simeq(\mathcal{O}_{Y},0,h_{0,Y},C_{0,Y}).  

2.2.4 Compatible harmonic metrics

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on YY. A holomorphic symmetric pairing CC of EE is called a symmetric pairing of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) if θ\theta is self-adjoint with respect to CC. (See Definition 2.23.) When (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is equipped with a symmetric pairing CC, we say that a harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is compatible with CC if CC is a real structure of the harmonic bundle (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h), i.e., h|Ph_{|P} is compatible with C|PC_{|P} for any PYP\in Y. Let Harm(E,¯E,θ;C)\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C) denote the set of harmonic metrics hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC.

From hiHarm(E,¯E,θ;C)h_{i}\in\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C) (i=1,2)(i=1,2), we obtain the real structures κi\kappa_{i} of EE, and GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundles (E,κi,,κi,C,κi)(E_{{\mathbb{R}},\kappa_{i}},\nabla_{{\mathbb{R}},\kappa_{i}},C_{{\mathbb{R}},\kappa_{i}}).

Proposition 2.29

If [θ,s(h1,h2)]=0[\theta,s(h_{1},h_{2})]=0 and ¯E(s(h1,h2))=0\overline{\partial}_{E}(s(h_{1},h_{2}))=0, then s(h1,h2)1/2s(h_{1},h_{2})^{1/2} induces an isomorphism of GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundles (E,κ2,,κ2,C,κ2)(E,κ1,,κ1,C,κ1)(E_{{\mathbb{R}},\kappa_{2}},\nabla_{{\mathbb{R}},\kappa_{2}},C_{{\mathbb{R}},\kappa_{2}})\simeq(E_{{\mathbb{R}},\kappa_{1}},\nabla_{{\mathbb{R}},\kappa_{1}},C_{{\mathbb{R}},\kappa_{1}}).

Proof   According to Lemma 2.12, s(h1,h2)1/2s(h_{1},h_{2})^{1/2} induces an isomorphism (E,κ2,C,κ2)(E,κ1,C,κ1)(E_{{\mathbb{R}},\kappa_{2}},C_{{\mathbb{R}},\kappa_{2}})\simeq(E_{{\mathbb{R}},\kappa_{1}},C_{{\mathbb{R}},\kappa_{1}}). Because s(h1,h2)s(h_{1},h_{2}) is self-adjoint with respect to h1h_{1}, we obtain [θh1,s(h1,h2)]=0[\theta^{\dagger}_{h_{1}},s(h_{1},h_{2})]=0, E,h1(s(h1,h2))=0\partial_{E,h_{1}}(s(h_{1},h_{2}))=0, and h1(s(h1,h2))=0\nabla_{h_{1}}(s(h_{1},h_{2}))=0. The eigenvalues of s(h1,h2)s(h_{1},h_{2}) are constant. We obtain the eigen decomposition E=a>0EaE=\bigoplus_{a>0}E_{a} of s(h1,h2)s(h_{1},h_{2}), which is orthogonal with respect to both hih_{i}, and h2=ah1h_{2}=ah_{1} on EaE_{a}. The decomposition is compatible with ¯E\overline{\partial}_{E} and θ\theta. We obtain that θh2=θh1\theta^{\dagger}_{h_{2}}=\theta^{\dagger}_{h_{1}} and E,h2=E,h1\partial_{E,h_{2}}=\partial_{E,h_{1}}. We obtain 𝔻h11=𝔻h21\mathbb{D}^{1}_{h_{1}}=\mathbb{D}^{1}_{h_{2}}, and 𝔻h11(s(h1,h2)1/2)=0\mathbb{D}^{1}_{h_{1}}(s(h_{1},h_{2})^{1/2})=0. Hence, s(h1,h2)1/2s(h_{1},h_{2})^{1/2} induces an isomorphism of GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundles.  

2.2.5 Canonical harmonic metric in the regular semisimple case

Definition 2.30

The Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is called regular semisimple if the following holds for any PYP\in Y.

  • Let (YP,z)(Y_{P},z) be a holomorphic coordinate neighbourhood around PP. Let fPf_{P} be the endomorphism of E|YPE_{|Y_{P}} determined by θ=fPdz\theta=f_{P}\,dz. Then, fP|Qf_{P|Q} (QYP)(Q\in Y_{P}) are regular semisimple, i.e., the multiplicity of each eigenvalue of fP|Qf_{P|Q} is 11.  

Proposition 2.31

Suppose that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple and that it is equipped with a non-degenerate symmetric pairing. Then, there exists a unique harmonic metric hcanh^{\mathop{\rm can}\nolimits} of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) satisfying the following conditions.

  • hcanh^{\mathop{\rm can}\nolimits} is compatible with CC, i.e., hcanHarm(E,¯E,θ;C)h^{\mathop{\rm can}\nolimits}\in\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C).

  • Let PP be any point of YY. Let (YP,z)(Y_{P},z) and fPf_{P} be as in Definition 2.30. Then, the eigen decomposition of fP|Qf_{P|Q} is orthogonal with respect to h0|Qh_{0|Q}.

The metric hcanh^{\mathop{\rm can}\nolimits} is called the canonical metric of (E,¯E,θ,C)(E,\overline{\partial}_{E},\theta,C).

Proof   By Lemma 2.14, there exists a Hermitian metric hcanh^{\mathop{\rm can}\nolimits} of EE satisfying the conditions. We can easily check that it is a harmonic metric of the Higgs bundle.  

Remark 2.32

In general, there are many other harmonic metric of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC.  

2.3 An existence theorem of compatible harmonic metrics

2.3.1 Statement

Let XX be any Riemann surface. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on XX.

Definition 2.33

The Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is called generically regular semisimple if there exists a discrete subset ZXZ\subset X such that (E,¯E,θ)|XZ(E,\overline{\partial}_{E},\theta)_{|X\setminus Z} is regular semisimple.  

We shall prove the following theorem in §2.3.32.3.5.

Theorem 2.34

We assume that (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) is generically regular semisimple, and that it is equipped with a non-degenerate symmetric pairing CC. Then, there exists a harmonic metric hh of the Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) which is compatible with CC, i.e., Harm(E,¯E,θ;C)\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C)\neq\emptyset.

Remark 2.35

If XX is compact, the existence of a non-degenerate pairing implies deg(E)=0\deg(E)=0, and we can obtain the polystability of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) by using the argument in the proof of Theorem 4.12. Hence, we can prove Theorem 2.34 by using the classical theorem of Hitchin [5] and Simpson [17] (see Theorem 1.1).  

2.3.2 Compact case

First, let us study the case where XX is compact. Indeed, in this case, we can also obtain the uniqueness of a harmonic metric compatible with CC.

Proposition 2.36

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a generically regular semisimple Higgs bundle equipped with a non-degenerate symmetric pairing CC on a compact Riemann surface XX. Then, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is polystable of degree 0 and has a unique harmonic metric compatible with CC.

Proof   Because EE is equipped with a non-degenerate symmetric pairing CC, we obtain det(E)det(E)𝒪X\det(E)\otimes\det(E)\simeq\mathcal{O}_{X}, which implies deg(E)=0\deg(E)=0. Let EEE^{\prime}\subset E be a subbundle such that θ(E)EΩX1\theta(E^{\prime})\subset E^{\prime}\otimes\Omega^{1}_{X}. Let PXZP\in X\setminus Z. Let (XP,z)(X_{P},z) be a holomorphic coordinate neighbourhood around PP in XZX\setminus Z. We obtain the endomorphism ff of E|XPE_{|X_{P}} such that θ=fdz\theta=f\,dz. There exist holomorphic functions α1,,αrankE\alpha_{1},\ldots,\alpha_{\mathop{\rm rank}\nolimits E} on XPX_{P} and the eigen decomposition (E|XP,f)=i=1rankE(Ei,αiidEi)(E_{|X_{P}},f)=\bigoplus_{i=1}^{\mathop{\rm rank}\nolimits E}(E_{i},\alpha_{i}\mathop{\rm id}\nolimits_{E_{i}}). It is orthogonal with respect to C|XPC_{|X_{P}}, and hence the restriction of CC to each EiE_{i} is non-degenerate. Because E|XPE^{\prime}_{|X_{P}} is the direct sum of some of EiE_{i}, the restriction of CC to E|XPE^{\prime}_{|X_{P}} is also non-degenerate. We obtain that the restriction CC^{\prime} of CC to EE^{\prime} is non-degenerate on XZX\setminus Z. Hence, we obtain a monomorphism det(C):det(E)det(E)𝒪X\det(C^{\prime}):\det(E^{\prime})\otimes\det(E^{\prime})\to\mathcal{O}_{X}. It implies that deg(E)0\deg(E^{\prime})\leq 0. Moreover, if deg(E)=0\deg(E^{\prime})=0, we obtain that CC^{\prime} is non-degenerate on XX. We obtain the orthogonal decomposition E=EEE=E^{\prime}\oplus E^{\prime\bot} with respect to CC, which is compatible with θ\theta. Hence, we obtain that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is polystable.

There exists a decomposition

(E,¯E,θ)=(Ej,¯Ej,θj)(E,\overline{\partial}_{E},\theta)=\bigoplus(E_{j},\overline{\partial}_{E_{j}},\theta_{j}) (7)

into stable Higgs bundles of degree 0, which is orthogonal with respect to CC. Because of the generic regular semisimplicity, we obtain (Ej,¯Ej,θj)≄(Ek,¯Ek,θk)(E_{j},\overline{\partial}_{E_{j}},\theta_{j})\not\simeq(E_{k},\overline{\partial}_{E_{k}},\theta_{k}) (jk)(j\neq k). By the theorem of Hitchin and Simpson, there exists a harmonic metric h1h_{1} of (E,¯E,θ)(E,\overline{\partial}_{E},\theta). It induces a harmonic metric h1h_{1}^{\lor} of (E,¯E,θ)(E^{\lor},\overline{\partial}_{E^{\lor}},\theta^{\lor}). The decomposition (7) is orthogonal with respect to both h1h_{1} and ΨC(h1)\Psi_{C}^{\ast}(h_{1}^{\lor}). There exists cj>0c_{j}>0 such that ΨC(h1)=cj2h1\Psi_{C}^{\ast}(h_{1}^{\lor})=c_{j}^{2}h_{1} on EjE_{j}. We set h2=cjh1|Ejh_{2}=\bigoplus c_{j}\cdot h_{1|E_{j}}. Then, h2h_{2} is compatible with CC. The uniqueness is also clear.  

2.3.3 Local estimate in the regular semisimple case

For R>0R>0, we set U(R):={z||z|<R}U(R):=\{z\in{\mathbb{C}}\,|\,|z|<R\}. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle of rank rr on U(R)U(R). We obtain the endomorphism ff of EE by θ=fdz\theta=f\,dz. Assume the following.

  • There exist holomorphic functions α1,,αr\alpha_{1},\ldots,\alpha_{r} and a decomposition

    (E,f)=(Ei,αiidEi).(E,f)=\bigoplus(E_{i},\alpha_{i}\mathop{\rm id}\nolimits_{E_{i}}). (8)
  • There exist 0<A1,A20<A_{1},A_{2} such that |αi|<A1|\alpha_{i}|<A_{1} and |αiαj|>A2|\alpha_{i}-\alpha_{j}|>A_{2} (ij)(i\neq j).

Let CC be a holomorphic non-degenerate symmetric bilinear form of the Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta). The decomposition (8) is orthogonal with respect to CC. We have the canonical harmonic metric hcanh^{\mathop{\rm can}\nolimits} of (E,¯E,θ,C)(E,\overline{\partial}_{E},\theta,C) as in Proposition 2.31.

Proposition 2.37

Let 0<R1<R0<R_{1}<R. There exist positive constants CiC_{i} (i=1,2)(i=1,2) depending only on AjA_{j} (j=1,2)(j=1,2) and R1,RR_{1},R such that the following holds on U(R1)U(R_{1}) for any hHarm(E,¯E,θ;C)h\in\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C):

|f|hC1,|s(hcan,h)|hcan+|s(hcan,h)1|hcanC2.|f|_{h}\leq C_{1},\quad|s(h^{\mathop{\rm can}\nolimits},h)|_{h^{\mathop{\rm can}\nolimits}}+|s(h^{\mathop{\rm can}\nolimits},h)^{-1}|_{h^{\mathop{\rm can}\nolimits}}\leq C_{2}.

Proof   The estimate for |f|h|f|_{h} is given in [19, Lemma 2.7]. We obtain the estimate for s(hcan,h)s(h^{\mathop{\rm can}\nolimits},h) from Proposition 2.15.  

2.3.4 Local estimate in the generically regular semisimple case

Let XX, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) and CC be as in Theorem 2.34. Assume that XX is non-compact. We choose a Kähler metric gXg_{X} of XX. There exists a discrete subset ZXZ\subset X such that (E,¯E,θ)|XZ(E,\overline{\partial}_{E},\theta)_{|X\setminus Z} is regular semisimple. Let K1XK_{1}\subset X be any relatively compact open subset of XX. For simplicity, we assume that ZK1=Z\cap\partial K_{1}=\emptyset. Let K2K_{2} be a neighbourhood of the closure K¯1\overline{K}_{1} in XX such that K2Z=K1ZK_{2}\cap Z=K_{1}\cap Z. Let h1h_{1} be any Hermitian metric of EE which is compatible with CC. (See Lemma 2.16.)

Proposition 2.38

There exists B(K1,K2)>0B(K_{1},K_{2})>0 such that the following inequality holds on K1K_{1} for any hHarm((E,¯E,θ;C)|K2)h\in\mathop{\rm Harm}\nolimits((E,\overline{\partial}_{E},\theta;C)_{|K_{2}}):

|s(h1,h)|h1+|s(h1,h)1|h1B(K1,K2).\bigl{|}s(h_{1},h)\bigr{|}_{h_{1}}+\bigl{|}s(h_{1},h)^{-1}\bigr{|}_{h_{1}}\leq B(K_{1},K_{2}).

Proof   Let PK1P\in\partial K_{1}. There exists a relatively compact neighbourhood XPX_{P} of PP in K2K_{2} such that (E,¯E,θ,h)|XP(E,\overline{\partial}_{E},\theta,h)_{|X_{P}} is regular semisimple. Let hXZcanh^{\mathop{\rm can}\nolimits}_{X\setminus Z} denote the canonical harmonic metric of (E,¯E,θ)|XZ(E,\overline{\partial}_{E},\theta)_{|X\setminus Z} compatible with C|XZC_{|X\setminus Z}. By Proposition 2.37, there exists B(XP)>0B(X_{P})>0 such that the following holds on XPX_{P} for any hHarm((E,¯E,θ;C)|K2)h\in\mathop{\rm Harm}\nolimits((E,\overline{\partial}_{E},\theta;C)_{|K_{2}}):

|s(hXZcan,h)|hXZcan+|s(hXZcan,h)1|hXZcanB(XP).\bigl{|}s(h^{\mathop{\rm can}\nolimits}_{X\setminus Z},h)\bigr{|}_{h^{\mathop{\rm can}\nolimits}_{X\setminus Z}}+\bigl{|}s(h^{\mathop{\rm can}\nolimits}_{X\setminus Z},h)^{-1}\bigr{|}_{h^{\mathop{\rm can}\nolimits}_{X\setminus Z}}\leq B(X_{P}).

Let NN be a relatively compact neighbourhood of K1\partial K_{1} in PK1XP\bigcup_{P\in\partial K_{1}}X_{P}. There exists B(N)>0B(N)>0 depending only on NN such that the following holds on NN for any hHarm((E,¯E,θ;C)|K2)h\in\mathop{\rm Harm}\nolimits((E,\overline{\partial}_{E},\theta;C)_{|K_{2}}):

|s(hXZcan,h)|hXZcan+|s(hXZcan,h)1|hXZcanB(N).\bigl{|}s(h^{\mathop{\rm can}\nolimits}_{X\setminus Z},h)\bigr{|}_{h^{\mathop{\rm can}\nolimits}_{X\setminus Z}}+\bigl{|}s(h^{\mathop{\rm can}\nolimits}_{X\setminus Z},h)^{-1}\bigr{|}_{h^{\mathop{\rm can}\nolimits}_{X\setminus Z}}\leq B(N).

There exists B(N)>0B^{\prime}(N)>0 depending only on NN such that the following holds on NN for any hHarm((E,¯E,θ;C)|K2)h\in\mathop{\rm Harm}\nolimits((E,\overline{\partial}_{E},\theta;C)_{|K_{2}}):

|s(h1,h)|h1+|s(h1,h)1|h1B(N).\bigl{|}s(h_{1},h)\bigr{|}_{h_{1}}+\bigl{|}s(h_{1},h)^{-1}\bigr{|}_{h_{1}}\leq B^{\prime}(N).

We recall the following inequality [17, Lemma 3.1] on K2K_{2} for any hHarm((E,¯E,θ;C)|K2)h\in\mathop{\rm Harm}\nolimits((E,\overline{\partial}_{E},\theta;C)_{|K_{2}}):

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

There exists a function β\beta on K2K_{2} such that 1Λ¯β=|ΛF(h1)|h1\sqrt{-1}\Lambda\overline{\partial}\partial\beta=\bigl{|}\Lambda F(h_{1})\bigr{|}_{h_{1}}. We obtain

1Λ¯(logTr(s(h1,h))β)0\sqrt{-1}\Lambda\overline{\partial}\partial\Bigl{(}\log\mathop{\rm Tr}\nolimits\bigl{(}s(h_{1},h)\bigr{)}-\beta\Bigr{)}\leq 0

on K2K_{2}. By the maximum principle, we obtain

maxK1(logTr(s(h1,h))β)maxN(logTr(s(h1,h))β).\max_{K_{1}}\Bigl{(}\log\mathop{\rm Tr}\nolimits\bigl{(}s(h_{1},h)\bigr{)}-\beta\Bigr{)}\leq\max_{N}\Bigl{(}\log\mathop{\rm Tr}\nolimits\bigl{(}s(h_{1},h)\bigr{)}-\beta\Bigr{)}.

We also note that det(s(h1,h))=1\det(s(h_{1},h))=1. Hence, we obtain the claim of the lemma.  

2.3.5 Proof of Theorem 2.34

We have the isomorphism of the Higgs bundles ΨC:(E,¯E,θ)(E,¯E,θ)\Psi_{C}:(E,\overline{\partial}_{E},\theta)\simeq(E^{\lor},\overline{\partial}_{E^{\lor}},\theta^{\lor}).

Let {Xi}\{X_{i}\} be a smooth exhausting family of XX as in [10, Definition 2.5]. Let h1h_{1} be a Hermitian metric of EE which is compatible with CC. We have the induced metric h1h_{1} on EE^{\lor}. The morphism ΨC:EE\Psi_{C}:E\longrightarrow E^{\lor} is an isometry with respect to h1h_{1} and h1h_{1}^{\lor}.

By [3, Theorem 2] and [10, Proposition 2.1], there exists a unique harmonic metric hXih_{X_{i}} of (E,¯E,θ)|Xi(E,\overline{\partial}_{E},\theta)_{|X_{i}} such that hXi|Xi=h1|Xih_{X_{i}|\partial X_{i}}=h_{1|\partial X_{i}}. We have the induced harmonic metric hXih_{X_{i}}^{\lor} of (E,¯E,θ)|Xi(E^{\lor},\overline{\partial}_{E^{\lor}},\theta^{\lor})_{|X_{i}}. Because ΨC|Xi\Psi_{C|\partial X_{i}} is isometric with respect to hXi|Xih_{X_{i}|\partial X_{i}} and hXi|Xih_{X_{i}|\partial X_{i}}^{\lor}, we obtain that ΨC|Xi\Psi_{C|X_{i}} is isometric with respect to hXih_{X_{i}} and hXih_{X_{i}}^{\lor}. By [10, Proposition 2.6] and Proposition 2.38, {hXi}\{h_{X_{i}}\} has a convergent subsequence whose limit is denoted by hh. Then, hh is a harmonic metric compatible with CC.  

2.4 Some sufficient conditions for generically regular semisimplicity

2.4.1 Spectral curves

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle of rank rr on a Riemann surface XX. We may regard EE as a module over the sheaf of algebras SymΘX\mathop{\rm Sym}\nolimits\Theta_{X}, where ΘX\Theta_{X} denote the tangent sheaf of XX. There exists the coherent 𝒪TX\mathcal{O}_{T^{\ast}X}-module (E,θ)\mathcal{F}(E,\theta) with an isomorphism of SymΘX\mathop{\rm Sym}\nolimits\Theta_{X}-modules π(E,θ)E\pi_{\ast}\mathcal{F}(E,\theta)\simeq E where π:TXX\pi:T^{\ast}X\to X denote the projection. The support ΣE,θ\Sigma_{E,\theta} of (E,θ)\mathcal{F}(E,\theta) is called the spectral curve of (E,θ)(E,\theta). (See [1, 6].) Let us recall some related notions in a way convenient to us.

For any PXP\in X, let ιP:TPXTX\iota_{P}:T^{\ast}_{P}X\to T^{\ast}X denote the inclusion. We obtain the 𝒪TPX\mathcal{O}_{T^{\ast}_{P}X}-module ιP(E,θ)\iota_{P}^{\ast}\mathcal{F}(E,\theta). We have the decomposition by the supports

ιP(E,θ)=QTPXΣE,θιP(E,θ)Q,\iota_{P}^{\ast}\mathcal{F}(E,\theta)=\bigoplus_{Q\in T^{\ast}_{P}X\cap\Sigma_{E,\theta}}\iota_{P}^{\ast}\mathcal{F}(E,\theta)_{Q},

where the support of ιP(E,θ)Q\iota_{P}^{\ast}\mathcal{F}(E,\theta)_{Q} is {Q}\{Q\}. Each ιP(E,θ)Q\iota_{P}^{\ast}\mathcal{F}(E,\theta)_{Q} is naturally a finite dimensional {\mathbb{C}}-vector space. We set 𝔪(Q):=dimιP(E,θ)Q\mathfrak{m}(Q):=\dim_{{\mathbb{C}}}\iota_{P}^{\ast}\mathcal{F}(E,\theta)_{Q}. We obtain a map 𝔪:ΣE,θ>0\mathfrak{m}:\Sigma_{E,\theta}\to{\mathbb{Z}}_{>0}. We have π(Q)=P𝔪(Q)=r\sum_{\pi(Q)=P}\mathfrak{m}(Q)=r for any PXP\in X. In particular, we obtain |ΣE,θTPX|r|\Sigma_{E,\theta}\cap T_{P}^{\ast}X|\leq r. We set n(E,θ)=maxPX|Σ(E,θ)TPX|rn(E,\theta)=\max_{P\in X}|\Sigma(E,\theta)\cap T_{P}^{\ast}X|\leq r. The generically regular semisimplicity condition for (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is equivalent to n(E,θ)=rn(E,\theta)=r.

We set D(E,θ):={PX||ΣE,θTPX|<n(E,θ)}D(E,\theta):=\bigl{\{}P\in X\,\big{|}\,|\Sigma_{E,\theta}\cap T_{P}^{\ast}X|<n(E,\theta)\bigr{\}}. It is easy to see that D(E,θ)D(E,\theta) is discrete in XX. We set ΣE,θ:=ΣE,θπ1(D(E,θ))\Sigma_{E,\theta}^{\circ}:=\Sigma_{E,\theta}\setminus\pi^{-1}(D(E,\theta)). It is a closed complex submanifold of T(XD(E,θ))T^{\ast}(X\setminus D(E,\theta)), and the projection ΣE,θXD(E,θ)\Sigma_{E,\theta}^{\circ}\to X\setminus D(E,\theta) is a local homeomorphism. The function 𝔪\mathfrak{m} is locally constant on ΣE,θ>0\Sigma_{E,\theta}^{\circ}\to{\mathbb{Z}}_{>0}. We set (ΣE,θ)k:=ΣE,θ𝔪1(k)(\Sigma_{E,\theta}^{\circ})_{k}:=\Sigma_{E,\theta}^{\circ}\cap\mathfrak{m}^{-1}(k). Let (ΣE,θ)k(\Sigma_{E,\theta})_{k} denote the closure of (ΣE,θ)k(\Sigma_{E,\theta}^{\circ})_{k} in ΣE,θ\Sigma_{E,\theta}. We have ΣE,θ=k(ΣE,θ)k\Sigma_{E,\theta}=\bigcup_{k}(\Sigma_{E,\theta})_{k}.

A union of some irreducible components of ΣE,θ\Sigma_{E,\theta} is called a spectral subcurve of ΣE,θ\Sigma_{E,\theta}. For example, (ΣE,θ)k(\Sigma_{E,\theta})_{k} are spectral subcurves. A spectral subcurve of ΣE,θ\Sigma_{E,\theta} is a purely 11-dimensional complex analytic closed subset of ΣE,θ\Sigma_{E,\theta}. For a spectral subcurve Σ1\Sigma_{1}, we set Σ1:=Σ1π1(D(E,θ))\Sigma_{1}^{\circ}:=\Sigma_{1}\setminus\pi^{-1}(D(E,\theta)). It is easy to see that (i) Σ1\Sigma_{1}^{\circ} is a smooth complex submanifold of T(XD(E,θ))T^{\ast}(X\setminus D(E,\theta)), (ii) Σ1XD(E,θ)\Sigma_{1}^{\circ}\to X\setminus D(E,\theta) is a local homeomorphism, (iii) Σ1\Sigma_{1} is the closure of Σ1\Sigma_{1}^{\circ} in TXT^{\ast}X. For any PXD(E,θ)P\in X\setminus D(E,\theta), we set n(Σ1):=|Σ1TPX|n(\Sigma_{1}):=\bigl{|}\Sigma_{1}\cap T_{P}^{\ast}X\bigr{|} and r(Σ1):=QTPXΣ1𝔪(Q)r(\Sigma_{1}):=\sum_{Q\in T_{P}^{\ast}X\cap\Sigma_{1}}\mathfrak{m}(Q), which are independent of the choice of PP. In particular, we set n(k):=n((ΣE,θ)k)n(k):=n((\Sigma_{E,\theta})_{k}) and r(k):=r((ΣE,θ)k)r(k):=r((\Sigma_{E,\theta})_{k}). Note that r(k)=kn(k)r(k)=k\cdot n(k).

2.4.2 Characteristic polynomials and some Higgs subsheaves in the local case

Let (U,z)(U,z) be a connected holomorphic chart of XX, i.e., UU is a connected open subset of XX, and zz is a holomorphic coordinate on UU. Let fUf_{U} be the endomorphism of E|UE_{|U} defined by θ=fUdz\theta=f_{U}\,dz. We obtain the characteristic polynomial det(TidE|UfU)=j=0raj(z)Tj\det(T\mathop{\rm id}\nolimits_{E_{|U}}-f_{U})=\sum_{j=0}^{r}a_{j}(z)T^{j} which is a monic. Then, the polynomial

PE,θ,U(T)=j=0raj(z)(dz)rjTjj=0rH0(U,KUrj)TjP_{E,\theta,U}(T)=\sum_{j=0}^{r}a_{j}(z)(dz)^{r-j}T^{j}\in\bigoplus_{j=0}^{r}H^{0}(U,K_{U}^{r-j})T^{j}

is independent of the choice of a coordinate zz on UU. It is called the characteristic polynomial of (E,¯E,θ)|U(E,\overline{\partial}_{E},\theta)_{|U}. Under the isomorphism U×TUU\times{\mathbb{C}}\simeq T^{\ast}U induced by (z,T)(z,Tdz)(z,T)\longmapsto(z,T\,dz), the spectral curve ΣE,θTU\Sigma_{E,\theta}\cap T^{\ast}U of (E,¯E,θ)|U(E,\overline{\partial}_{E},\theta)_{|U} is equal to {(z,T)U×|aj(z)Tj=0}\{(z,T)\in U\times{\mathbb{C}}\,|\,\sum a_{j}(z)T^{j}=0\}. For (z0,αdz)ΣE,θTU(z_{0},\alpha\,dz)\in\Sigma_{E,\theta}\cap T^{\ast}U, 𝔪(z0,αdz)\mathfrak{m}(z_{0},\alpha\,dz) is equal to the multiplicity of the root α\alpha of the polynomial aj(z0)Tj\sum a_{j}(z_{0})T^{j}.

Let Σ1ΣE,θTU\Sigma_{1}\subset\Sigma_{E,\theta}\cap T^{\ast}U be a spectral subcurve. For z0UD(E,θ)z_{0}\in U\setminus D(E,\theta), we obtain the complex numbers cjΣ1(z0)c_{j}^{\Sigma_{1}}(z_{0}) (0jn(Σ1))(0\leq j\leq n(\Sigma_{1})) and c~jΣ1(z0)\widetilde{c}_{j}^{\Sigma_{1}}(z_{0}) (0jr(Σ1))(0\leq j\leq r(\Sigma_{1})) by setting

j=0n(Σ1)cjΣ1(z0)Tj=(z0,αdz)Σ1(Tα),j=0r(Σ1)c~jΣ1(z0)Tj=(z0,αdz)Σ1(Tα)𝔪(z0,αdz).\sum_{j=0}^{n(\Sigma_{1})}c_{j}^{\Sigma_{1}}(z_{0})T^{j}=\prod_{(z_{0},\alpha\,dz)\in\Sigma_{1}}(T-\alpha),\quad\quad\sum_{j=0}^{r(\Sigma_{1})}\widetilde{c}_{j}^{\Sigma_{1}}(z_{0})T^{j}=\prod_{(z_{0},\alpha\,dz)\in\Sigma_{1}}(T-\alpha)^{\mathfrak{m}(z_{0},\alpha\,dz)}.

We obtain holomorphic functions cjΣ1c_{j}^{\Sigma_{1}} and c~jΣ1\widetilde{c}_{j}^{\Sigma_{1}} on UD(E,θ)U\setminus D(E,\theta). Because the eigenvalues of fUf_{U} are locally bounded, they extend to holomorphic functions on UU, which are also denoted by cjΣ1c_{j}^{\Sigma_{1}} and c~jΣ1\widetilde{c}_{j}^{\Sigma_{1}}. We obtain the following endomorphisms of the 𝒪U\mathcal{O}_{U}-module E|UE_{|U}:

FΣ1=j=0n(Σ1)cjΣ1fUj,F~Σ1=j=0r(Σ1)c~jΣ1fUj.F^{\Sigma_{1}}=\sum_{j=0}^{n(\Sigma_{1})}c^{\Sigma_{1}}_{j}f_{U}^{j},\quad\quad\widetilde{F}^{\Sigma_{1}}=\sum_{j=0}^{r(\Sigma_{1})}\widetilde{c}^{\Sigma_{1}}_{j}f_{U}^{j}.

We obtain the 𝒪U\mathcal{O}_{U}-submodules (Σ1,0)=KerFΣ1\mathcal{F}(\Sigma_{1},0)=\mathop{\rm Ker}\nolimits F^{\Sigma_{1}} and (Σ1,1)=KerF~Σ1\mathcal{F}(\Sigma_{1},1)=\mathop{\rm Ker}\nolimits\widetilde{F}^{\Sigma_{1}}. Note that (Σ1,i)0\mathcal{F}(\Sigma_{1},i)\neq 0 (i=0,1)(i=0,1).

Lemma 2.39

(Σ1,i)\mathcal{F}(\Sigma_{1},i) (i=0,1)(i=0,1) are holomorphic subbundles of E|UE_{|U}, i.e., E|U/(Σ1,i)E_{|U}\big{/}\mathcal{F}(\Sigma_{1},i) (i=0,1)(i=0,1) are torsion-free 𝒪U\mathcal{O}_{U}-modules. We also have fU((Σ1,i))(Σ1,i)f_{U}(\mathcal{F}(\Sigma_{1},i))\subset\mathcal{F}(\Sigma_{1},i).

Proof   We obtain the first claim from E|U/(Σ1,0)ImFΣ1E|UE_{|U}\big{/}\mathcal{F}(\Sigma_{1},0)\simeq\mathop{\rm Im}\nolimits F^{\Sigma_{1}}\subset E_{|U} and E|U/(Σ1,1)ImF~Σ1E|UE_{|U}\big{/}\mathcal{F}(\Sigma_{1},1)\simeq\mathop{\rm Im}\nolimits\widetilde{F}^{\Sigma_{1}}\subset E_{|U}. We obtain the second claim from the commutativity [f,FΣ1]=[f,F~Σ1]=0[f,F^{\Sigma_{1}}]=[f,\widetilde{F}^{\Sigma_{1}}]=0.  

In particular, (Σ1,i)\mathcal{F}(\Sigma_{1},i) are Higgs subbundles of (E,θ)|U(E,\theta)_{|U}. Let θ(Σ1,i)\theta(\Sigma_{1},i) denote the Higgs field of (Σ1,i)\mathcal{F}(\Sigma_{1},i) induced by θ|U\theta_{|U}.

Lemma 2.40

We have det(Tid(Σ1,1)fU|(Σ1,1))=j=0r(Σ1)c~jΣ1Tj\det(T\mathop{\rm id}\nolimits_{\mathcal{F}(\Sigma_{1},1)}-f_{U|\mathcal{F}(\Sigma_{1},1)})=\sum_{j=0}^{r(\Sigma_{1})}\widetilde{c}^{\Sigma_{1}}_{j}T^{j}. As a result, we obtain P(Σ1,1),θ(Σ1,1)(T)=j=0r(Σ1)c~jΣ1(dz)r(Σ1)jTjP_{\mathcal{F}(\Sigma_{1},1),\theta(\Sigma_{1},1)}(T)=\sum_{j=0}^{r(\Sigma_{1})}\widetilde{c}^{\Sigma_{1}}_{j}(dz)^{r(\Sigma_{1})-j}T^{j}.

Proof   It is easy to see det(Tid(Σ1,1)fU|(Σ1,1))|UD(E,θ)=(j=0r(Σ1)c~jΣ1Tj)|UD(E,θ)\det(T\mathop{\rm id}\nolimits_{\mathcal{F}(\Sigma_{1},1)}-f_{U|\mathcal{F}(\Sigma_{1},1)})_{|U\setminus D(E,\theta)}=\Bigl{(}\sum_{j=0}^{r(\Sigma_{1})}\widetilde{c}^{\Sigma_{1}}_{j}T^{j}\Bigr{)}_{|U\setminus D(E,\theta)}. Then, we obtain the claim of the lemma.  

Remark 2.41

The rank of (Σ1,0)\mathcal{F}(\Sigma_{1},0) is not necessarily equal to n(Σ1)n(\Sigma_{1}). The characteristic polynomial of fU|(Σ1,0)f_{U|\mathcal{F}(\Sigma_{1},0)} is not necessarily equal to j=0n(Σ1)cjΣ1Tj\sum_{j=0}^{n(\Sigma_{1})}c_{j}^{\Sigma_{1}}T^{j}.  

We set ΣE,θ,k,U:=(ΣE,θ)kTU\Sigma_{E,\theta,k,U}:=(\Sigma_{E,\theta})_{k}\cap T^{\ast}U and

PE,θ,k,U(T)=j=0n(k)cjΣE,θ,k,U(z)(dz)n(k)jTj.P_{E,\theta,k,U}(T)=\sum_{j=0}^{n(k)}c^{\Sigma_{E,\theta,k,U}}_{j}(z)(dz)^{n(k)-j}T^{j}.

Note that by definition,

P(ΣE,θ,k,U,1),θ(ΣE,θ,k,U,1),U(T)=j=0r(k)c~jΣE,θ,k,U(z)(dz)r(k)jTj=PE,θ,k,U(T)k.P_{\mathcal{F}(\Sigma_{E,\theta,k,U},1),\theta(\Sigma_{E,\theta,k,U},1),U}(T)=\sum_{j=0}^{r(k)}\widetilde{c}^{\Sigma_{E,\theta,k,U}}_{j}(z)(dz)^{r(k)-j}T^{j}=P_{E,\theta,k,U}(T)^{k}.

We then have the factorization PE,θ,U(T)=k1PE,θ,k,U(T)kP_{E,\theta,U}(T)=\prod_{k\geq 1}P_{E,\theta,k,U}(T)^{k}.

2.4.3 Some sufficient conditions for generically regular semisimplicity

There uniquely exists

PE,θ(T)j=0rH0(X,KXrj)Tj(resp. PE,θ,k(T)j=0n(k)H0(X,KXn(k)j)Tj)P_{E,\theta}(T)\in\bigoplus_{j=0}^{r}H^{0}(X,K_{X}^{r-j})T^{j}\quad\left(\mbox{{r}esp. }P_{E,\theta,k}(T)\in\bigoplus_{j=0}^{n(k)}H^{0}(X,K_{X}^{n(k)-j})T^{j}\right)

such that the restriction of PE,θ(T)P_{E,\theta}(T) (resp. PE,θ,k(T)P_{E,\theta,k}(T)) to UU is PE,θ,U(T)P_{E,\theta,U}(T) (resp. PE,θ,k,U(T)P_{E,\theta,k,U}(T)) for any holomorphic chart (U,z)(U,z). We have the factorization PE,θ(T)=k1PE,θ,k(T)kP_{E,\theta}(T)=\prod_{k\geq 1}P_{E,\theta,k}(T)^{k}. The polynomial PE,θ(T)P_{E,\theta}(T) is called the characteristic polynomial of the Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta).

Proposition 2.42

The Higgs bundle (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple if the characteristic polynomial is irreducible.

Proof   The irreducibility of PE,θ(T)P_{E,\theta}(T) implies that PE,θ(T)=PE,θ,1(T)P_{E,\theta}(T)=P_{E,\theta,1}(T). Hence, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple.  

Proposition 2.43

Suppose that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) has no proper Higgs subbundle. Then, the characteristic polynomial PE,θP_{E,\theta} is irreducible. In particular, (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple.

Proof   Let Σ1ΣE,θ\Sigma_{1}\subset\Sigma_{E,\theta} be a non-empty spectral subcurve. We obtain non-zero Higgs subbundles ((Σ1,i),θ(Σ1,i))(\mathcal{F}(\Sigma_{1},i),\theta(\Sigma_{1},i)) of (E,θ)(E,\theta) such that (Σ1,i)|U=(Σ1π1(U),i)\mathcal{F}(\Sigma_{1},i)_{|U}=\mathcal{F}(\Sigma_{1}\cap\pi^{-1}(U),i) for any holomorphic chart (U,z)(U,z), where (Σ1π1(U),i)\mathcal{F}(\Sigma_{1}\cap\pi^{-1}(U),i) are as in Lemma 2.39. Because EE has no proper Higgs subbundle, we obtain (Σ1,0)=(Σ1,1)=E\mathcal{F}(\Sigma_{1},0)=\mathcal{F}(\Sigma_{1},1)=E. In particular, we obtain Σ1=ΣE,θ\Sigma_{1}=\Sigma_{E,\theta}, i.e., ΣE,θ\Sigma_{E,\theta} is irreducible. There exists k1k\geq 1 such that ΣE,θ=(ΣE,θ)k\Sigma_{E,\theta}=(\Sigma_{E,\theta})_{k}. Because ΣE,θ\Sigma_{E,\theta} is irreducible, the polynomial PE,θ,k(T)P_{E,\theta,k}(T) is irreducible. It remains to prove k=1k=1.

Let φ:X~ΣE,θ\varphi:\widetilde{X}\to\Sigma_{E,\theta} be a normalization. Because ΣE,θ\Sigma_{E,\theta} is irreducible, X~\widetilde{X} is connected. We set φ~:=πφ:X~X\widetilde{\varphi}:=\pi\circ\varphi:\widetilde{X}\to X. We obtain the Higgs bundle (E~,¯E~,θ~):=φ~(E,¯E,θ)(\widetilde{E},\overline{\partial}_{\widetilde{E}},\widetilde{\theta}):=\widetilde{\varphi}^{\ast}(E,\overline{\partial}_{E},\theta) on X~\widetilde{X}. The spectral curve ΣE~,θ~\Sigma_{\widetilde{E},\widetilde{\theta}} is equal to the image of ΣE,θ×XX~φ~(TX)\Sigma_{E,\theta}\times_{X}\widetilde{X}\subset\widetilde{\varphi}^{\ast}(T^{\ast}X) by the naturally induced morphism φ~(TX)TX~\widetilde{\varphi}^{\ast}(T^{\ast}X)\to T^{\ast}\widetilde{X}. We obtain the section sφ:X~ΣE~,θ~s_{\varphi}:\widetilde{X}\to\Sigma_{\widetilde{E},\widetilde{\theta}} induced by φ\varphi. Let Σ~1\widetilde{\Sigma}_{1} denote the image of sφs_{\varphi}. We obtain the Higgs subbundle ~:=(Σ~1,0)E~\widetilde{\mathcal{F}}:=\mathcal{F}(\widetilde{\Sigma}_{1},0)\subset\widetilde{E} with the induced Higgs field θ~\theta_{\widetilde{\mathcal{F}}}.

Lemma 2.44

There exists a holomorphic subbundle ~\mathcal{L}\subset\widetilde{\mathcal{F}} such that rank=1\mathop{\rm rank}\nolimits\mathcal{L}=1.

Proof   First, we consider the case when X~\widetilde{X} is non-compact. Because any holomorphic vector bundle on X~\widetilde{X} has a holomorphic trivialization (see [4, Theorem 30.1]), we obtain the claim of the lemma in this case. If X~\widetilde{X} is compact, let 𝒪X~(1)\mathcal{O}_{\widetilde{X}}(1) be an ample line bundle. There exists N>0N>0 such that ~(N):=~𝒪X~(N)\widetilde{\mathcal{F}}(N):=\widetilde{\mathcal{F}}\otimes\mathcal{O}_{\widetilde{X}}(N) has a non-trivial section uu. We obtain the 𝒪X~\mathcal{O}_{\widetilde{X}}-submodule ~(N)\mathcal{L}^{\prime}\subset\widetilde{\mathcal{F}}(N) generated by uu. Let ′′~(N)\mathcal{L}^{\prime\prime}\subset\widetilde{\mathcal{F}}(N) denote the 𝒪X~\mathcal{O}_{\widetilde{X}}-submodule obtained as the pull back of the torsion-part of ~(N)/\widetilde{\mathcal{F}}(N)\bigl{/}\mathcal{L}^{\prime} by the projection. Then, =′′𝒪X~(N)\mathcal{L}=\mathcal{L}^{\prime\prime}\otimes\mathcal{O}_{\widetilde{X}}(-N) satisfies the desired condition.  

We may naturally regard sφs_{\varphi} as a holomorphic 11-form on X~\widetilde{X}. Because θ~sφidE~\widetilde{\theta}-s_{\varphi}\mathop{\rm id}\nolimits_{\widetilde{E}} vanishes on ~\widetilde{\mathcal{F}}, \mathcal{L} is a Higgs subbundle of E~\widetilde{E}.

We set D:=D(E,θ)D:=D(E,\theta) and D~:=φ~1(D)\widetilde{D}:=\widetilde{\varphi}^{-1}(D). Let 𝒪X(D)\mathcal{O}_{X}(\ast D) denote the sheaf of meromorphic functions on XX whose poles are contained in DD. Similarly, let 𝒪X~(D~)\mathcal{O}_{\widetilde{X}}(\ast\widetilde{D}) denote the sheaf of meromorphic functions on X~\widetilde{X} whose poles are contained in D~\widetilde{D}. The generalized eigen decomposition induces the projection φ~(E𝒪X(D),θ)(~𝒪X~(D~),θ~)\widetilde{\varphi}^{\ast}(E\otimes\mathcal{O}_{X}(\ast D),\theta)\to(\widetilde{\mathcal{F}}\otimes\mathcal{O}_{\widetilde{X}}(\ast\widetilde{D}),\theta_{\widetilde{\mathcal{F}}}). By the adjunction, we obtain ψ:E𝒪X(D)φ~(~𝒪X~(D~))=φ~(~)𝒪X(D)\psi:E\otimes\mathcal{O}_{X}(\ast D)\longrightarrow\widetilde{\varphi}_{\ast}(\widetilde{\mathcal{F}}\otimes\mathcal{O}_{\widetilde{X}}(\ast\widetilde{D}))=\widetilde{\varphi}_{\ast}(\widetilde{\mathcal{F}})\otimes\mathcal{O}_{X}(\ast D) compatible with the Higgs fields. Because the restriction ψ|XD\psi_{|X\setminus D} is an isomorphism, we obtain that ψ\psi is an isomorphism of locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-modules. We obtain the coherent 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule φ~()𝒪X(D)E𝒪X(D)\widetilde{\varphi}_{\ast}(\mathcal{L})\otimes\mathcal{O}_{X}(\ast D)\subset E\otimes\mathcal{O}_{X}(\ast D). We obtain the holomorphic subbundle 𝒢:=(φ~()𝒪X(D))EE\mathcal{G}:=\bigl{(}\widetilde{\varphi}_{\ast}(\mathcal{L})\otimes\mathcal{O}_{X}(\ast D)\bigr{)}\cap E\subset E which is preserved by θ\theta. Because EE has no proper Higgs subbundle, we obtain 𝒢=E\mathcal{G}=E. By considering the restriction of the characteristic polynomial to XDX\setminus D, we obtain n(k)=rn(k)=r and thus k=1k=1, i.e., (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple.  

3 Good filtered Higgs bundles with symmetric pairing

3.1 Pairings of filtered bundles

3.1.1 Pairings of locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-modules

Let XX be a Riemann surface. Let DXD\subset X be a discrete subset. Let 𝒪X(D)\mathcal{O}_{X}(\ast D) denote the sheaf of meromorphic functions on XX whose poles are contained in DD.

Let 𝒱\mathcal{V} be a locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-module of finite rank. Let 𝒱\mathcal{V}^{\lor} denote the dual of 𝒱\mathcal{V}, i.e.,

𝒱=om𝒪X(D)(𝒱,𝒪X(D)).\mathcal{V}^{\lor}={\mathcal{H}om}_{\mathcal{O}_{X}(\ast D)}(\mathcal{V},\mathcal{O}_{X}(\ast D)).

The determinant bundle of 𝒱\mathcal{V} is denoted by det(𝒱)\det(\mathcal{V}), i.e., det(𝒱)=rank𝒱𝒱\det(\mathcal{V})=\bigwedge^{\mathop{\rm rank}\nolimits\mathcal{V}}\mathcal{V}. There exists a natural isomorphism det(𝒱)det(𝒱)\det(\mathcal{V}^{\lor})\simeq\det(\mathcal{V})^{\lor}. For a morphism f:𝒱1𝒱2f:\mathcal{V}_{1}\to\mathcal{V}_{2} of locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-modules, f:𝒱2𝒱1f^{\lor}:\mathcal{V}_{2}^{\lor}\to\mathcal{V}_{1}^{\lor} denotes the dual of ff. If rank𝒱1=rank𝒱2\mathop{\rm rank}\nolimits\mathcal{V}_{1}=\mathop{\rm rank}\nolimits\mathcal{V}_{2}, det(f):det(𝒱1)det(𝒱2)\det(f):\det(\mathcal{V}_{1})\to\det(\mathcal{V}_{2}) denotes the induced morphism.

For locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-modules 𝒱i\mathcal{V}_{i} (i=1,2)(i=1,2), a pairing CC of 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} is a morphism C:𝒱1𝒪X(D)𝒱2𝒪X(D)C:\mathcal{V}_{1}\otimes_{\mathcal{O}_{X}(\ast D)}\mathcal{V}_{2}\to\mathcal{O}_{X}(\ast D). It induces a morphism ΨC:𝒱1𝒱2\Psi_{C}:\mathcal{V}_{1}\to\mathcal{V}_{2}^{\lor} by ΨC(u)(v)=C(uv)\Psi_{C}(u)(v)=C(u\otimes v). Let ex:𝒱1𝒱2𝒱2𝒱1\mathop{\rm ex}\nolimits:\mathcal{V}_{1}\otimes\mathcal{V}_{2}\simeq\mathcal{V}_{2}\otimes\mathcal{V}_{1} denote the natural morphism defined by ex(v1v2)=v2v1\mathop{\rm ex}\nolimits(v_{1}\otimes v_{2})=v_{2}\otimes v_{1}. We obtain the pairing Cex:𝒱2𝒱1𝒪X(D)C\circ\mathop{\rm ex}\nolimits:\mathcal{V}_{2}\otimes\mathcal{V}_{1}\to\mathcal{O}_{X}(\ast D). We have ΨCex=ΨC\Psi_{C\circ\mathop{\rm ex}\nolimits}=\Psi_{C}^{\lor}. If rank𝒱1=rank𝒱2\mathop{\rm rank}\nolimits\mathcal{V}_{1}=\mathop{\rm rank}\nolimits\mathcal{V}_{2}, we obtain the induced pairing det(C):det(𝒱1)det(𝒱2)𝒪X(D)\det(C):\det(\mathcal{V}_{1})\otimes\det(\mathcal{V}_{2})\to\mathcal{O}_{X}(\ast D). We have det(ΨC)=Ψdet(C)\det(\Psi_{C})=\Psi_{\det(C)}.

A pairing CC is called non-degenerate if ΨC\Psi_{C} is an isomorphism. It is equivalent to the condition that CexC\circ\mathop{\rm ex}\nolimits is non-degenerate. It is also equivalent to the condition that det(C)\det(C) is non-degenerate. If CC is non-degenerate, we obtain the pairing CC^{\lor} of 𝒱2\mathcal{V}_{2}^{\lor} and 𝒱1\mathcal{V}_{1}^{\lor} defined by C=C(ΨC1ΨCex1)C^{\lor}=C\circ\bigl{(}\Psi_{C}^{-1}\otimes\Psi_{C\circ\mathop{\rm ex}\nolimits}^{-1}\bigr{)}, i.e., the composition of the following morphism:

𝒱2𝒱1ΨC1ΨCex1𝒱1𝒱2C𝒪X(D).\begin{CD}\mathcal{V}_{2}^{\lor}\otimes\mathcal{V}_{1}^{\lor}@>{\Psi_{C}^{-1}\otimes\Psi_{C\circ\mathop{\rm ex}\nolimits}^{-1}}>{}>\mathcal{V}_{1}\otimes\mathcal{V}_{2}@>{C}>{}>\mathcal{O}_{X}(\ast D).\end{CD}

A pairing CC of a locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-module 𝒱\mathcal{V} is a morphism C:𝒱𝒱𝒪X(D)C:\mathcal{V}\otimes\mathcal{V}\to\mathcal{O}_{X}(\ast D), i.e., a pairing of 𝒱\mathcal{V} and 𝒱\mathcal{V}. It is called symmetric if Cex=CC\circ\mathop{\rm ex}\nolimits=C, i.e., C(v1v2)=C(v2v1)C(v_{1}\otimes v_{2})=C(v_{2}\otimes v_{1}) for any local sections viv_{i} of 𝒱\mathcal{V}. We obtain the induced pairing det(C)\det(C) of det(𝒱)\det(\mathcal{V}). If CC is non-degenerate, we obtain the induced pairing CC^{\lor} of 𝒱\mathcal{V}^{\lor}.

3.1.2 Filtered bundles

Let us recall the notion of filtered bundles by following [17, 18]. For any sheaf \mathcal{F} on XX and QXQ\in X, let Q\mathcal{F}_{Q} denote the stalk of \mathcal{F} at QQ. In this subsection, 𝒱\mathcal{V} and 𝒱i\mathcal{V}_{i} (i=1,2)(i=1,2) denote locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-modules of finite rank.

For PDP\in D, a filtered bundle 𝒫(𝒱P)\mathcal{P}_{\ast}(\mathcal{V}_{P}) over 𝒱P\mathcal{V}_{P} is an increasing sequence of free 𝒪X,P\mathcal{O}_{X,P}-submodules 𝒫a(𝒱P)\mathcal{P}_{a}(\mathcal{V}_{P}) (a)(a\in{\mathbb{R}}) of 𝒱P\mathcal{V}_{P} such that (i) 𝒫a(𝒱P)(P)=𝒱P\mathcal{P}_{a}(\mathcal{V}_{P})(\ast P)=\mathcal{V}_{P}, (ii) 𝒫a(𝒱P)=b>a𝒫a(𝒱P)\mathcal{P}_{a}(\mathcal{V}_{P})=\bigcap_{b>a}\mathcal{P}_{a}(\mathcal{V}_{P}) for any aa\in{\mathbb{R}}, (iii) for any aa\in{\mathbb{R}} and nn\in{\mathbb{Z}}, we have 𝒫a+n(𝒱P)=𝒫a(𝒱P)𝒪X,P(nP)\mathcal{P}_{a+n}(\mathcal{V}_{P})=\mathcal{P}_{a}(\mathcal{V}_{P})\otimes\mathcal{O}_{X,P}(n\cdot P). When a tuple of filtered bundles 𝒫(𝒱P)\mathcal{P}_{\ast}(\mathcal{V}_{P}) (PD)(P\in D) is provided, we obtain locally free 𝒪X\mathcal{O}_{X}-submodules 𝒫𝒂(𝒱)𝒱\mathcal{P}_{{\boldsymbol{a}}}(\mathcal{V})\subset\mathcal{V} (𝒂D)({\boldsymbol{a}}\in{\mathbb{R}}^{D}) characterized by the conditions (i) 𝒫𝒂(𝒱)𝒪X𝒪X(D)=𝒱\mathcal{P}_{{\boldsymbol{a}}}(\mathcal{V})\otimes_{\mathcal{O}_{X}}\mathcal{O}_{X}(\ast D)=\mathcal{V}, (ii) 𝒫𝒂(𝒱)P=𝒫a(P)(𝒱P)\mathcal{P}_{{\boldsymbol{a}}}(\mathcal{V})_{P}=\mathcal{P}_{a(P)}(\mathcal{V}_{P}), where a(P)a(P) denotes the PP-component of 𝒂D{\boldsymbol{a}}\in{\mathbb{R}}^{D}. Such a tuple 𝒫(𝒱)=(𝒫𝒂𝒱|𝒂D)\mathcal{P}_{\ast}(\mathcal{V})=(\mathcal{P}_{{\boldsymbol{a}}}\mathcal{V}\,|\,{\boldsymbol{a}}\in{\mathbb{R}}^{D}) is called a filtered bundle over 𝒱\mathcal{V}. We also say that 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) is a filtered bundle on (X,D)(X,D).

Let 𝒱𝒱\mathcal{V}^{\prime}\subset\mathcal{V} be a locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule of 𝒱\mathcal{V}. When a filtered bundle 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) over 𝒱\mathcal{V} is provided, by setting 𝒫a(𝒱P):=𝒱P𝒫a(𝒱)\mathcal{P}_{a}(\mathcal{V}^{\prime}_{P}):=\mathcal{V}^{\prime}_{P}\cap\mathcal{P}_{a}(\mathcal{V}), we obtain the induced filtered bundle 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}^{\prime}) over 𝒱\mathcal{V}^{\prime}.

When filtered bundles 𝒫(𝒱i)\mathcal{P}_{\ast}(\mathcal{V}_{i}) (i=1,2)(i=1,2) over 𝒱i\mathcal{V}_{i} are provided, a morphism f:𝒫(𝒱1)𝒫(𝒱2)f:\mathcal{P}_{\ast}(\mathcal{V}_{1})\to\mathcal{P}_{\ast}(\mathcal{V}_{2}) is an 𝒪X(D)\mathcal{O}_{X}(\ast D)-homomorphism f:𝒱1𝒱2f:\mathcal{V}_{1}\to\mathcal{V}_{2} such that f(𝒫a(𝒱1,P))𝒫a(𝒱2,P)f(\mathcal{P}_{a}(\mathcal{V}_{1,P}))\subset\mathcal{P}_{a}(\mathcal{V}_{2,P}) for any PDP\in D and aa\in{\mathbb{R}}.

When filtered bundles 𝒫(𝒱i)\mathcal{P}_{\ast}(\mathcal{V}_{i}) over 𝒱i\mathcal{V}_{i} are provided, for any aa\in{\mathbb{R}} and PDP\in D we define

𝒫a((𝒱1𝒱2)P)=𝒫a𝒱1,P𝒫a𝒱2,P,𝒫a((𝒱1𝒱2)P)=c1+c2a𝒫c1(𝒱1,P)𝒪c2(𝒱2,P).\mathcal{P}_{a}\bigl{(}(\mathcal{V}_{1}\oplus\mathcal{V}_{2})_{P}\bigr{)}=\mathcal{P}_{a}\mathcal{V}_{1,P}\oplus\mathcal{P}_{a}\mathcal{V}_{2,P},\quad\mathcal{P}_{a}\bigl{(}(\mathcal{V}_{1}\otimes\mathcal{V}_{2})_{P}\bigr{)}=\sum_{c_{1}+c_{2}\leq a}\mathcal{P}_{c_{1}}(\mathcal{V}_{1,P})\otimes\mathcal{O}_{c_{2}}(\mathcal{V}_{2,P}).

We also define

𝒫a(om𝒪X(D)(𝒱1,𝒱2)P)={fom𝒪X(D)P(𝒱1,P,𝒱2,P)|f(𝒫b(𝒱1,P))𝒫b+a(𝒱2,P)b}.\mathcal{P}_{a}\bigl{(}{\mathcal{H}om}_{\mathcal{O}_{X}(\ast D)}(\mathcal{V}_{1},\mathcal{V}_{2})_{P}\bigr{)}=\bigl{\{}f\in{\mathcal{H}om}_{\mathcal{O}_{X}(\ast D)_{P}}(\mathcal{V}_{1,P},\mathcal{V}_{2,P})\,\big{|}\,f(\mathcal{P}_{b}(\mathcal{V}_{1,P}))\subset\mathcal{P}_{b+a}(\mathcal{V}_{2,P})\,\,\forall b\in{\mathbb{R}}\bigr{\}}.

In this way, we obtain filtered bundles 𝒫(𝒱1𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{1}\oplus\mathcal{V}_{2}) over 𝒱1𝒱2\mathcal{V}_{1}\oplus\mathcal{V}_{2}, 𝒫(𝒱1𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{1}\otimes\mathcal{V}_{2}) over 𝒱1𝒱2\mathcal{V}_{1}\otimes\mathcal{V}_{2}, and 𝒫om(𝒱1,𝒱2)\mathcal{P}_{\ast}{\mathcal{H}om}(\mathcal{V}_{1},\mathcal{V}_{2}) over om(𝒱1,𝒱2){\mathcal{H}om}(\mathcal{V}_{1},\mathcal{V}_{2}). The filtered bundles are denoted by 𝒫(𝒱1)𝒫(𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{1})\oplus\mathcal{P}_{\ast}(\mathcal{V}_{2}), 𝒫(𝒱1)𝒫(𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{1})\otimes\mathcal{P}_{\ast}(\mathcal{V}_{2}), and om(𝒫𝒱1,𝒫𝒱2){\mathcal{H}om}(\mathcal{P}_{\ast}\mathcal{V}_{1},\mathcal{P}_{\ast}\mathcal{V}_{2}), respectively.

Remark 3.1

Even if 𝒱=𝒱𝒱′′\mathcal{V}=\mathcal{V}^{\prime}\oplus\mathcal{V}^{\prime\prime} holds as locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-modules, 𝒫(𝒱)=𝒫(𝒱)𝒫(𝒱′′)\mathcal{P}_{\ast}(\mathcal{V})=\mathcal{P}_{\ast}(\mathcal{V}^{\prime})\oplus\mathcal{P}_{\ast}(\mathcal{V}^{\prime\prime}) does not necessarily hold for the induced filtered bundles 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}^{\prime}) and 𝒫(𝒱′′)\mathcal{P}_{\ast}(\mathcal{V}^{\prime\prime}). If 𝒫(𝒱)=𝒫(𝒱)𝒫(𝒱′′)\mathcal{P}_{\ast}(\mathcal{V})=\mathcal{P}_{\ast}(\mathcal{V}^{\prime})\oplus\mathcal{P}_{\ast}(\mathcal{V}^{\prime\prime}) holds, we say that the filtered bundle 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) is compatible with the decomposition 𝒱=𝒱𝒱′′\mathcal{V}=\mathcal{V}^{\prime}\oplus\mathcal{V}^{\prime\prime}.  

We set r=rank𝒱r=\mathop{\rm rank}\nolimits\mathcal{V}. From a filtered bundle 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} over 𝒱\mathcal{V}, we obtain a filtered bundle 𝒫(𝒱)r\mathcal{P}_{\ast}(\mathcal{V})^{\otimes r} over 𝒱r\mathcal{V}^{\otimes r}. There exists the decomposition 𝒱r=det(𝒱)(𝒱r)\mathcal{V}^{\otimes r}=\det(\mathcal{V})\oplus(\mathcal{V}^{\otimes r})^{\bot} in a way compatible with the natural action of the rr-th symmetric group 𝔖r\mathfrak{S}_{r}. We obtain the filtered bundles 𝒫(det(𝒱))\mathcal{P}_{\ast}(\det(\mathcal{V})) and 𝒫((𝒱r))\mathcal{P}_{\ast}((\mathcal{V}^{\otimes\,r})^{\bot}) over det(𝒱)\det(\mathcal{V}) and (𝒱r)(\mathcal{V}^{\otimes\,r})^{\bot}, respectively, for which 𝒫(𝒱)r=𝒫(det(𝒱))𝒫((𝒱r))\mathcal{P}_{\ast}(\mathcal{V})^{\otimes r}=\mathcal{P}_{\ast}(\det(\mathcal{V}))\oplus\mathcal{P}_{\ast}((\mathcal{V}^{\otimes\,r})^{\bot}) holds.

Let 𝒫(0)(𝒪X(D))\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)) denote the canonical filtered bundle over 𝒪X(D)\mathcal{O}_{X}(\ast D), i.e.,

𝒫𝒂(0)(𝒪X(D))=𝒪X(PD[a(P)]P),\mathcal{P}^{(0)}_{{\boldsymbol{a}}}(\mathcal{O}_{X}(\ast D))=\mathcal{O}_{X}\Bigl{(}\sum_{P\in D}[a(P)]P\Bigr{)},

where we set [c]:=max{n|nc}[c]:=\max\{n\in{\mathbb{Z}}\,|\,n\leq c\} for any cc\in{\mathbb{R}}. When a filtered bundle 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) over 𝒱\mathcal{V} is provided, let 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}^{\lor}) denote om(𝒫𝒱,𝒫(0)(𝒪X(D))){\mathcal{H}om}(\mathcal{P}_{\ast}\mathcal{V},\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D))).

3.1.3 Pairings of filtered bundles

Let 𝒫(𝒱i)\mathcal{P}_{\ast}(\mathcal{V}_{i}) (i=1,2)(i=1,2) be filtered bundles on (X,D)(X,D). A pairing CC of 𝒫(𝒱1)\mathcal{P}_{\ast}(\mathcal{V}_{1}) and 𝒫(𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{2}) is a morphism

C:𝒫(𝒱1)𝒫(𝒱2)𝒫(0)(𝒪X(D)).C:\mathcal{P}_{\ast}(\mathcal{V}_{1})\otimes\mathcal{P}_{\ast}(\mathcal{V}_{2})\longrightarrow\mathcal{P}_{\ast}^{(0)}(\mathcal{O}_{X}(\ast D)).

We obtain the pairing CexC\circ\mathop{\rm ex}\nolimits of 𝒫(𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{2}) and 𝒫(𝒱1)\mathcal{P}_{\ast}(\mathcal{V}_{1}). If rank𝒱1=rank𝒱2\mathop{\rm rank}\nolimits\mathcal{V}_{1}=\mathop{\rm rank}\nolimits\mathcal{V}_{2}, we obtain

det(C):det(𝒫𝒱1)det(𝒫𝒱2)𝒫(0)(𝒪X(D)).\det(C):\det(\mathcal{P}_{\ast}\mathcal{V}_{1})\otimes\det(\mathcal{P}_{\ast}\mathcal{V}_{2})\longrightarrow\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). (9)

From a pairing CC of 𝒫(𝒱1)\mathcal{P}_{\ast}(\mathcal{V}_{1}) and 𝒫(𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{2}), we obtain the following morphism of filtered bundles:

ΨC:𝒫(𝒱1)𝒫(𝒱2).\Psi_{C}:\mathcal{P}_{\ast}(\mathcal{V}_{1})\longrightarrow\mathcal{P}_{\ast}(\mathcal{V}_{2}^{\lor}). (10)
Definition 3.2

CC is called perfect if the morphism (10) is an isomorphism of filtered bundles.  

Note that if rank𝒱=1\mathop{\rm rank}\nolimits\mathcal{V}=1, CC is perfect if and only if (9) is an isomorphism.

Lemma 3.3

CC is perfect if and only if the following induced morphism is an isomorphism:

det(C):det(𝒫𝒱1)det(𝒫𝒱2)𝒫(0)(𝒪X(D)).\det(C):\det(\mathcal{P}_{\ast}\mathcal{V}_{1})\otimes\det(\mathcal{P}_{\ast}\mathcal{V}_{2})\longrightarrow\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). (11)

Proof   The morphism (10) is an isomorphism if and only if the induced morphism

det(ΨC):det(𝒫𝒱1)det(𝒫𝒱2)\det(\Psi_{C}):\det(\mathcal{P}_{\ast}\mathcal{V}_{1})\longrightarrow\det(\mathcal{P}_{\ast}\mathcal{V}_{2}^{\lor})

is an isomorphism, which is equivalent to the condition that (11) is an isomorphism.  

Let 𝒱i𝒱i\mathcal{V}_{i}^{\prime}\subset\mathcal{V}_{i} be locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodules. For simplicity, we also assume that 𝒱i\mathcal{V}_{i}^{\prime} are saturated, i.e., 𝒱i/𝒱i\mathcal{V}_{i}/\mathcal{V}_{i}^{\prime} are also locally free. From a pairing CC of 𝒫𝒱1\mathcal{P}_{\ast}\mathcal{V}_{1} and 𝒫(𝒱2)\mathcal{P}_{\ast}(\mathcal{V}_{2}), we obtain the induced pairing CC^{\prime} of 𝒫𝒱1\mathcal{P}_{\ast}\mathcal{V}_{1}^{\prime} and 𝒫𝒱2\mathcal{P}_{\ast}\mathcal{V}_{2}^{\prime}. There exist the following natural morphisms:

𝒱1i1𝒱1ΨC𝒱2i2(𝒱2).\begin{CD}\mathcal{V}_{1}^{\prime}@>{i_{1}}>{}>\mathcal{V}_{1}@>{\Psi_{C}}>{}>\mathcal{V}_{2}^{\lor}@>{i_{2}^{\lor}}>{}>(\mathcal{V}_{2}^{\prime})^{\lor}.\end{CD} (12)

Here, i1i_{1} denotes the natural inclusion, and i2i_{2}^{\lor} denotes the dual of the natural inclusion i2:𝒱2𝒱2i_{2}:\mathcal{V}_{2}^{\prime}\to\mathcal{V}_{2}. Note that ΨC=i2ΨCi1\Psi_{C^{\prime}}=i_{2}^{\lor}\circ\Psi_{C}\circ i_{1}. Let 𝒰1𝒱1\mathcal{U}_{1}\subset\mathcal{V}_{1} denote the kernel of i2ΨCi_{2}^{\lor}\circ\Psi_{C}. We have the induced filtered bundle 𝒫𝒰1\mathcal{P}_{\ast}\mathcal{U}_{1} over 𝒰1\mathcal{U}_{1}. The following lemma is obvious.

Lemma 3.4

If CC and CC^{\prime} are perfect, we have the decomposition of the filtered bundles 𝒫𝒱1=𝒫𝒱1𝒫𝒰1\mathcal{P}_{\ast}\mathcal{V}_{1}=\mathcal{P}_{\ast}\mathcal{V}_{1}^{\prime}\oplus\mathcal{P}_{\ast}\mathcal{U}_{1}.  

3.1.4 Symmetric pairings of filtered bundles

Let CC be a symmetric pairing of a filtered bundle 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} on (X,D)(X,D). We have the induced pairing

det(C):det(𝒫𝒱)det(𝒫𝒱)𝒫(0)(𝒪X(D)).\det(C):\det(\mathcal{P}_{\ast}\mathcal{V})\otimes\det(\mathcal{P}_{\ast}\mathcal{V})\to\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). (13)

We obtain the following lemma as a special case of Lemma 3.3.

Lemma 3.5

CC is perfect if and only if (13) is an isomorphism.  

Let 𝒱𝒱\mathcal{V}^{\prime}\subset\mathcal{V} be a saturated locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule. Let (𝒱)C(\mathcal{V}^{\prime})^{\bot\,C} denote the kernel of the composition of the following morphisms:

𝒱ΨC𝒱i(𝒱),\begin{CD}\mathcal{V}@>{\Psi_{C}}>{}>\mathcal{V}^{\lor}@>{i^{\lor}}>{}>(\mathcal{V}^{\prime})^{\lor},\end{CD}

where ii^{\lor} denote the dual of the inclusion i:𝒱𝒱i:\mathcal{V}^{\prime}\to\mathcal{V}. We have the filtered bundle 𝒫(𝒱)C\mathcal{P}_{\ast}(\mathcal{V}^{\prime})^{\bot\,C} over (𝒱)C(\mathcal{V}^{\prime})^{\bot\,C}. Let CC^{\prime} denote the induced symmetric pairing of 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}^{\prime}). The following lemma is a special case of Lemma 3.4.

Lemma 3.6

If CC and CC^{\prime} are perfect, we obtain the decomposition 𝒫𝒱=𝒫𝒱𝒫(𝒱)C\mathcal{P}_{\ast}\mathcal{V}=\mathcal{P}_{\ast}\mathcal{V}^{\prime}\oplus\mathcal{P}_{\ast}(\mathcal{V}^{\prime})^{\bot\,C}.  

Corollary 3.7

Suppose that (13) is an isomorphism. We also assume that

det(C):det(𝒫𝒱)det(𝒫𝒱)𝒫(0)(𝒪X(D))\det(C^{\prime}):\det(\mathcal{P}_{\ast}\mathcal{V}^{\prime})\otimes\det(\mathcal{P}_{\ast}\mathcal{V}^{\prime})\to\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)) (14)

is an isomorphism. Then, we have the decomposition of the filtered bundles 𝒫𝒱=𝒫𝒱𝒫(𝒱)C\mathcal{P}_{\ast}\mathcal{V}=\mathcal{P}_{\ast}\mathcal{V}^{\prime}\oplus\mathcal{P}_{\ast}(\mathcal{V}^{\prime})^{\bot\,C}.  

3.1.5 Compact case

We assume that XX is compact. Recall that for any filtered bundle 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} on (X,D)(X,D), we obtain deg(𝒫𝒱)\deg(\mathcal{P}_{\ast}\mathcal{V})\in{\mathbb{R}} as follows (see [17, 18] and also [14]):

deg(𝒫𝒱)=deg(𝒫𝒂𝒱)PDa(P)1<aa(P)adimGra𝒫(𝒱P).\deg(\mathcal{P}_{\ast}\mathcal{V})=\deg(\mathcal{P}_{{\boldsymbol{a}}}\mathcal{V})-\sum_{P\in D}\sum_{a(P)-1<a\leq a(P)}a\cdot\dim_{{\mathbb{C}}}\mathop{\rm Gr}\nolimits^{\mathcal{P}}_{a}(\mathcal{V}_{P}).

Here, we set Gra𝒫(𝒱P):=𝒫a(𝒱P)/b<a𝒫b(𝒱P)\mathop{\rm Gr}\nolimits^{\mathcal{P}}_{a}(\mathcal{V}_{P}):=\mathcal{P}_{a}(\mathcal{V}_{P})\Big{/}\sum_{b<a}\mathcal{P}_{b}(\mathcal{V}_{P}). It is independent of the choice of 𝒂D{\boldsymbol{a}}\in{\mathbb{R}}^{D}. Note that deg(𝒫𝒱)=deg(det(𝒫𝒱))\deg(\mathcal{P}_{\ast}\mathcal{V})=\deg(\det(\mathcal{P}_{\ast}\mathcal{V})).

Lemma 3.8

Let 𝒫𝒱i\mathcal{P}_{\ast}\mathcal{V}_{i} (i=1,2)(i=1,2) be filtered bundles of rank 11 on (X,D)(X,D). If there exists a non-zero morphism f:𝒫𝒱1𝒫𝒱2f:\mathcal{P}_{\ast}\mathcal{V}_{1}\to\mathcal{P}_{\ast}\mathcal{V}_{2}, then deg(𝒫𝒱1)deg(𝒫𝒱2)\deg(\mathcal{P}_{\ast}\mathcal{V}_{1})\leq\deg(\mathcal{P}_{\ast}\mathcal{V}_{2}). If moreover deg(𝒫𝒱1)=deg(𝒫𝒱2)\deg(\mathcal{P}_{\ast}\mathcal{V}_{1})=\deg(\mathcal{P}_{\ast}\mathcal{V}_{2}), then ff is an isomorphism.

Proof   Though this is well known, we include a sketch of the proof for the convenience of the readers. By the morphism ff, we can regard 𝒱1\mathcal{V}_{1} as an 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule of 𝒱2\mathcal{V}_{2}. There exists a finite subset ZXDZ\subset X\setminus D and function m:Z>0m:Z\to{\mathbb{Z}}_{>0} such that 𝒱1=𝒱2(QZm(Q)Q)\mathcal{V}_{1}=\mathcal{V}_{2}(-\sum_{Q\in Z}m(Q)Q). For each PDP\in D, we take a non-zero element vP𝒱1,P=𝒱2,Pv_{P}\in\mathcal{V}_{1,P}=\mathcal{V}_{2,P}. We obtain the numbers b(P,i):=min{b|vP𝒫b(𝒱i,P)}b(P,i):=\min\{b\in{\mathbb{R}}\,|\,v_{P}\in\mathcal{P}_{b}(\mathcal{V}_{i,P})\}. Let 𝒱i𝒱i\mathcal{V}^{\prime}_{i}\subset\mathcal{V}_{i} be the locally free 𝒪X\mathcal{O}_{X}-submodules determined by the conditions (i) 𝒱i𝒪X(D)=𝒱i\mathcal{V}^{\prime}_{i}\otimes\mathcal{O}_{X}(\ast D)=\mathcal{V}_{i}, (ii) 𝒱i,P=𝒪X,PvP\mathcal{V}^{\prime}_{i,P}=\mathcal{O}_{X,P}v_{P}. By the definition, we obtain

deg(𝒫𝒱i)=deg(𝒱i)PDb(P,i).\deg(\mathcal{P}_{\ast}\mathcal{V}_{i})=\deg(\mathcal{V}_{i}^{\prime})-\sum_{P\in D}b(P,i).

Because 𝒱1=𝒱2(m(Q)Q)\mathcal{V}_{1}^{\prime}=\mathcal{V}_{2}^{\prime}(-\sum m(Q)Q) and b(P,1)b(P,2)b(P,1)\geq b(P,2) (PD)(P\in D), we obtain

deg(𝒫𝒱1)=deg(𝒱1)PDb(P,1)=deg(𝒱2)QZm(Q)PDb(P,1)deg(𝒱2)PDb(P,2)=deg(𝒫𝒱2).\deg(\mathcal{P}_{\ast}\mathcal{V}_{1})=\deg(\mathcal{V}_{1}^{\prime})-\sum_{P\in D}b(P,1)=\deg(\mathcal{V}_{2}^{\prime})-\sum_{Q\in Z}m(Q)-\sum_{P\in D}b(P,1)\leq\deg(\mathcal{V}_{2}^{\prime})-\sum_{P\in D}b(P,2)=\deg(\mathcal{P}_{\ast}\mathcal{V}_{2}).

If deg(𝒫𝒱1)=deg(𝒫𝒱2)\deg(\mathcal{P}_{\ast}\mathcal{V}_{1})=\deg(\mathcal{P}_{\ast}\mathcal{V}_{2}), we obtain Z=Z=\emptyset and b(P,1)=b(P,2)b(P,1)=b(P,2) (PD)(P\in D), and hence 𝒫𝒱1=𝒫𝒱2\mathcal{P}_{\ast}\mathcal{V}_{1}=\mathcal{P}_{\ast}\mathcal{V}_{2}.  

Proposition 3.9

Let CC be a symmetric pairing of a filtered bundle 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} on (X,D)(X,D). Then, we have either (i) det(C)=0\det(C)=0, or (ii) det(𝒫𝒱)0\det(\mathcal{P}_{\ast}\mathcal{V})\leq 0. If det(C)0\det(C)\neq 0 and deg(𝒫𝒱)=0\deg(\mathcal{P}_{\ast}\mathcal{V})=0, then CC is perfect.

Proof   We obtain

det(C):det(𝒫𝒱)det(𝒫𝒱)𝒫(0)(𝒪X(D)).\det(C):\det(\mathcal{P}_{\ast}\mathcal{V})\otimes\det(\mathcal{P}_{\ast}\mathcal{V})\to\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). (15)

By Lemma 3.8, if det(C)0\det(C)\neq 0, we obtain

2deg(𝒫𝒱)=deg(det(𝒫𝒱)det(𝒫𝒱))deg(𝒫(0)(𝒪X(D)))=0.2\deg(\mathcal{P}_{\ast}\mathcal{V})=\deg\bigl{(}\det(\mathcal{P}_{\ast}\mathcal{V})\otimes\det(\mathcal{P}_{\ast}\mathcal{V})\bigr{)}\leq\deg(\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)))=0.

If moreover deg(𝒫𝒱)=0\deg(\mathcal{P}_{\ast}\mathcal{V})=0 holds, then (15) is an isomorphism. Hence, CC is perfect by Lemma 3.5.  

As a complement, we remark the following.

Lemma 3.10

If a filtered bundle 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} on (X,D)(X,D) has a perfect symmetric pairing, then we obtain deg(𝒫𝒱)=0\deg(\mathcal{P}_{\ast}\mathcal{V})=0.

Proof   Because det(𝒫𝒱)det(𝒫𝒱)𝒫(0)(𝒪X(D))\det(\mathcal{P}_{\ast}\mathcal{V})\otimes\det(\mathcal{P}_{\ast}\mathcal{V})\simeq\mathcal{P}^{(0)}(\mathcal{O}_{X}(\ast D)), we obtain deg(𝒫𝒱)=deg(det(𝒫𝒱))=0\deg(\mathcal{P}_{\ast}\mathcal{V})=\deg(\det(\mathcal{P}_{\ast}\mathcal{V}))=0.  

3.1.6 Harmonic metrics in the rank one case

We continue to assume that XX is compact. Let h(0)h^{(0)} denote the harmonic metric of 𝒪XD\mathcal{O}_{X\setminus D} defined by h(0)(1,1)=1h^{(0)}(1,1)=1. It is adapted to 𝒫(0)(𝒪X(D))\mathcal{P}_{\ast}^{(0)}(\mathcal{O}_{X}(\ast D)), i.e., 𝒫h(0)(𝒪XD)=𝒫(0)(𝒪X(D))\mathcal{P}^{h^{(0)}}_{\ast}(\mathcal{O}_{X\setminus D})=\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)).

Let 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} be a filtered bundle on (X,D)(X,D) of rank 11 such that deg(𝒫𝒱)=0\deg(\mathcal{P}_{\ast}\mathcal{V})=0 equipped with a non-zero pairing:

C:𝒫(𝒱)𝒫(𝒱)𝒫(0)(𝒪X(D)).C:\mathcal{P}_{\ast}(\mathcal{V})\otimes\mathcal{P}_{\ast}(\mathcal{V})\longrightarrow\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). (16)

Note that (16) is an isomorphism by Proposition 3.9.

Lemma 3.11

There uniquely exists a Hermitian metric hh of 𝒱|XD\mathcal{V}_{|X\setminus D} such that (i) hh is harmonic, i.e., the Chern connection of hh is flat, (ii) hh is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}, (iii) hh=h(0)h\otimes h=h^{(0)} under the isomorphism C:𝒫𝒱𝒫𝒱𝒫(0)(𝒪X(D))C:\mathcal{P}_{\ast}\mathcal{V}\otimes\mathcal{P}_{\ast}\mathcal{V}\simeq\mathcal{P}_{\ast}^{(0)}(\mathcal{O}_{X}(\ast D)). Note that the condition (iii) means that hh is compatible with C|XDC_{|X\setminus D}.

Proof   It is well known that there exists a harmonic metric h1h_{1} of 𝒪XD\mathcal{O}_{X\setminus D} satisfying the conditions (i) and (ii). We obtain the metric h1h1h_{1}\otimes h_{1} of 𝒪XD\mathcal{O}_{X\setminus D} by the isomorphism (16), which is adapted to 𝒫(0)(𝒪X(D))\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). By the uniqueness of harmonic metrics, there exists c>0c>0 such that h1h1=ch(0)h_{1}\otimes h_{1}=ch^{(0)}. Then, h=c1/2h1h=c^{-1/2}h_{1} satisfies the desired conditions.

 

3.2 Symmetric pairings of good filtered Higgs bundles

3.2.1 Preliminary

Let XX be a Riemann surface. For each PDP\in D, let (XP,zP)(X_{P},z_{P}) be a holomorphic coordinate neighbourhood around PP such that zP(P)=0z_{P}(P)=0. By the coordinate, we may regard XPX_{P} as a neighbourhood of 0 in {\mathbb{C}}. For a positive integer ee, let φe:\varphi_{e}:{\mathbb{C}}\to{\mathbb{C}} be defined by φe(ζ)=ζe\varphi_{e}(\zeta)=\zeta^{e}. Let XP(e)=φe1(XP)X^{(e)}_{P}=\varphi_{e}^{-1}(X_{P}). Let φP,e:XP(e)XP\varphi_{P,e}:X^{(e)}_{P}\to X_{P} denote the induced morphism. Let P(e)XP(e)P^{(e)}\in X^{(e)}_{P} denote the inverse image of PP. On XP(e)X^{(e)}_{P}, we choose a holomorphic function zP,ez_{P,e} such that zP,ee=φP,e(zP)z_{P,e}^{e}=\varphi_{P,e}^{\ast}(z_{P}). Let GalP(e)\mathop{\rm Gal}\nolimits^{(e)}_{P} be the Galois group of the ramified covering XP(e)XPX^{(e)}_{P}\to X_{P}.

3.2.2 Meromorphic Higgs bundles

Let DXD\subset X be a discrete subset. Let 𝒱\mathcal{V} be a locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-module of rank rr. A Higgs field of 𝒱\mathcal{V} is a morphism θ:𝒱𝒱ΩX1\theta:\mathcal{V}\to\mathcal{V}\otimes\Omega^{1}_{X}. Such a tuple (𝒱,θ)(\mathcal{V},\theta) is called a meromorphic Higgs bundle on (X,D)(X,D). We obtain the Higgs field θ\theta^{\lor} of 𝒱\mathcal{V}^{\lor} as the dual of θ\theta. We also obtain the Higgs field tr(θ)\mathop{\rm tr}\nolimits(\theta) of det(𝒱)\det(\mathcal{V}). A morphism g:(𝒱1,θ1)(𝒱2,θ2)g:(\mathcal{V}_{1},\theta_{1})\to(\mathcal{V}_{2},\theta_{2}) of a meromorphic Higgs bundles is an 𝒪X(D)\mathcal{O}_{X}(\ast D)-homomorphism g:𝒱1𝒱2g:\mathcal{V}_{1}\to\mathcal{V}_{2} such that θ2g=gθ1\theta_{2}\circ g=g\circ\theta_{1}.

Let (𝒱,θ)(\mathcal{V},\theta) be a meromorphic Higgs bundle on (X,D)(X,D). Let PDP\in D. We obtain the endomorphism fPf_{P} by θ=fPdzP/zP\theta=f_{P}\,dz_{P}/z_{P} around PP. We say that (𝒱,θ)(\mathcal{V},\theta) is regular at PP if there exists a free 𝒪X,P\mathcal{O}_{X,P}-submodule 𝒱P\mathcal{L}\subset\mathcal{V}_{P} such that (i) (P)=𝒱P\mathcal{L}(\ast P)=\mathcal{V}_{P}, (ii) θ\theta is logarithmic with respect to \mathcal{L}, i.e., θ()zP1ΩX,P1\theta(\mathcal{L})\subset z_{P}^{-1}\mathcal{L}\otimes\Omega^{1}_{X,P}. The second condition is equivalent to that fP()f_{P}(\mathcal{L})\subset\mathcal{L}. Note that we have the characteristic polynomial det(Tid𝒱fP)=Tr+j=0r1aj(zP)Tj\det(T\mathop{\rm id}\nolimits_{\mathcal{V}}-f_{P})=T^{r}+\sum_{j=0}^{r-1}a_{j}(z_{P})T^{j}.

Lemma 3.12

(𝒱,θ)(\mathcal{V},\theta) is regular at PP if and only if aja_{j} are holomorphic at PP.

Proof   The “only if” part is obvious. Suppose that aja_{j} are holomorphic at PP. We choose a non-zero v𝒱Pv\in\mathcal{V}_{P}. We consider the lattice \mathcal{L} generated by fj(v)f^{j}(v) (j=0,,r1)(j=0,\ldots,r-1). Because fr+j=0r1aj(zP)fj=0f^{r}+\sum_{j=0}^{r-1}a_{j}(z_{P})f^{j}=0, we obtain fr(v)f^{r}(v)\in\mathcal{L}. Hence, f()f(\mathcal{L})\subset\mathcal{L}.  

The following lemma is well known and easy to prove.

Lemma 3.13

If (𝒱,θ)(\mathcal{V},\theta) is regular at PP, there exists a decomposition

(𝒱P,θ)=α(𝒱P,α,θP,α)(\mathcal{V}_{P},\theta)=\bigoplus_{\alpha\in{\mathbb{C}}}(\mathcal{V}_{P,\alpha},\theta_{P,\alpha}) (17)

such that the following holds.

  • Let fP,αf_{P,\alpha} be the endomorphism satisfying θP,ααid𝒱P,αdzP/zP=fP,αdzP/zP\theta_{P,\alpha}-\alpha\,\mathop{\rm id}\nolimits_{\mathcal{V}_{P,\alpha}}dz_{P}/z_{P}=f_{P,\alpha}dz_{P}/z_{P}. Let det(TidfP,α)=aα,j(zP)Tj\det(T\mathop{\rm id}\nolimits-f_{P,\alpha})=\sum a_{\alpha,j}(z_{P})T^{j} denote the characteristic polynomial of fP,αf_{P,\alpha}. Then, aα,ja_{\alpha,j} are holomorphic at PP, and aα,j(P)=0a_{\alpha,j}(P)=0 unless j=rank𝒱P,αj=\mathop{\rm rank}\nolimits\mathcal{V}_{P,\alpha}.  

We recall the following lemma, which is an analogue of the Hukuhara-Levelt-Turrittin theorem for meromorphic flat bundles, but easier to prove by using only standard arguments in linear algebra.

Lemma 3.14

There exist e>0e\in{\mathbb{Z}}_{>0} and a decomposition

(φP,e(𝒱)P(e),φP,e(θ))=𝔞zP,e1[zP,e1](𝒱P,𝔞(e),θP,𝔞(e))(\varphi_{P,e}^{\ast}(\mathcal{V})_{P^{(e)}},\varphi_{P,e}^{\ast}(\theta))=\bigoplus_{\mathfrak{a}\in z_{P,e}^{-1}{\mathbb{C}}[z_{P,e}^{-1}]}(\mathcal{V}^{(e)}_{P,\mathfrak{a}},\theta^{(e)}_{P,\mathfrak{a}}) (18)

such that (𝒱P,𝔞(e),θP,𝔞(e)d𝔞id)(\mathcal{V}^{(e)}_{P,\mathfrak{a}},\theta^{(e)}_{P,\mathfrak{a}}-d\mathfrak{a}\mathop{\rm id}\nolimits) are regular. Indeed, ee divides r!r!.  

3.2.3 Good filtered Higgs bundles

Let (𝒱,θ)(\mathcal{V},\theta) be a meromorphic Higgs bundle on (X,D)(X,D). Let 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} be a filtered bundle over 𝒱\mathcal{V}. A filtered Higgs bundle (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) is called regular at PDP\in D if θ\theta is logarithmic with respect to each 𝒫a(𝒱P)\mathcal{P}_{a}(\mathcal{V}_{P}) (a)(a\in{\mathbb{R}}). If (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) is regular at PP, the decomposition (17) is compatible with the filtration, i.e., we obtain the decomposition of filtered bundles at PP for each aa\in{\mathbb{R}}:

𝒫a(𝒱P)=α𝒫a(𝒱P,α).\mathcal{P}_{a}(\mathcal{V}_{P})=\bigoplus_{\alpha\in{\mathbb{C}}}\mathcal{P}_{a}(\mathcal{V}_{P,\alpha}).

Recall that for any e>0e\in{\mathbb{Z}}_{>0}, we obtain the filtered bundle 𝒫(φP,e(𝒱)P(e))\mathcal{P}_{\ast}\bigl{(}\varphi_{P,e}^{\ast}(\mathcal{V})_{P^{(e)}}\bigr{)} over φP,e(𝒱)P(e)\varphi_{P,e}^{\ast}(\mathcal{V})_{P^{(e)}} defined as follows:

𝒫a(φP,e(𝒱)P(e))=n+ebazP,enφP,e(𝒫b(𝒱P))\mathcal{P}_{a}\Bigl{(}\varphi_{P,e}^{\ast}(\mathcal{V})_{P^{(e)}}\Bigr{)}=\sum_{n+eb\leq a}z_{P,e}^{-n}\varphi_{P,e}^{\ast}\bigl{(}\mathcal{P}_{b}(\mathcal{V}_{P})\bigr{)}

A filtered Higgs bundle (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) is called good at PDP\in D if the decomposition (18) is compatible with the filtration, i.e.,

𝒫a(φP,e(𝒱)P(e))=𝒫a(𝒱P,𝔞(e)),\mathcal{P}_{a}\Bigl{(}\varphi_{P,e}^{\ast}(\mathcal{V})_{P^{(e)}}\Bigr{)}=\bigoplus\mathcal{P}_{a}(\mathcal{V}^{(e)}_{P,\mathfrak{a}}),

and moreover θP,𝔞d𝔞id\theta_{P,\mathfrak{a}}-d\mathfrak{a}\mathop{\rm id}\nolimits are logarithmic with respect to 𝒫a(𝒱P,𝔞(e))\mathcal{P}_{a}(\mathcal{V}^{(e)}_{P,\mathfrak{a}}) for any aa\in{\mathbb{R}}.

3.2.4 Symmetric pairings of good filtered Higgs bundles

Definition 3.15

A symmetric pairing CC of a good filtered Higgs bundle (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) (resp. a meromorphic Higgs bundle (𝒱,θ)(\mathcal{V},\theta)) is a symmetric pairing CC of 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} (resp. 𝒱\mathcal{V}) such that C(θid)=C(idθ)C(\theta\otimes\mathop{\rm id}\nolimits)=C(\mathop{\rm id}\nolimits\otimes\theta).  

In case, we obtain the induced morphism ΨC:(𝒫𝒱,θ)(𝒫𝒱,θ)\Psi_{C}:(\mathcal{P}_{\ast}\mathcal{V},\theta)\to(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}) (resp. ΨC:(𝒱,θ)(𝒱,θ)\Psi_{C}:(\mathcal{V},\theta)\to(\mathcal{V}^{\lor},\theta^{\lor})). We also obtain a symmetric pairing det(C)\det(C) of (𝒫𝒱,trθ)(\mathcal{P}_{\ast}\mathcal{V},\mathop{\rm tr}\nolimits\theta).

3.2.5 Wild harmonic bundle with a real structure

Let (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) be a wild harmonic bundle on (X,D)(X,D). Around PDP\in D, let zPz_{P} denote a holomorphic local coordinate such that zP(P)=0z_{P}(P)=0. For an open neighbourhood UU of PP, let 𝒫ah(E)(U)\mathcal{P}^{h}_{a}(E)(U) denote the space of holomorphic sections ss of E|U{P}E_{|U\setminus\{P\}} satisfying |s|h=O(|zP|aϵ)|s|_{h}=O(|z_{P}|^{-a-\epsilon}) for any ϵ>0\epsilon>0. In this way, we obtain the associated good filtered Higgs bundle (𝒫h(E),θ)(\mathcal{P}^{h}_{\ast}(E),\theta) on (X,D)(X,D) [18, 12]. Let CC be a real structure of the harmonic bundle (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h). (See §2.2.2.)

Lemma 3.16

CC induces a perfect symmetric pairing of (𝒫hE,θ)(\mathcal{P}^{h}_{\ast}E,\theta).

Proof   Let hh^{\lor} be the induced metric of EE^{\lor}. The dual of 𝒫h(E)\mathcal{P}^{h}_{\ast}(E) is naturally isomorphic to 𝒫h(E)\mathcal{P}^{h^{\lor}}_{\ast}(E^{\lor}). Because ΨC\Psi_{C} extends to an isomorphism 𝒫h(E)𝒫h(E)\mathcal{P}^{h}_{\ast}(E)\simeq\mathcal{P}^{h^{\lor}}_{\ast}(E^{\lor}), we obtain the claim of the lemma.  

We recall the following proposition [12].

Proposition 3.17

Suppose that XX is a compact Riemann surface.

  • (𝒫h(E),θ)(\mathcal{P}_{\ast}^{h}(E),\theta) is polystable of degree 0.

  • If 𝒫h1(E)=𝒫h2(E)\mathcal{P}^{h_{1}}_{\ast}(E)=\mathcal{P}^{h_{2}}_{\ast}(E) for hiHarm(E,¯E,θ;C)h_{i}\in\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C), there exists a decomposition (𝒫h1(E),θ)=i=1m(𝒫𝒱i,θi)(\mathcal{P}^{h_{1}}_{\ast}(E),\theta)=\bigoplus_{i=1}^{m}(\mathcal{P}_{\ast}\mathcal{V}_{i},\theta_{i}) such that (i) E=𝒱i|XDE=\bigoplus\mathcal{V}_{i|X\setminus D} is orthogonal with respect to both h1h_{1} and h2h_{2}, (ii) h2=aih1h_{2}=a_{i}h_{1} for some ai>0a_{i}>0 on 𝒱i|XD\mathcal{V}_{i|X\setminus D}. In particular, h1h_{1} and h2h_{2} are mutually bounded, and we have ¯Es(h1,h2)=0\overline{\partial}_{E}s(h_{1},h_{2})=0 and [θ,s(h1,h2)]=0[\theta,s(h_{1},h_{2})]=0.  

We obtain the following corollary from Proposition 2.29.

Corollary 3.18

In Proposition 3.17, the GL(n,)\mathop{\rm GL}\nolimits(n,{\mathbb{R}})-harmonic bundles associated with (E,¯E,θ,hi;C)(E,\overline{\partial}_{E},\theta,h_{i};C) are naturally isomorphic.  

3.3 Kobayashi-Hitchin correspondence

Suppose that XX is compact.

3.3.1 Basic polystable objects (1)

Let (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) be a stable good filtered Higgs bundle of degree 0 on (X,D)(X,D) such that (𝒫𝒱,θ)(𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\simeq(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}). Let P1P_{1} be a pairing

𝒫𝒱𝒫𝒱𝒫(0)(𝒪X(D))\mathcal{P}_{\ast}\mathcal{V}\otimes\mathcal{P}_{\ast}\mathcal{V}\to\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D))

such that ΨP1\Psi_{P_{1}} induces an isomorphism (𝒫𝒱,θ)(𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\simeq(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}). Note that if P1P_{1}^{\prime} is such another pairing, then there exists α\alpha\in{\mathbb{C}} such that P1=αP1P_{1}^{\prime}=\alpha P_{1} by the stability assumption.

Lemma 3.19

Either one of P1ex=P1P_{1}\circ\mathop{\rm ex}\nolimits=P_{1} or P1ex=P1P_{1}\circ\mathop{\rm ex}\nolimits=-P_{1} holds.

Proof   Because of the stability condition, there exists β\beta\in{\mathbb{C}} such that ΨP1=βΨP1\Psi_{P_{1}}^{\lor}=\beta\Psi_{P_{1}}. Because (ΨP1)=ΨP1(\Psi_{P_{1}}^{\lor})^{\lor}=\Psi_{P_{1}}, we obtain β2=1\beta^{2}=1. Because ΨP1ex=ΨP1\Psi_{P_{1}\circ\mathop{\rm ex}\nolimits}=\Psi_{P_{1}}^{\lor}, we obtain the claim of the lemma.  

Let CC_{{\mathbb{C}}^{\ell}} denote the symmetric bilinear form of {\mathbb{C}}^{\ell} defined by C(𝒙,𝒚)=i=1xiyiC_{{\mathbb{C}}^{\ell}}({\boldsymbol{x}},\boldsymbol{y})=\sum_{i=1}^{\ell}x_{i}y_{i} for 𝒙=(xi){\boldsymbol{x}}=(x_{i}) and 𝒚=(yi)\boldsymbol{y}=(y_{i}). Let ω2k\omega_{{\mathbb{C}}^{2k}} denote the skew-symmetric bilinear form of 2k{\mathbb{C}}^{2k} defined by ω2k(𝒙,𝒚)=i=1k(x2i1y2ix2iy2i1)\omega_{{\mathbb{C}}^{2k}}({\boldsymbol{x}},\boldsymbol{y})=\sum_{i=1}^{k}(x_{2i-1}y_{2i}-x_{2i}y_{2i-1}). If P1P_{1} is symmetric, we obtain a perfect symmetric pairing P1CP_{1}\otimes C_{{\mathbb{C}}^{\ell}} of (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell}. If P1P_{1} is skew-symmetric, we obtain a perfect symmetric pairing P1ω2kP_{1}\otimes\omega_{{\mathbb{C}}^{2k}} of (𝒫𝒱,θ)2k(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{2k}.

Lemma 3.20

Suppose that (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell} is equipped with a perfect symmetric pairing CC.

  • If P1P_{1} is symmetric, there exists an automorphism ρ\rho of {\mathbb{C}}^{\ell} such that (id𝒱ρ)C=P1C(\mathop{\rm id}\nolimits_{\mathcal{V}}\otimes\rho)^{\ast}C=P_{1}\otimes C_{{\mathbb{C}}^{\ell}}.

  • If P1P_{1} is skew-symmetric, \ell is an even number 2k2k, and there exists an automorphism ρ\rho of 2k{\mathbb{C}}^{2k} such that (id𝒱ρ)C=P1ω2k(\mathop{\rm id}\nolimits_{\mathcal{V}}\otimes\rho)^{\ast}C=P_{1}\otimes\omega_{{\mathbb{C}}^{2k}}.

Proof   There exists a non-degenerate bilinear form C1C_{1} of {\mathbb{C}}^{\ell} such that C=P1C1C=P_{1}\otimes C_{1}. If P1P_{1} is symmetric, then C1C_{1} is symmetric. By using an orthonormal frame of {\mathbb{C}}^{\ell} with respect to C1C_{1}, we obtain the first claim. If P1P_{1} is skew-symmetric, C1C_{1} is skew-symmetric. In particular, =2k\ell=2k for a positive integer kk. By using a symplectic base of 2k{\mathbb{C}}^{2k}, we obtain the second claim.  

Lemma 3.21

There exists a unique harmonic metric h0h_{0} of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} such that (i) ΨP1\Psi_{P_{1}} is isometric with respect to h0h_{0} and h0h_{0}^{\lor}, (ii) h0h_{0} is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}.

Proof   Let h1h_{1} be a harmonic metric of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} which is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}. Let h1h_{1}^{\lor} denote the induced harmonic metric of (𝒱,θ)|XD(\mathcal{V}^{\lor},\theta^{\lor})_{|X\setminus D}, which is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}^{\lor}. Note that both (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) and (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}) are stable of degree 0, and that ΨP1:(𝒫𝒱,θ)(𝒫𝒱,θ)\Psi_{P_{1}}:(\mathcal{P}_{\ast}\mathcal{V},\theta)\longrightarrow(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}) is an isomorphism. Hence, there exists c>0c>0 such that ΨP1(h1)=c2h1\Psi_{P_{1}}^{\ast}(h_{1}^{\lor})=c^{2}h_{1}. We set h0=ch1h_{0}=ch_{1}. Because h0=c1h1h_{0}^{\lor}=c^{-1}h_{1}^{\lor}, h0h_{0} has the desired property. The uniqueness is also clear.  

The following lemma follows from the uniqueness of harmonic metrics adapted to parabolic structure.

Lemma 3.22

  • For any Hermitian metric hh_{{\mathbb{C}}^{\ell}} of {\mathbb{C}}^{\ell}, h0hh_{0}\otimes h_{{\mathbb{C}}^{\ell}} is a harmonic metric of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} which is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}\otimes{\mathbb{C}}^{\ell}. Conversely, for any harmonic metric hh of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} which is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}, there exists a Hermitian metric hh_{{\mathbb{C}}^{\ell}} of {\mathbb{C}}^{\ell} such that h=h0hh=h_{0}\otimes h_{{\mathbb{C}}^{\ell}}.

  • If P1P_{1} is symmetric (resp. skew-symmetric), a harmonic metric h0hh_{0}\otimes h_{{\mathbb{C}}^{\ell}} of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} is compatible with P1CP_{1}\otimes C_{{\mathbb{C}}^{\ell}} (resp. P1ωP_{1}\otimes\omega_{{\mathbb{C}}^{\ell}}) if and only if hh_{{\mathbb{C}}^{\ell}} is compatible with CC_{{\mathbb{C}}^{\ell}} (resp. ω\omega_{{\mathbb{C}}^{\ell}}).  

See §2.1.4 and §2.1.7 for the ambiguity of hh_{{\mathbb{C}}^{\ell}} compatible with CC_{{\mathbb{C}}^{\ell}} or ω\omega_{{\mathbb{C}}^{\ell}}.

3.3.2 Basic polystable objects (2)

Let (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) be a stable good filtered Higgs bundle of degree 0 on (X,D)(X,D). Assume that (𝒫𝒱,θ)≄(𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\not\simeq(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}). We set 𝒫𝒱~=𝒫𝒱𝒫(𝒱)\mathcal{P}_{\ast}\widetilde{\mathcal{V}}=\mathcal{P}_{\ast}\mathcal{V}\oplus\mathcal{P}_{\ast}(\mathcal{V}^{\lor}) which is equipped with the Higgs field θ~=θθ\widetilde{\theta}=\theta\oplus\theta^{\lor}. We have the naturally defined perfect pairing

𝒫𝒱𝒫(𝒱)𝒫(0)(𝒪X(D)).\mathcal{P}_{\ast}\mathcal{V}\otimes\mathcal{P}_{\ast}(\mathcal{V}^{\lor})\longrightarrow\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)).

It induces a symmetric product C~(𝒫𝒱,θ)\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V},\theta)} of (𝒫𝒱~,θ~)(\mathcal{P}_{\ast}\widetilde{\mathcal{V}},\widetilde{\theta}).

Remark 3.23

If (𝒫𝒱,θ)(𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\simeq(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}), (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) has a symmetric or skew symmetric pairing P1P_{1}. If P1P_{1} is symmetric (skew-symmetric), (𝒫𝒱~,θ~,C~(𝒫𝒱,θ))(\mathcal{P}_{\ast}\widetilde{\mathcal{V}},\widetilde{\theta},\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V},\theta)}) is isomorphic to (𝒫𝒱,θ)2(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{2} with P1C2P_{1}\otimes C_{{\mathbb{C}}^{2}} (resp. P1ω2P_{1}\otimes\omega_{{\mathbb{C}}^{2}}).  

Lemma 3.24

Suppose that ((𝒫𝒱,θ)1)((𝒫𝒱,θ)2)\bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell_{1}}\bigr{)}\oplus\bigl{(}(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor})\otimes{\mathbb{C}}^{\ell_{2}}\bigr{)} is equipped with a perfect symmetric pairing CC. Then, we have 1=2\ell_{1}=\ell_{2}, and there exists an isomorphism (𝒫𝒱~,θ~)1((𝒫𝒱,θ)1)((𝒫𝒱,θ)2)(\mathcal{P}_{\ast}\widetilde{\mathcal{V}},\widetilde{\theta})\otimes{\mathbb{C}}^{\ell_{1}}\simeq\bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell_{1}}\bigr{)}\oplus\bigl{(}(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor})\otimes{\mathbb{C}}^{\ell_{2}}\bigr{)} under which C~(𝒫𝒱,θ)C1=C\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V},\theta)}\otimes C_{{\mathbb{C}}^{\ell_{1}}}=C.

Proof   There exist 11-dimensional subspaces LiiL_{i}\subset{\mathbb{C}}^{\ell_{i}} such that the restriction of CC to ((𝒫𝒱,θ)L1)((𝒫𝒱,θ)L2)\Bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes L_{1}\Bigr{)}\oplus\Bigl{(}(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor})\otimes L_{2}\Bigr{)} is not 0. Because (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) and (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}) are stable, and because (𝒫𝒱,θ)≄(𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta)\not\simeq(\mathcal{P}_{\ast}\mathcal{V}^{\lor},\theta^{\lor}), the restriction is equal to αC~(𝒫𝒱,θ)\alpha\cdot\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V},\theta)} for a non-zero α\alpha. In particular, it is perfect. We obtain the decomposition of filtered bundles which is orthogonal with respect to CC:

(𝒫(𝒱)1)(𝒫(𝒱)2)=((𝒫(𝒱)L1)(𝒫(𝒱)L2))𝒫𝒱.\Bigl{(}\mathcal{P}_{\ast}(\mathcal{V})\otimes{\mathbb{C}}^{\ell_{1}}\Bigr{)}\oplus\Bigl{(}\mathcal{P}_{\ast}(\mathcal{V}^{\lor})\otimes{\mathbb{C}}^{\ell_{2}}\Bigr{)}=\Bigl{(}\bigl{(}\mathcal{P}_{\ast}(\mathcal{V})\otimes L_{1}\bigr{)}\oplus\bigl{(}\mathcal{P}_{\ast}(\mathcal{V}^{\lor})\otimes L_{2}\bigr{)}\Bigr{)}\oplus\mathcal{P}_{\ast}\mathcal{V}^{\prime}.

It is preserved by the Higgs field, and 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}^{\prime} with the induced Higgs field is isomorphic to (𝒫𝒱,θ)11(𝒫𝒱,θ)21(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell_{1}-1}\oplus(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell_{2}-1}. Hence, we obtain the claim of the lemma by an easy induction.  

By using CC_{{\mathbb{C}}^{\ell}}, we identify {\mathbb{C}}^{\ell} and the dual space ()({\mathbb{C}}^{\ell})^{\lor}. Then, the perfect symmetric bilinear form C~(𝒫𝒱,θ)C\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V},\theta)}\otimes C_{{\mathbb{C}}^{\ell}} on

(𝒫𝒱~,θ~)=((𝒫𝒱,θ))((𝒫𝒱,θ)())(\mathcal{P}_{\ast}\widetilde{\mathcal{V}},\widetilde{\theta})\otimes{\mathbb{C}}^{\ell}=\bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell}\bigr{)}\oplus\bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)^{\lor}\otimes({\mathbb{C}}^{\ell})^{\lor}\bigr{)}

is identified with the symmetric pairing induced by the natural pairing

((𝒫𝒱,θ))((𝒫𝒱,θ)())𝒫(0)(𝒪X(D)).\bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)\otimes{\mathbb{C}}^{\ell}\bigr{)}\otimes\bigl{(}(\mathcal{P}_{\ast}\mathcal{V},\theta)^{\lor}\otimes({\mathbb{C}}^{\ell})^{\lor}\bigr{)}\longrightarrow\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)).

Let h0h_{0} be any harmonic metric of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} which is adapted to 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}). Note that for any harmonic metric h0h^{\prime}_{0} of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} which is adapted to 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}), there exists c>0c>0 such that h0=ch0h^{\prime}_{0}=ch_{0}. We obtain the induced harmonic metric h0h_{0}^{\lor} of (𝒱,θ)|XD(\mathcal{V}^{\lor},\theta^{\lor})_{|X\setminus D} which is adapted to 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}^{\lor}).

Lemma 3.25

  • Let hh_{{\mathbb{C}}^{\ell}} be any Hermitian metric of {\mathbb{C}}^{\ell}. Let hh^{\lor}_{{\mathbb{C}}^{\ell}} denote the induced Hermitian metric on ()({\mathbb{C}}^{\ell})^{\lor}. Then, (h0h)(h0h)(h_{0}\otimes h_{{\mathbb{C}}^{\ell}})\oplus(h_{0}^{\lor}\otimes h^{\lor}_{{\mathbb{C}}^{\ell}}) is a harmonic metric of (𝒱~,θ~)|XD(\widetilde{\mathcal{V}},\widetilde{\theta})_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} such that (i) it is adapted to 𝒫𝒱~\mathcal{P}_{\ast}\widetilde{\mathcal{V}}, (ii) it is compatible with C~(𝒫𝒱,θ)C\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V},\theta)}\otimes C_{{\mathbb{C}}^{\ell}}.

  • Conversely, let hh be any harmonic metric of (𝒱~,θ~)|XD(\widetilde{\mathcal{V}},\widetilde{\theta})_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} satisfying the above conditions (i) and (ii). Then, there exists a Hermitian metric hh_{{\mathbb{C}}^{\ell}} of {\mathbb{C}}^{\ell} such that h=(h0h)(h0h)h=(h_{0}\otimes h_{{\mathbb{C}}^{\ell}})\oplus(h_{0}^{\lor}\otimes h^{\lor}_{{\mathbb{C}}^{\ell}}).

Proof   The first claim is clear. Let hh be as in the second claim. By the uniqueness of harmonic metrics to a parabolic structure (see Proposition 3.17), 𝒱|XD\mathcal{V}_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} and 𝒱|XD\mathcal{V}^{\lor}_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} are orthogonal with respect to hh. Let h𝒱h_{\mathcal{V}} and h𝒱h_{\mathcal{V}^{\lor}} denote the restrictions of hh to 𝒱|XD\mathcal{V}_{|X\setminus D}\otimes{\mathbb{C}}^{\ell} and 𝒱|XD\mathcal{V}^{\lor}_{|X\setminus D}\otimes{\mathbb{C}}^{\ell}, respectively. By the uniqueness again, there exists a Hermitian metric hh_{{\mathbb{C}}^{\ell}} of {\mathbb{C}}^{\ell} such that h𝒱=h0hh_{\mathcal{V}}=h_{0}\otimes h_{{\mathbb{C}}^{\ell}}. There also exists a Hermitian metric h()h^{\prime}_{({\mathbb{C}}^{\ell})^{\lor}} of ()({\mathbb{C}}^{\ell})^{\lor} such that h𝒱=h0h()h_{\mathcal{V}^{\lor}}=h_{0}^{\lor}\otimes h^{\prime}_{({\mathbb{C}}^{\ell})^{\lor}}. By the compatibility condition, we obtain that h()=hh^{\prime}_{({\mathbb{C}}^{\ell})^{\lor}}=h_{{\mathbb{C}}^{\ell}}^{\lor}.  

3.3.3 Polystable objects

Let (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) be a polystable good filtered Higgs bundle of degree 0 on (X,D)(X,D) with a perfect symmetric pairing CC.

Proposition 3.26

There exist stable good filtered Higgs bundles (𝒫𝒱i(0),θi(0))(\mathcal{P}_{\ast}\mathcal{V}^{(0)}_{i},\theta^{(0)}_{i}) (i=1,,p(0))(i=1,\ldots,p(0)), (𝒫𝒱i(1),θi(1))(\mathcal{P}_{\ast}\mathcal{V}^{(1)}_{i},\theta^{(1)}_{i}) (i=1,,p(1))(i=1,\ldots,p(1)), and (𝒫𝒱i(2),θi(2))(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i}) (i=1,,p(2))(i=1,\ldots,p(2)) of degree 0 on (X,D)(X,D) such that the following holds.

  • (𝒫𝒱i(0),θi(0))(\mathcal{P}_{\ast}\mathcal{V}^{(0)}_{i},\theta^{(0)}_{i}) is equipped with a symmetric perfect pairing Pi(0)P^{(0)}_{i}.

  • (𝒫𝒱i(1),θi(1))(\mathcal{P}_{\ast}\mathcal{V}^{(1)}_{i},\theta^{(1)}_{i}) is equipped with a skew-symmetric perfect pairing Pi(1)P^{(1)}_{i}.

  • (𝒫𝒱i(2),θi(2))≄(𝒫𝒱i(2),θi(2))(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})\not\simeq(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})^{\lor}.

  • There exist positive integers (a,i)\ell(a,i) and an isomorphism

    (𝒫𝒱,θ)i=1p(0)(𝒫𝒱i(0),θi(0))(0,i)i=1p(1)(𝒫𝒱i(1),θi(1))2(1,i)i=1p(2)(((𝒫𝒱i(2),θi(2))(2,i))((𝒫𝒱i(2),θi(2))((2,i)))).(\mathcal{P}_{\ast}\mathcal{V},\theta)\simeq\bigoplus_{i=1}^{p(0)}(\mathcal{P}_{\ast}\mathcal{V}^{(0)}_{i},\theta^{(0)}_{i})\otimes{\mathbb{C}}^{\ell(0,i)}\oplus\bigoplus_{i=1}^{p(1)}(\mathcal{P}_{\ast}\mathcal{V}^{(1)}_{i},\theta^{(1)}_{i})\otimes{\mathbb{C}}^{2\ell(1,i)}\oplus\\ \bigoplus_{i=1}^{p(2)}\Bigl{(}\bigl{(}(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})\otimes{\mathbb{C}}^{\ell(2,i)}\bigr{)}\oplus\bigl{(}(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})^{\lor}\otimes({\mathbb{C}}^{\ell(2,i)})^{\lor}\bigr{)}\Bigr{)}. (19)

    Under this isomorphism, CC is identified with the direct sum of P1(0)C(0,i)P^{(0)}_{1}\otimes C_{{\mathbb{C}}^{\ell(0,i)}}, P1(1)ω2(1,i)P^{(1)}_{1}\otimes\omega_{{\mathbb{C}}^{2\ell(1,i)}} and C~(𝒫𝒱i(2),θi(2))C(2,i)\widetilde{C}_{(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})}\otimes C_{{\mathbb{C}}^{\ell(2,i)}}.

  • (𝒫𝒱i(a),θi(a))≄(𝒫𝒱j(a),θj(a))(\mathcal{P}_{\ast}\mathcal{V}^{(a)}_{i},\theta^{(a)}_{i})\not\simeq(\mathcal{P}_{\ast}\mathcal{V}^{(a)}_{j},\theta^{(a)}_{j}) (ij)(i\neq j) for a=0,1,2a=0,1,2, and (𝒫𝒱i(2),θi(2))≄(𝒫𝒱j(2),θj(2))(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})\not\simeq(\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{j},\theta^{(2)}_{j})^{\lor} for any i,ji,j.

Proof   It follows from Lemma 3.20 and Lemma 3.24.  

Let hi(a)h^{(a)}_{i} (a=0,1)(a=0,1) be unique harmonic metrics of (𝒱i(a),θi(a))|XD(\mathcal{V}^{(a)}_{i},\theta^{(a)}_{i})_{|X\setminus D} such that (i) hi(a)h^{(a)}_{i} are adapted to 𝒫𝒱i(a)\mathcal{P}_{\ast}\mathcal{V}^{(a)}_{i}, (ii) ΨPi(a)\Psi_{P^{(a)}_{i}} are isometric with respect to hi(a)h^{(a)}_{i} and (hi(a))(h^{(a)}_{i})^{\lor}. Let hi(2)h^{(2)}_{i} be harmonic metrics of (𝒱i(2),θi(2))|XD(\mathcal{V}^{(2)}_{i},\theta^{(2)}_{i})_{|X\setminus D} adapted to 𝒫𝒱i(2)\mathcal{P}_{\ast}\mathcal{V}^{(2)}_{i}.

Proposition 3.27

There exists a harmonic metric hh of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} such that (i) it is adapted to 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V}, (ii) it is compatible with CC. More precisely, the following holds.

  • Let h(0,i)h_{{\mathbb{C}}^{\ell(0,i)}} be Hermitian metrics of (0,i){\mathbb{C}}^{\ell(0,i)} compatible with C(0,i)C_{{\mathbb{C}}^{\ell(0,i)}}. Let h2(1,i)h_{{\mathbb{C}}^{2\ell(1,i)}} be Hermitian metrics of 2(1,i){\mathbb{C}}^{2\ell(1,i)} compatible with ω2(0,i)\omega_{{\mathbb{C}}^{2\ell(0,i)}}. Let h(2,i)h_{{\mathbb{C}}^{\ell(2,i)}} be any Hermitian metric of (2,i){\mathbb{C}}^{\ell(2,i)}. Then,

    i=1p(0)(hi(0)h(0,i))i=1p(1)(hi(1)h2(1,i))i=1p(2)((hi(2)h(2,i))((hi(2))(h(2,i))))\bigoplus_{i=1}^{p(0)}\bigl{(}h^{(0)}_{i}\otimes h_{{\mathbb{C}}^{\ell(0,i)}}\bigr{)}\oplus\bigoplus_{i=1}^{p(1)}\bigl{(}h^{(1)}_{i}\otimes h_{{\mathbb{C}}^{2\ell(1,i)}}\bigr{)}\oplus\bigoplus_{i=1}^{p(2)}\Bigl{(}\bigl{(}h^{(2)}_{i}\otimes h_{{\mathbb{C}}^{\ell(2,i)}}\bigr{)}\oplus\bigl{(}(h^{(2)}_{i})^{\lor}\otimes(h_{{\mathbb{C}}^{\ell(2,i)}})^{\lor}\bigr{)}\Bigr{)} (20)

    is a harmonic metric of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} satisfying the conditions (i) and (ii).

  • Conversely, if hh is a harmonic metric of (𝒱,θ)|XD(\mathcal{V},\theta)_{|X\setminus D} satisfying the conditions (i) and (ii), hh is of the form (20).

Proof   The first claim is obvious. The second claim follows from Lemma 3.22 and Lemma 3.25.  

3.3.4 An equivalence

Let (E,¯E,θ,h)(E,\overline{\partial}_{E},\theta,h) be a wild harmonic bundle on (X,D)(X,D) with a real structure CC. As explained in §3.2.5, we obtain a good filtered Higgs bundle (𝒫h(E),θ)(\mathcal{P}_{\ast}^{h}(E),\theta) on (X,D)(X,D) equipped with a perfect symmetric pairing CC. It is polystable of degree 0.

Theorem 3.28

The above construction induces an equivalence between the following objects.

  • Wild harmonic bundles on (X,D)(X,D) equipped with a real structure.

  • Polystable good filtered Higgs bundles of degree 0 on (X,D)(X,D) equipped with a perfect symmetric pairing.

Proof   We obtain the converse construction from Proposition 3.27.  

4 Generically regular semisimple case

4.1 Prolongation of regular semisimple wild Higgs bundles on a punctured disc

4.1.1 Regular semisimple wild Higgs bundles

Let UU be a neighbourhood of 0 in {\mathbb{C}}. For each e>0e\in{\mathbb{Z}}_{>0}, let φe:\varphi_{e}:{\mathbb{C}}\longrightarrow{\mathbb{C}} be given defined by φe(ζ)=ζe\varphi_{e}(\zeta)=\zeta^{e}. We set U(e)=φ1(U)U^{(e)}=\varphi^{-1}(U) on which we set ze=ζz_{e}=\zeta. We have zee=φe(z)z_{e}^{e}=\varphi_{e}^{\ast}(z).

Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on U{0}U\setminus\{0\} of rank rr equipped with a non-degenerate symmetric pairing CC. We obtain the endomorphism ff determined by θ=fdz/z\theta=f\,dz/z. We obtain the characteristic polynomial det(TidEf)=Tr+j=0r1aj(z)Tj\det(T\mathop{\rm id}\nolimits_{E}-f)=T^{r}+\sum_{j=0}^{r-1}a_{j}(z)T^{j}, where aj(z)a_{j}(z) are holomorphic function on U{0}U\setminus\{0\}. We assume the following condition.

  • (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple, i.e., for each z0U{0}z_{0}\in U\setminus\{0\}, the polynomial Tr+j=0r1aj(z0)TjT^{r}+\sum_{j=0}^{r-1}a_{j}(z_{0})T^{j} has rr-distinct roots.

  • (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is wild, i.e., aj(z)a_{j}(z) are meromorphic at z=0z=0.

There exist a divisor ee of r!r! and holomorphic functions α1,,αr\alpha_{1},\ldots,\alpha_{r} on U(e){0}U^{(e)}\setminus\{0\} such that

Tr+j=0r1φe(aj)Tj=j=1r(Tαj(ze)),T^{r}+\sum_{j=0}^{r-1}\varphi_{e}^{\ast}(a_{j})T^{j}=\prod_{j=1}^{r}(T-\alpha_{j}(z_{e})),

and that αiαj\alpha_{i}-\alpha_{j} (ij)(i\neq j) are nowhere vanishing on U(e){0}U^{(e)}\setminus\{0\}. Because aj(z)a_{j}(z) are meromorphic at z=0z=0, we obtain that αj(z)\alpha_{j}(z) are meromorphic at z=0z=0. We set E(e):=φe(E)E^{(e)}:=\varphi_{e}^{\ast}(E), which is equipped with the endomorphism f(e):=φe(f)f^{(e)}:=\varphi_{e}^{\ast}(f). There exists the eigen decomposition

(E(e),f(e))=i=1r(Ei(e),αiidEi(e)).(E^{(e)},f^{(e)})=\bigoplus_{i=1}^{r}(E^{(e)}_{i},\alpha_{i}\mathop{\rm id}\nolimits_{E^{(e)}_{i}}).

It is orthogonal with respect to C(e):=φe(C)C^{(e)}:=\varphi_{e}^{\ast}(C). The restriction of C(e)C^{(e)} to Ei(e)E^{(e)}_{i} are denoted by Ci(e)C^{(e)}_{i}.

Let Gale\mathop{\rm Gal}\nolimits_{e} denote the Galois group of the covering U(e){0}U{0}U^{(e)}\setminus\{0\}\to U\setminus\{0\}, which is a cyclic group of order ee. The pull back φ(E,f,C)\varphi^{\ast}(E,f,C) is naturally equivariant with respect to Gale\mathop{\rm Gal}\nolimits_{e}. There exists the naturally induced Gale\mathop{\rm Gal}\nolimits_{e}-action on {1,,r}\{1,\ldots,r\} such that b(αi)=αb(i)b^{\ast}(\alpha_{i})=\alpha_{b(i)} and b(Ei(e))=Eb(i)(e)b^{\ast}(E^{(e)}_{i})=E^{(e)}_{b(i)} for any bGaleb\in\mathop{\rm Gal}\nolimits_{e}.

4.1.2 Canonical meromorphic extension

Let ι:U{0}U\iota:U\setminus\{0\}\to U denote the inclusion. There exists a natural inclusion 𝒪U(0)ι(𝒪U{0})\mathcal{O}_{U}(\ast 0)\subset\iota_{\ast}(\mathcal{O}_{U\setminus\{0\}}). We naturally regard EE as an 𝒪U{0}\mathcal{O}_{U\setminus\{0\}}-module, and we obtain the ι(𝒪U{0})\iota_{\ast}(\mathcal{O}_{U\setminus\{0\}})-module ι(E)\iota_{\ast}(E). We obtain the morphisms ι(f):ι(E)ι(E)\iota_{\ast}(f):\iota_{\ast}(E)\to\iota_{\ast}(E) and ι(C):ι(E)ι(E)ι(𝒪U{0})\iota_{\ast}(C):\iota_{\ast}(E)\otimes\iota_{\ast}(E)\longrightarrow\iota_{\ast}(\mathcal{O}_{U\setminus\{0\}}).

For any locally free 𝒪U{0}\mathcal{O}_{U\setminus\{0\}}-module \mathcal{F}, a locally free 𝒪U(0)\mathcal{O}_{U}(\ast 0)-submodule 1ιU\mathcal{F}_{1}\subset\iota_{U\ast}\mathcal{F} is called a meromorphic extension of \mathcal{F} if ιU(1)=\iota_{U}^{\ast}(\mathcal{F}_{1})=\mathcal{F}.

Proposition 4.1

There exists a unique meromorphic extension E~\widetilde{E} of EE such that (i) ι(f)(E~)E~\iota_{\ast}(f)(\widetilde{E})\subset\widetilde{E}, (ii) ι(C)(E~E~)𝒪U(0)\iota_{\ast}(C)(\widetilde{E}\otimes\widetilde{E})\subset\mathcal{O}_{U}(\ast 0).

Proof   Let ιe:U(e){0}U(e)\iota_{e}:U^{(e)}\setminus\{0\}\to U^{(e)} denote the inclusion. A meromorphic extension of a locally free 𝒪U(e){0}\mathcal{O}_{U^{(e)}\setminus\{0\}}-module \mathcal{F} is defined to be a locally free 𝒪U(e)(0)\mathcal{O}_{U^{(e)}}(\ast 0)-submodule 1ιe()\mathcal{F}_{1}\subset\iota_{e\ast}(\mathcal{F}) such that ιe(1)=\iota_{e}^{\ast}(\mathcal{F}_{1})=\mathcal{F}.

Lemma 4.2

For each ii, there exists a unique meromorphic extension E~i(e)\widetilde{E}^{(e)}_{i} of Ei(e)E^{(e)}_{i} such that Ci(e)C^{(e)}_{i} induces an isomorphism

E~i(e)E~i(e)𝒪U(e)(0).\widetilde{E}^{(e)}_{i}\otimes\widetilde{E}^{(e)}_{i}\simeq\mathcal{O}_{U^{(e)}}(\ast 0). (21)

Proof   There exists a holomorphic frame ui(e)u^{(e)}_{i} of Ei(e)E^{(e)}_{i} on U(e){0}U^{(e)}\setminus\{0\}. We obtain the holomorphic function g=Ci(e)(ui(e),ui(e))g=C^{(e)}_{i}(u^{(e)}_{i},u^{(e)}_{i}). There exist a holomorphic function g1g_{1} such that g12=gg_{1}^{2}=g or g12=zegg_{1}^{2}=z_{e}g. We set u~i(e):=g11ui(e)\widetilde{u}^{(e)}_{i}:=g_{1}^{-1}u^{(e)}_{i}, and E~i(e):=𝒪U(e)(0)u~i(e)ιe(Ei(e))\widetilde{E}^{(e)}_{i}:=\mathcal{O}_{U^{(e)}}(\ast 0)\cdot\widetilde{u}^{(e)}_{i}\subset\iota_{e\ast}(E^{(e)}_{i}). Then, E~i(e)\widetilde{E}^{(e)}_{i} satisfies (21).

Let E~i(e)\widetilde{E}^{\prime(e)}_{i} be another meromorphic extension of Ei(e)E^{(e)}_{i} satisfying (21). There exists a holomorphic frame vi(e)v^{(e)}_{i} of E~i(e)\widetilde{E}^{\prime(e)}_{i} on a neighbourhood 𝒰\mathcal{U} of 0 in U(e)U^{(e)}. We obtain the holomorphic function g2g_{2} on 𝒰{0}\mathcal{U}\setminus\{0\} defined by vi(e)=g2u~i(e)v^{(e)}_{i}=g_{2}\cdot\widetilde{u}^{(e)}_{i}. Because both Ci(e)(vi(e)vi(e))C^{(e)}_{i}(v^{(e)}_{i}\otimes v^{(e)}_{i}) and Ci(e)(u~i(e)u~i(e))C^{(e)}_{i}(\widetilde{u}^{(e)}_{i}\otimes\widetilde{u}^{(e)}_{i}) are meromorphic at ze=0z_{e}=0, we obtain that g22g_{2}^{2} is meromorphic at ze=0z_{e}=0, and hence g2g_{2} is meromorphic at ze=0z_{e}=0. It implies E~i(e)=E~i(e)\widetilde{E}^{\prime(e)}_{i}=\widetilde{E}^{(e)}_{i}.  

We obtain a meromorphic extension E~(e):=E~i(e)ιe(E(e))\widetilde{E}^{(e)}:=\bigoplus\widetilde{E}^{(e)}_{i}\subset\iota_{e\ast}(E^{(e)}) of E(e)E^{(e)}. It is naturally Gale\mathop{\rm Gal}\nolimits_{e}-equivariant. By the construction, ιe(f(e))(E~(e))E~(e)\iota_{e\ast}(f^{(e)})\bigl{(}\widetilde{E}^{(e)}\bigr{)}\subset\widetilde{E}^{(e)} and ιe(C(e))(E~(e)E~(e))𝒪U(e)(0)\iota_{e\ast}(C^{(e)})\bigl{(}\widetilde{E}^{(e)}\otimes\widetilde{E}^{(e)}\bigr{)}\subset\mathcal{O}_{U^{(e)}}(\ast 0). As the descent (see [14, §2.3.2]), we obtain a locally free 𝒪U(0)\mathcal{O}_{U}(\ast 0)-module E~\widetilde{E} with a Gale\mathop{\rm Gal}\nolimits_{e}-equivariant isomorphism φe(E~)E~(e)\varphi_{e}^{\ast}(\widetilde{E})\simeq\widetilde{E}^{(e)}. It is easy to see that E~\widetilde{E} is a meromorphic extension of EE with the desired property.

Let E~\widetilde{E}^{\prime} be another meromorphic extension with the desired property. We obtain a meromorphic extension E~(e):=φe(E~)\widetilde{E}^{\prime(e)}:=\varphi_{e}^{\ast}(\widetilde{E}^{\prime}) of E(e)E^{(e)} such that ιe(f(e))(E~(e))E~(e)\iota_{e\ast}(f^{(e)})\bigl{(}\widetilde{E}^{\prime(e)}\bigr{)}\subset\widetilde{E}^{\prime(e)} and ιe(C(e))(E~(e)E~(e))𝒪U(e)(0)\iota_{e\ast}(C^{(e)})\bigl{(}\widetilde{E}^{\prime(e)}\otimes\widetilde{E}^{\prime(e)}\bigr{)}\subset\mathcal{O}_{U^{(e)}}(\ast 0). Let f(e)f^{\prime\,(e)} and C(e)C^{\prime(e)} denote the induced endomorphism and the pairing of E~(e)\widetilde{E}^{\prime(e)}.

For a sheaf \mathcal{F} on U(e)U^{(e)}, let 0\mathcal{F}_{0} denote the stalk of \mathcal{F} at 0. Note that 𝔎=𝒪U(e)(0)0\mathfrak{K}=\mathcal{O}_{U^{(e)}}(\ast 0)_{0} is a field. We obtain a 𝔎\mathfrak{K}-vector space E~0(e)\widetilde{E}^{\prime(e)}_{0} with the linear endomorphism f0(e)f^{\prime(e)}_{0} and the symmetric bilinear pairing C0(e)C^{\prime(e)}_{0}. Because the characteristic polynomial of f0(e)f^{\prime(e)}_{0} is Tr+j=0r1φe(aj)Tj𝔎[T]T^{r}+\sum_{j=0}^{r-1}\varphi_{e}^{\ast}(a_{j})T^{j}\in\mathfrak{K}[T], the eigenvalues of f0(e)f^{\prime(e)}_{0} are α1,,αr𝔎\alpha_{1},\ldots,\alpha_{r}\in\mathfrak{K}, and there exists the eigen decomposition E~0(e)=i=1r(E~0(e))i\widetilde{E}^{\prime(e)}_{0}=\bigoplus_{i=1}^{r}(\widetilde{E}^{\prime(e)}_{0})_{i}, where f0(e)αiidf^{\prime(e)}_{0}-\alpha_{i}\mathop{\rm id}\nolimits are 0 on (E~0(e))i(\widetilde{E}^{\prime(e)}_{0})_{i}. Hence, there exists the decomposition of the 𝒪U(0)\mathcal{O}_{U}(\ast 0)-module E~(e)=E~i(e)\widetilde{E}^{\prime(e)}=\bigoplus\widetilde{E}^{\prime(e)}_{i} such that ιe(f(e))αiidE~(e)\iota_{e\ast}(f^{(e)})-\alpha_{i}\mathop{\rm id}\nolimits_{\widetilde{E}^{\prime(e)}} are 0 on E~i(e)\widetilde{E}^{\prime(e)}_{i}. Each E~i(e)\widetilde{E}^{\prime(e)}_{i} are meromorphic extension of Ei(e)E^{(e)}_{i}, and C(e)C^{\prime(e)} induces an isomorphism E~i(e)E~i(e)𝒪U(e)(0)\widetilde{E}^{\prime(e)}_{i}\otimes\widetilde{E}^{\prime(e)}_{i}\simeq\mathcal{O}_{U^{(e)}}(\ast 0). By the uniqueness in Lemma 4.2, we obtain E~i(e)=E~i(e)\widetilde{E}^{\prime(e)}_{i}=\widetilde{E}_{i}^{(e)}, and hence E~(e)=E~(e)\widetilde{E}^{\prime(e)}=\widetilde{E}^{(e)}. It implies E~=E~\widetilde{E}^{\prime}=\widetilde{E}.  

The induced endomorphism E~E~\widetilde{E}\to\widetilde{E} and the pairing E~E~𝒪U(0)\widetilde{E}\otimes\widetilde{E}\to\mathcal{O}_{U}(\ast 0) are denoted by f~\widetilde{f} and C~\widetilde{C}, respectively. Note that there exists the eigen decomposition

φe(E~,f~)=i=1r(E~i,αiidE~i).\varphi_{e}^{\ast}(\widetilde{E},\widetilde{f})=\bigoplus_{i=1}^{r}(\widetilde{E}_{i},\alpha_{i}\mathop{\rm id}\nolimits_{\widetilde{E}_{i}}). (22)

It is orthogonal with respect to φeC~\varphi_{e}^{\ast}\widetilde{C}.

4.1.3 Canonical filtered extension and regular semisimplicity at 0

Definition 4.3

A filtered bundle 𝒫E~\mathcal{P}_{\ast}\widetilde{E} is called a good filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC if the following condition is satisfied:

(i)

(𝒫E~,θ)(\mathcal{P}_{\ast}\widetilde{E},\theta) is a good filtered Higgs bundle.

(ii)

CC is a perfect pairing of 𝒫(E~)\mathcal{P}_{\ast}(\widetilde{E}).  

Lemma 4.4

There exists a unique filtered bundle 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}) over E~\widetilde{E} satisfying the conditions (i), (ii) and the following additional condition.

(iii)

φe(𝒫canE~)\varphi_{e}^{\ast}(\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}\widetilde{E}) is compatible with the decomposition (22), i.e.,

φe(𝒫canE~)=𝒫E~i(e).\varphi_{e}^{\ast}(\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}\widetilde{E})=\bigoplus\mathcal{P}_{\ast}\widetilde{E}^{(e)}_{i}. (23)

The filtered bundle 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}) is called the canonical filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC.

Proof   There exists a unique filtered bundle 𝒫E~i(e)\mathcal{P}_{\ast}\widetilde{E}^{(e)}_{i} such that 𝒫E~i(e)𝒫E~i(e)𝒫(0)(𝒪U(e)(0))\mathcal{P}_{\ast}\widetilde{E}^{(e)}_{i}\otimes\mathcal{P}_{\ast}\widetilde{E}^{(e)}_{i}\simeq\mathcal{P}_{\ast}^{(0)}(\mathcal{O}_{U^{(e)}}(\ast 0)). We obtain the filtered bundle 𝒫(E~(e))\mathcal{P}_{\ast}(\widetilde{E}^{(e)}) by the right hand side of (23). The uniqueness of such 𝒫E~i(e)\mathcal{P}_{\ast}\widetilde{E}^{(e)}_{i} implies that 𝒫(E~(e))\mathcal{P}_{\ast}(\widetilde{E}^{(e)}) is Gale\mathop{\rm Gal}\nolimits_{e}-equivariant, and we obtain a filtered bundle 𝒫(E~)\mathcal{P}_{\ast}(\widetilde{E}) over E~\widetilde{E} as the descent of 𝒫(E~(e))\mathcal{P}_{\ast}(\widetilde{E}^{(e)}), which has the desired property. The uniqueness is clear.  

Definition 4.5

We say that (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at 0 if the following condition is satisfied.

  • (αiαj)1(\alpha_{i}-\alpha_{j})^{-1} (ij)(i\neq j) are holomorphic at 0.  

Proposition 4.6

If (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at 0, any good filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC is equal to 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}).

Proof   If (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at 0, then 𝒫(E~(e))\mathcal{P}_{\ast}(\widetilde{E}^{(e)}) has to be compatible with the decomposition (22). Hence, the proposition is clear.  

Remark 4.7

If (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is not regular semisimple at 0, there may exist many good filtered extensions of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC, in general. (See Proposition 4.26.)  

4.1.4 Compatible harmonic metrics

Let hHarm(E,¯E,θ;C)h\in\mathop{\rm Harm}\nolimits(E,\overline{\partial}_{E},\theta;C). We obtain the good filtered Higgs bundle (𝒫hE,θ)(\mathcal{P}^{h}_{\ast}E,\theta) on (U,0)(U,0) with a perfect symmetric pairing CC as in Lemma 3.16. We also obtain the locally free 𝒪U(0)\mathcal{O}_{U}(\ast 0)-module 𝒫h(E)=a𝒫ah(E)\mathcal{P}^{h}(E)=\bigcup_{a\in{\mathbb{R}}}\mathcal{P}^{h}_{a}(E).

Proposition 4.8

The 𝒪U(0)\mathcal{O}_{U}(\ast 0)-module 𝒫hE\mathcal{P}^{h}E equals to the canonical meromorphic extension E~\widetilde{E}. If (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at 0, then we have 𝒫h(E)=𝒫can(E~)\mathcal{P}^{h}_{\ast}(E)=\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}).

Proof   Because CC is a perfect pairing of 𝒫h(E)\mathcal{P}^{h}_{\ast}(E), we obtain Proposition 4.8 from the uniqueness in Proposition 4.1 and Proposition 4.6.  

4.2 Classification of harmonic metrics by good filtered extensions

4.2.1 Setting

Let XX be a compact Riemann surface with a finite subset DD. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be a Higgs bundle on XDX\setminus D with a non-degenerate symmetric pairing CC. We assume the following conditions.

  • The Higgs bundle is wild at each point of DD.

  • (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple.

Lemma 4.9

There exists a finite subset Z1XDZ_{1}\subset X\setminus D such that (E,¯E,θ)|X(DZ1)(E,\overline{\partial}_{E},\theta)_{|X\setminus(D\cup Z_{1})} is regular semisimple.

Proof   Let PP be any point of DD. Let (XP,z)(X_{P},z) be a holomorphic coordinate neighbourhood around PP. We obtain the endomorphism ff of E|XP{P}E_{|X_{P}\setminus\{P\}} by θ=fdz/z\theta=f\,dz/z. We obtain the characteristic polynomial Pz(T)=Tr+j=0r1aj(z)=det(TidEf)P_{z}(T)=T^{r}+\sum_{j=0}^{r-1}a_{j}(z)=\det(T\mathop{\rm id}\nolimits_{E}-f). Let DiscT(Pz(T))\mathop{\rm Disc}\nolimits_{T}(P_{z}(T)) be the discriminant of Pz(T)P_{z}(T), which is a holomorphic function on XP{P}X_{P}\setminus\{P\}. Because (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple, there exists a discrete subset ZPXP{P}Z_{P}\subset X_{P}\setminus\{P\} such that DiscT(Pz(T))0\mathop{\rm Disc}\nolimits_{T}(P_{z}(T))\neq 0 unless zZPz\in Z_{P}. Because aj(z)a_{j}(z) are meromorphic at z=0z=0, DiscT(Pz(T))\mathop{\rm Disc}\nolimits_{T}(P_{z}(T)) is meromorphic at z=0z=0. Hence, we obtain the finiteness of ZPZ_{P}.  

Let ι:XDX\iota:X\setminus D\to X denote the inclusion. A meromorphic extension of a locally free 𝒪XD\mathcal{O}_{X\setminus D}-module \mathcal{F} is defined to a locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule 1ι()\mathcal{F}_{1}\subset\iota_{\ast}(\mathcal{F}) such that ι(1)=\iota^{\ast}(\mathcal{F}_{1})=\mathcal{F}. We obtain the following proposition from Proposition 4.1.

Proposition 4.10

There exists a unique meromorphic extension E~\widetilde{E} of EE such that ι(θ)(E~)E~ΩX1\iota_{\ast}(\theta)(\widetilde{E})\subset\widetilde{E}\otimes\Omega^{1}_{X} and ι(C)(E~E~)𝒪X(D)\iota_{\ast}(C)(\widetilde{E}\otimes\widetilde{E})\subset\mathcal{O}_{X}(\ast D). Such E~\widetilde{E} is called the canonical meromorphic extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC.  

Let θ~:E~E~ΩX1\widetilde{\theta}:\widetilde{E}\to\widetilde{E}\otimes\Omega^{1}_{X} and C~:E~E~𝒪X(D)\widetilde{C}:\widetilde{E}\otimes\widetilde{E}\to\mathcal{O}_{X}(\ast D) denote the induced morphisms.

4.2.2 Good filtered extensions

Definition 4.11

A good filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC is a filtered bundle 𝒫E~\mathcal{P}_{\ast}\widetilde{E} over E~\widetilde{E} such that (i) (𝒫E~,θ~)(\mathcal{P}_{\ast}\widetilde{E},\widetilde{\theta}) is a good filtered Higgs bundle of degree 0, (ii) C~\widetilde{C} is a symmetric pairing of 𝒫E~\mathcal{P}_{\ast}\widetilde{E}.  

In this definition, we do not assume that (𝒫E~,θ~)(\mathcal{P}_{\ast}\widetilde{E},\widetilde{\theta}) is polystable nor that C~\widetilde{C} is perfect. Note that if C~\widetilde{C} is perfect the condition deg(𝒫E~)=0\deg(\mathcal{P}_{\ast}\widetilde{E})=0 is automatically satisfied, as remarked in Lemma 3.10.

Theorem 4.12

The following holds.

  • (𝒫E~,θ~)(\mathcal{P}_{\ast}\widetilde{E},\widetilde{\theta}) is polystable, and C~\widetilde{C} is perfect.

  • Let E~1E~\widetilde{E}_{1}\subset\widetilde{E} be a saturated locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule such that (i) θ~(E~1)E~1ΩX1\widetilde{\theta}(\widetilde{E}_{1})\subset\widetilde{E}_{1}\otimes\Omega^{1}_{X}, (ii) deg(𝒫E~1)=0\deg(\mathcal{P}_{\ast}\widetilde{E}_{1})=0. Then, the restriction of C~\widetilde{C} to 𝒫E~1\mathcal{P}_{\ast}\widetilde{E}_{1} is also perfect, and we obtain the decomposition of good filtered Higgs bundles

    (𝒫E~,θ~)=(𝒫E~1,θ~1)(𝒫E~1C,θ~1C).(\mathcal{P}_{\ast}\widetilde{E},\widetilde{\theta})=(\mathcal{P}_{\ast}\widetilde{E}_{1},\widetilde{\theta}_{1})\oplus(\mathcal{P}_{\ast}\widetilde{E}_{1}^{\bot\,C},\widetilde{\theta}_{1}^{\bot\,C}). (24)

Proof   Let E~1E~\widetilde{E}_{1}\subset\widetilde{E} be a saturated locally free 𝒪X(D)\mathcal{O}_{X}(\ast D)-submodule such that θ(E~1)E~1\theta(\widetilde{E}_{1})\subset\widetilde{E}_{1}. We obtain the induced pairing C~1:𝒫(E~1)𝒫(E~1)𝒫(0)(𝒪X(D))\widetilde{C}_{1}:\mathcal{P}_{\ast}(\widetilde{E}_{1})\otimes\mathcal{P}_{\ast}(\widetilde{E}_{1})\to\mathcal{P}_{\ast}^{(0)}\bigl{(}\mathcal{O}_{X}(\ast D)\bigr{)}.

There exists a finite subset Z2XDZ_{2}\subset X\setminus D such that (E,¯E,θ)|X(DZ2)(E,\overline{\partial}_{E},\theta)_{|X\setminus(D\cup Z_{2})} is regular semisimple, and that C|X(DZ2)C_{|X\setminus(D\cup Z_{2})} is non-degenerate.

Lemma 4.13

C1|X(DZ2)C_{1|X\setminus(D\cup Z_{2})} is a non-degenerate symmetric pairing of E~1|X(DZ2)\widetilde{E}_{1|X\setminus(D\cup Z_{2})}.

Proof   Let QQ be any point of X(DZ2)X\setminus(D\cup Z_{2}). Let (XQ,z)(X_{Q},z) be a holomorphic coordinate neighbourhood around QQ. Let ff be the endomorphism of E~|XQ\widetilde{E}_{|X_{Q}} defined by θ=fdz\theta=f\,dz. There exists the eigen decomposition E|Q=αEQ,αE_{|Q}=\bigoplus_{\alpha\in{\mathbb{C}}}E_{Q,\alpha} of f|Qf_{|Q}. The decomposition is orthogonal with respect to C|QC_{|Q}. Because CC is non-degenerate, the restriction of CC to each EQ,αE_{Q,\alpha} is also non-degenerate. Because E~1|Q\widetilde{E}_{1|Q} is a direct sum of some eigen spaces, C~1|Q\widetilde{C}_{1|Q} is non-degenerate.  

We set r1:=rankE~1r_{1}:=\mathop{\rm rank}\nolimits\widetilde{E}_{1}. We obtain det(E~1)r1E~\det(\widetilde{E}_{1})\subset\bigwedge^{r_{1}}\widetilde{E}, and the filtered bundle 𝒫det(E~1)\mathcal{P}_{\ast}\det(\widetilde{E}_{1}). We obtain the induced pairing

det(C1):𝒫det(E~1)𝒫det(E~1)𝒫(0)(𝒪X(D)).\det(C_{1}):\mathcal{P}_{\ast}\det(\widetilde{E}_{1})\otimes\mathcal{P}_{\ast}\det(\widetilde{E}_{1})\longrightarrow\mathcal{P}^{(0)}_{\ast}\bigl{(}\mathcal{O}_{X}(\ast D)\bigr{)}. (25)

It is non-zero by Lemma 4.13. Hence, by Proposition 3.9, we obtain deg(𝒫E~1)0\deg(\mathcal{P}_{\ast}\widetilde{E}_{1})\leq 0. If moreover deg(𝒫E~1)=0\deg(\mathcal{P}_{\ast}\widetilde{E}_{1})=0 holds, then C~1\widetilde{C}_{1} is perfect. In particular, C~\widetilde{C} is perfect.

If E~1E~\widetilde{E}_{1}\neq\widetilde{E}, we obtain the decomposition (24) by Lemma 3.6. By an easy induction, we obtain the polystability of (𝒫E~,θ~)(\mathcal{P}_{\ast}\widetilde{E},\widetilde{\theta}), and Theorem 4.12 is proved.  

There exists a filtered bundle 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}) over E~\widetilde{E} such that for each PDP\in D the restriction of 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}) to a neighbourhood of PP is equal to the canonical filtered extension in Lemma 4.4.

Lemma 4.14

We have deg(𝒫canE~)=0\deg(\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}\widetilde{E})=0. Hence, 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}) is a good filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta).

Proof   By the construction, C~\widetilde{C} induces an isomorphism det(𝒫canE~)det(𝒫canE~)𝒫(0)(𝒪X(D))\det(\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}\widetilde{E})\otimes\det(\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}\widetilde{E})\simeq\mathcal{P}^{(0)}_{\ast}(\mathcal{O}_{X}(\ast D)). Hence, we obtain deg(𝒫canE~)=0\deg(\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}\widetilde{E})=0.  

4.2.3 Classification and uniqueness

Theorem 4.15

There exists the bijection between the following objects. The correspondence is induced by h𝒫h(E)h\mapsto\mathcal{P}^{h}_{\ast}(E) as in Proposition 4.8.

  • Harmonic metrics of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC.

  • Good filtered extensions of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC.

Proof   Let 𝒫E~\mathcal{P}_{\ast}\widetilde{E} be a good filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC. Note that (𝒫E~,θ)(\mathcal{P}_{\ast}\widetilde{E},\theta) is polystable of degree 0. There exists a decomposition

(𝒫E~,θ)=i=1m𝒫(𝒱i,θi)(\mathcal{P}_{\ast}\widetilde{E},\theta)=\bigoplus_{i=1}^{m}\mathcal{P}_{\ast}(\mathcal{V}_{i},\theta_{i})

into stable good filtered Higgs bundles of degree 0. Because (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is generically regular semisimple, we have 𝒫(𝒱i,θi)≄𝒫(𝒱j,θj)\mathcal{P}_{\ast}(\mathcal{V}_{i},\theta_{i})\not\simeq\mathcal{P}_{\ast}(\mathcal{V}_{j},\theta_{j}) unless i=ji=j. Because the spectral curves of (𝒱i,θi)(\mathcal{V}_{i},\theta_{i}) and (𝒱i,θi)(\mathcal{V}_{i}^{\lor},\theta_{i}^{\lor}) are the same, we obtain that 𝒫(𝒱i,θi)≄𝒫(𝒱j,θj)\mathcal{P}_{\ast}(\mathcal{V}^{\lor}_{i},\theta^{\lor}_{i})\not\simeq\mathcal{P}_{\ast}(\mathcal{V}_{j},\theta_{j}) unless i=ji=j. By Proposition 3.26, each (𝒫𝒱i,θi)(\mathcal{P}_{\ast}\mathcal{V}_{i},\theta_{i}) has a symmetric perfect pairing PiP_{i}, and we may assume that C~\widetilde{C} is the direct sum of PiP_{i}. By Proposition 3.27 together with Lemma 3.22, there exists a unique harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC.  

Theorem 4.16

If moreover (E,¯E,θ)(E,\overline{\partial}_{E},\theta) is regular semisimple at each point of DD in the sense of Definition 4.5, there exists a unique harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC. In this case, 𝒫h(E)\mathcal{P}^{h}_{\ast}(E) is equal to 𝒫can(E~)\mathcal{P}^{\mathop{\rm can}\nolimits}_{\ast}(\widetilde{E}).

Proof   It follows from Proposition 4.6 and Theorem 4.15.  

4.2.4 Complement

Let (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) be a good filtered Higgs bundle of degree 0 on (X,D)(X,D) with a symmetric pairing CC. Let (E,¯E,θ)(E,\overline{\partial}_{E},\theta) be the Higgs bundle on XDX\setminus D obtained as the restriction of (𝒱,θ)(\mathcal{V},\theta). By applying the argument in the proof of Theorem 4.12, we obtain the following theorem.

Theorem 4.17

Assume that there exists QXDQ\in X\setminus D such that the following holds.

  • C|QC_{|Q} is non-degenerate, and θ\theta is regular semisimple around QQ.

Then, the following holds.

  • (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) is polystable, and CC is perfect. In particular, C|XDC_{|X\setminus D} is non-degenerate.

  • 𝒫𝒱\mathcal{P}_{\ast}\mathcal{V} is a good filtered extension of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) with CC.

As a result, there exists a unique harmonic metric hh of (E,¯E,θ)(E,\overline{\partial}_{E},\theta) compatible with CC such that 𝒫hE=𝒫𝒱\mathcal{P}^{h}_{\ast}E=\mathcal{P}_{\ast}\mathcal{V}.  

4.3 Examples

Let XX be any Riemann surface. We set KX=ΩX1K_{X}=\Omega_{X}^{1}. For r>1r\in{\mathbb{Z}}_{>1}, we set 𝕂X,r:=i=1rKX(r2i+1)/2\mathbb{K}_{X,r}:=\bigoplus_{i=1}^{r}K_{X}^{(r-2i+1)/2}. Let qjq_{j} (j=2,,r)(j=2,\ldots,r) be holomorphic jj-differentials on XX. The multiplication of qjq_{j} induces

KX(r2i+1)/2KX(r2i+2(j1)+1)/2KX(jir).K_{X}^{(r-2i+1)/2}\to K_{X}^{(r-2i+2(j-1)+1)/2}\otimes K_{X}\quad(j\leq i\leq r).

We also have the multiplications of i(ri)/2i(r-i)/2 for i=1,,r1i=1,\ldots,r-1:

KX(r2i+1)/2KX(r2(i+1)+1)/2KX.K_{X}^{(r-2i+1)/2}\to K_{X}^{(r-2(i+1)+1)/2}\otimes K_{X}.

They define a Higgs field θ(𝒒)\theta(\boldsymbol{q}) of 𝕂X,r\mathbb{K}_{X,r}. The natural pairing KX(r2i+1)/2KX(r2i+1)/2𝒪XK_{X}^{(r-2i+1)/2}\otimes K_{X}^{-(r-2i+1)/2}\to\mathcal{O}_{X} induces a non-degenerate symmetric bilinear form C𝕂,X,rC_{\mathbb{K},X,r} of 𝕂X,r\mathbb{K}_{X,r}. It is a symmetric pairing of (𝕂X,r,θ(𝒒))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})). We have the following corollary of Theorem 2.34.

Corollary 4.18

If the Higgs bundle (𝕂X,r,θ(𝐪))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})) is generically regular semisimple, then there exists a harmonic metric of (𝕂X,r,θ(𝐪))(\mathbb{K}_{X,r},\theta(\boldsymbol{q})) which is compatible with C𝕂,X,rC_{\mathbb{K},X,r}. It is induced by an SL(r,)\mathop{\rm SL}\nolimits(r,{\mathbb{R}})-harmonic bundle.  

In some case, we can obtain a precise classification by using Theorem 4.15 and Theorem 4.16. Let us explain the case r=3r=3 and X=X={\mathbb{C}}.

4.3.1 Example 1

Let αi[z]\alpha_{i}\in{\mathbb{C}}[z] be polynomials such that α1+α2+α3=0\alpha_{1}+\alpha_{2}+\alpha_{3}=0 and that αiαj\alpha_{i}-\alpha_{j} (ij)(i\neq j) are not constantly 0. We set

q2=12(α1α2+α2α3+α3α1)(dz)2,q3=α1α2α3(dz)3.q_{2}=-\frac{1}{2}(\alpha_{1}\alpha_{2}+\alpha_{2}\alpha_{3}+\alpha_{3}\alpha_{1})(dz)^{2},\quad q_{3}=\alpha_{1}\alpha_{2}\alpha_{3}(dz)^{3}.

Let ff be the endomorphism of 𝕂,3\mathbb{K}_{{\mathbb{C}},3} defined by θ(q2,q3)=fdz\theta(q_{2},q_{3})=f\,dz. Because the eigenvalues of ff are αi\alpha_{i} (i=1,2,3)(i=1,2,3), the Higgs bundle is generically semisimple.

Proposition 4.19

In this case, (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) is regular semisimple at \infty. As a result, it has a unique harmonic metric hh compatible with C𝕂,,3C_{\mathbb{K},{\mathbb{C}},3}.

Proof   We set w=z1w=z^{-1}. Around w=0w=0, let gg be the endomorphism defined by gdw/w=θ(q2,q3)g\,dw/w=\theta(q_{2},q_{3}). Because g=w1fg=-w^{-1}f, the eigenvalues of gg are w1αi(w1)-w^{-1}\alpha_{i}(w^{-1}). Hence, (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) is regular semisimple at \infty. We obtain the uniqueness of compatible harmonic metric from Theorem 4.16.  

Because 𝒫h𝕂,3\mathcal{P}^{h}_{\ast}\mathbb{K}_{{\mathbb{C}},3} is equal to the canonical filtered extension, we can study the asymptotic behaviour of hh more closely by using a general theory of wild harmonic bundles [12].

4.3.2 Example 2

Let β\beta be a non-constant polynomial. We set

q2:=12(143β2)(dz)2,q3:=(23β1627β3)(dz)3.q_{2}:=-\frac{1}{2}\Bigl{(}1-\frac{4}{3}\beta^{2}\Bigr{)}(dz)^{2},\quad\quad q_{3}:=-\Bigl{(}\frac{2}{3}\beta-\frac{16}{27}\beta^{3}\Bigr{)}(dz)^{3}.

On a neighbourhood of \infty, we set

α1=13β+β(1β2)1/2,α2=13ββ(1β2)1/2,α3=23β.\alpha_{1}=\frac{1}{3}\beta+\beta(1-\beta^{-2})^{1/2},\quad\alpha_{2}=\frac{1}{3}\beta-\beta(1-\beta^{-2})^{1/2},\quad\alpha_{3}=-\frac{2}{3}\beta.

Then, we have

T3+(143β2)T+23β1627β3=(Tα1)(Tα2)(Tα3).T^{3}+\Bigl{(}1-\frac{4}{3}\beta^{2}\Bigr{)}T+\frac{2}{3}\beta-\frac{16}{27}\beta^{3}=(T-\alpha_{1})(T-\alpha_{2})(T-\alpha_{3}).

Note that

α2α3=ββ(1β2)1/2=12β1+O(β2).\alpha_{2}-\alpha_{3}=\beta-\beta(1-\beta^{-2})^{1/2}=\frac{1}{2}\beta^{-1}+O(\beta^{-2}).

Set w=z1w=z^{-1}. Let gg be the endomorphism of 𝕂,3|{0}\mathbb{K}_{{\mathbb{C}},3|{\mathbb{C}}\setminus\{0\}} defined by θ(q2,q2)=g(dw/w)\theta(q_{2},q_{2})=g(dw/w). The eigenvalues of gg are w1αi(w1)-w^{-1}\alpha_{i}(w^{-1}).

Lemma 4.20

If deg(β)=1\deg(\beta)=1, then (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) is regular semisimple at \infty. If deg(β)2\deg(\beta)\geq 2, (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) is not regular semisimple at \infty.  

If deg(β)=1\deg(\beta)=1, (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) has a unique harmonic metric by Theorem 4.16. In the case deg(β)=2\deg(\beta)=2, we can classify good filtered extensions of (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) by using Proposition 4.26 below, which implies the classification of harmonic metrics of (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) by Theorem 4.15. We set

q~2:=12(143β2),vi=(αi2q~2αi1).\widetilde{q}_{2}:=-\frac{1}{2}\Bigl{(}1-\frac{4}{3}\beta^{2}\Bigr{)},\quad v_{i}=\left(\begin{array}[]{c}\alpha_{i}^{2}-\widetilde{q}_{2}\\ \alpha_{i}\\ 1\end{array}\right).

We have g(vi)=w1αivig(v_{i})=-w^{-1}\alpha_{i}\cdot v_{i}. We also have C(v2,v2)=2+2β22β2(1β2)1/2C(v_{2},v_{2})=-2+2\beta^{2}-2\beta^{2}(1-\beta^{-2})^{1/2} and C(v3,v3)=1C(v_{3},v_{3})=1. For any good filtered extension 𝒫𝕂~,3\mathcal{P}_{\ast}\widetilde{\mathbb{K}}_{{\mathbb{C}},3}, there exists the decomposition of filtered bundles

𝒫𝕂~,3=𝒫(𝒪U()v1)𝒫(𝒪U()v2𝒪U()v3).\mathcal{P}_{\ast}\widetilde{\mathbb{K}}_{{\mathbb{C}},3}=\mathcal{P}_{\ast}(\mathcal{O}_{U_{\infty}}(\ast\infty)v_{1})\oplus\mathcal{P}_{\ast}\Bigl{(}\mathcal{O}_{U_{\infty}}(\ast\infty)v_{2}\oplus\mathcal{O}_{U_{\infty}}(\ast\infty)v_{3}\Bigr{)}.

By an appropriate normalization, 𝒪U()v2𝒪U()v2\mathcal{O}_{U_{\infty}}(\ast\infty)v_{2}\oplus\mathcal{O}_{U_{\infty}}(\ast\infty)v_{2} with the induced Higgs field and the symmetric product is isomorphic to the Higgs bundle (𝒱,θ)(\mathcal{V},\theta) with the symmetric product C0,0C_{0,0} in §4.4. Hence, there are good filtered extensions of types (I), (II) and (III-1) as in Proposition 4.26. In particular, the uniqueness of harmonic metrics of (𝕂,3,θ(q2,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2},q_{3})) does not hold in this case.

4.3.3 Example 3

For aa\in{\mathbb{C}}, we set q2,a=az2sin(z)(dz)2q_{2,a}=az^{2}\sin(z)(dz)^{2} and q3=(z+1)4cos(z)(dz)3q_{3}=(z+1)^{4}\cos(z)(dz)^{3}. Because (z2sin(z))|z=0=0(z^{2}\sin(z))_{|z=0}=0 and ((z+1)4cos(z))|z=0=1\bigl{(}(z+1)^{4}\cos(z)\bigr{)}_{|z=0}=1, it is easy to check that (𝕂,3,θ(q2,a,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2,a},q_{3})) is generically regular semisimple. Hence, there exists a harmonic metric of (𝕂,3,θ(q2,a,q3))(\mathbb{K}_{{\mathbb{C}},3},\theta(q_{2,a},q_{3})) compatible with C𝕂,,3C_{\mathbb{K},{\mathbb{C}},3}.

Remark 4.21

If a=0a=0, we can also prove the uniqueness of such a harmonic metric by using a result in [10].  

4.4 Appendix: Classification of regular filtered extensions in an easy case

Let UU be a neighbourhood of 0 in {\mathbb{C}}. We set 𝒱=𝒪U({0})e1𝒪U({0})e2\mathcal{V}=\mathcal{O}_{U}(\ast\{0\})e_{1}\oplus\mathcal{O}_{U}(\ast\{0\})e_{2}. We consider the Higgs field θ\theta of 𝒱\mathcal{V} given by

θe1=e1dz,θe2=e2(1)dz.\theta e_{1}=e_{1}\,\,dz,\quad\theta e_{2}=e_{2}\,\,(-1)dz.

For 𝒎=(m1,m2){0,1}2{\boldsymbol{m}}=(m_{1},m_{2})\in\{0,1\}^{2}, let C𝒎C_{{\boldsymbol{m}}} be the symmetric pairing of (𝒱,θ)(\mathcal{V},\theta) determined by C𝒎(ei,ei)=zmiC_{{\boldsymbol{m}}}(e_{i},e_{i})=z^{m_{i}} (i=1,2)(i=1,2) and C𝒎(e1,e2)=0C_{{\boldsymbol{m}}}(e_{1},e_{2})=0.

4.4.1 Logarithmic lattices

For 𝒏=(n1,n2)2{\boldsymbol{n}}=(n_{1},n_{2})\in{\mathbb{Z}}^{2}, we set

𝒱(𝒏):=𝒪Uzn1e1𝒪Uzn2e2𝒱.\mathcal{V}^{({\boldsymbol{n}})}:=\mathcal{O}_{U}z^{-n_{1}}e_{1}\oplus\mathcal{O}_{U}z^{-n_{2}}e_{2}\subset\mathcal{V}.

For 𝜶=(α1,α2)2{(0,0)}{\boldsymbol{\alpha}}=(\alpha_{1},\alpha_{2})\in{\mathbb{C}}^{2}\setminus\{(0,0)\} and 𝒏2{\boldsymbol{n}}\in{\mathbb{Z}}^{2}, we set

𝒱(𝒏,𝜶):=z𝒱(𝒏)+𝒪U(α1zn1e1+α2zn2e2)𝒱(𝒏).\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}:=z\mathcal{V}^{({\boldsymbol{n}})}+\mathcal{O}_{U}\cdot\bigl{(}\alpha_{1}z^{-n_{1}}e_{1}+\alpha_{2}z^{-n_{2}}e_{2}\bigr{)}\subset\mathcal{V}^{({\boldsymbol{n}})}.

Note that 𝒱(𝒏,(α1,0))=𝒱(n1,n21)\mathcal{V}^{({\boldsymbol{n}},(\alpha_{1},0))}=\mathcal{V}^{(n_{1},n_{2}-1)} and 𝒱(𝒏,(0,α2))=𝒱(n11,n2)\mathcal{V}^{({\boldsymbol{n}},(0,\alpha_{2}))}=\mathcal{V}^{(n_{1}-1,n_{2})}. We also note that 𝒱(𝒏,𝜶)=𝒱(𝒏,γ𝜶)\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}=\mathcal{V}^{({\boldsymbol{n}},\gamma{\boldsymbol{\alpha}})} for γ0\gamma\neq 0.

A lattice of 𝒱\mathcal{V} means a locally free 𝒪U\mathcal{O}_{U}-submodule 𝒰𝒱\mathcal{U}\subset\mathcal{V} such that 𝒰(0)=𝒱\mathcal{U}(\ast 0)=\mathcal{V}. It is called logarithmic if θ(𝒰)𝒰ΩU1(log0)\theta(\mathcal{U})\subset\mathcal{U}\otimes\Omega^{1}_{U}(\log 0). If 𝒰\mathcal{U} is a logarithmic lattice, we obtain the endomorphism Res𝒰(θ)\mathop{\rm Res}\nolimits_{\mathcal{U}}(\theta) of 𝒰|0=𝒰/z𝒰\mathcal{U}_{|0}=\mathcal{U}/z\mathcal{U}.

Both 𝒱(𝒏)\mathcal{V}^{({\boldsymbol{n}})} and 𝒱(𝒏,𝜶)\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})} are logarithmic. We have Res𝒱(𝒏)(θ)=0\mathop{\rm Res}\nolimits_{\mathcal{V}^{({\boldsymbol{n}})}}(\theta)=0. If 𝜶()2{\boldsymbol{\alpha}}\in({\mathbb{C}}^{\ast})^{2}, we also have Res𝒱(𝒏,𝜶)(θ)0\mathop{\rm Res}\nolimits_{\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}}(\theta)\neq 0 and Res𝒱(𝒏,𝜶)(θ)2=0\mathop{\rm Res}\nolimits_{\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}}(\theta)^{2}=0.

Lemma 4.22

Let 𝒰𝒱\mathcal{U}\subset\mathcal{V} be a logarithmic lattice. Then, we have 𝒰=𝒱(𝐧)\mathcal{U}=\mathcal{V}^{({\boldsymbol{n}})} for some 𝐧2{\boldsymbol{n}}\in{\mathbb{Z}}^{2} or 𝒰=𝒱(𝐧,𝛂)\mathcal{U}=\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})} for some (𝐧,𝛂)2×((2){(0,0)})({\boldsymbol{n}},{\boldsymbol{\alpha}})\in{\mathbb{Z}}^{2}\times\bigl{(}({\mathbb{C}}^{2})\setminus\{(0,0)\}\bigr{)}.

Proof   Let LiL_{i} (i=1,2)(i=1,2) be the image of 𝒰\mathcal{U} by the projection to 𝒪U(0)ei\mathcal{O}_{U}(\ast 0)\cdot e_{i}. There exists 𝒏2{\boldsymbol{n}}\in{\mathbb{Z}}^{2} such that Li=𝒪UznieiL_{i}=\mathcal{O}_{U}\cdot z^{-n_{i}}e_{i}. There exists 𝜶()2{\boldsymbol{\alpha}}\in({\mathbb{C}}^{\ast})^{2} and a holomorphic function β\beta on UU such that (i) β(0)0\beta(0)\neq 0, (ii) α1zn1e1+α2zn2βe2\alpha_{1}z^{-n_{1}}e_{1}+\alpha_{2}z^{-n_{2}}\beta e_{2} is contained in 𝒰\mathcal{U}. Because θ\theta is logarithmic with respect to 𝒰\mathcal{U}, α1zn1+1e1α2zn2+1βe2\alpha_{1}z^{-n_{1}+1}e_{1}-\alpha_{2}z^{-n_{2}+1}\beta e_{2} is also contained in 𝒰\mathcal{U}. We obtain z𝒱(𝒏)𝒰z\mathcal{V}^{({\boldsymbol{n}})}\subset\mathcal{U}, and hence 𝒱(𝒏,𝜶)𝒰𝒱(𝒏)\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}\subset\mathcal{U}\subset\mathcal{V}^{({\boldsymbol{n}})}. It implies 𝒰=𝒱(𝒏,𝜶),𝒱(𝒏)\mathcal{U}=\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})},\mathcal{V}^{({\boldsymbol{n}})}.  

Lemma 4.23

Let 𝒰\mathcal{U} be a logarithmic lattice. If z𝒱(𝐧)𝒰𝒱(𝐧)z\mathcal{V}^{({\boldsymbol{n}})}\subsetneq\mathcal{U}\subsetneq\mathcal{V}^{({\boldsymbol{n}})} for some 𝐧{\boldsymbol{n}}, then there exists 𝛂2{(0,0)}{\boldsymbol{\alpha}}\in{\mathbb{C}}^{2}\setminus\{(0,0)\} such that 𝒰=𝒱(𝐧,𝛂)\mathcal{U}=\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}. If z𝒱(𝐧,𝛂)𝒰𝒱(𝐧,𝛂)z\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}\subsetneq\mathcal{U}\subsetneq\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})} for some (𝐧,𝛂)2×()2({\boldsymbol{n}},{\boldsymbol{\alpha}})\in{\mathbb{Z}}^{2}\times({\mathbb{C}}^{\ast})^{2}, then we have 𝒰=z𝒱(𝐧)\mathcal{U}=z\mathcal{V}^{({\boldsymbol{n}})}.

Proof   The first claim is clear. Let us study the second claim. Let L𝒱|0(𝒏,𝜶)L\subset\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}_{|0} be the image of 𝒰\mathcal{U}. Because 𝒰\mathcal{U} is a logarithmic lattice, LL is preserved by Res𝒱(𝒏,𝜶)(θ)\mathop{\rm Res}\nolimits_{\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}}(\theta). Then, the second claim follows.  

4.4.2 Regular filtered Higgs bundles

Let (𝒏,𝜶)2×(2{(0,0)})({\boldsymbol{n}},{\boldsymbol{\alpha}})\in{\mathbb{Z}}^{2}\times({\mathbb{C}}^{2}\setminus\{(0,0)\}). For bb\in{\mathbb{R}}, let 𝒫(𝒏;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}};b)}_{\ast}(\mathcal{V}) and 𝒫(𝒏,𝜶;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V}) be the filtered bundles over 𝒱\mathcal{V} defined as follows (c)(c\in{\mathbb{R}}):

𝒫c(𝒏;b)(𝒱)=z[cb]𝒱(𝒏),𝒫c(𝒏,𝜶;b)(𝒱)=z[cb]𝒱(𝒏,𝜶).\mathcal{P}^{({\boldsymbol{n}};b)}_{c}(\mathcal{V})=z^{-[c-b]}\mathcal{V}^{({\boldsymbol{n}})},\quad\quad\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{c}(\mathcal{V})=z^{-[c-b]}\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}.

Here, [d]:=max{n|nd}[d]:=\max\{n\in{\mathbb{Z}}\,|\,n\leq d\}. For 𝒃=(b1,b2)2{\boldsymbol{b}}=(b_{1},b_{2})\in{\mathbb{R}}^{2} such that b11<b2<b1b_{1}-1<b_{2}<b_{1}, let 𝒫(𝒏,𝜶;𝒃)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}) be the filtered bundle defined as follows (cc\in{\mathbb{R}}, mm\in{\mathbb{Z}}):

𝒫c(𝒏,𝜶;𝒃)(𝒱)={zm𝒱(𝒏,𝜶)(m+b2c<m+b1)zm𝒱(𝒏)(m+b1c<m+1+b2).\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{c}(\mathcal{V})=\left\{\begin{array}[]{ll}z^{-m}\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}&(m+b_{2}\leq c<m+b_{1})\\ \\ z^{-m}\mathcal{V}^{({\boldsymbol{n}})}&(m+b_{1}\leq c<m+1+b_{2}).\end{array}\right.

We set 𝜹=(1,1){\boldsymbol{\delta}}=(1,1). We can check the following lemma directly from the definitions.

Lemma 4.24

The following holds for any mm\in{\mathbb{Z}}.

𝒫(𝒏;b)(𝒱)=𝒫(𝒏m𝜹;bm)(𝒱),𝒫(𝒏,𝜶;b)(𝒱)=𝒫(𝒏m𝜹,𝜶;bm)(𝒱),𝒫(𝒏,𝜶;𝒃)(𝒱)=𝒫(𝒏m𝜹,𝜶;𝒃m𝜹).\mathcal{P}^{({\boldsymbol{n}};b)}_{\ast}(\mathcal{V})=\mathcal{P}^{({\boldsymbol{n}}-m{\boldsymbol{\delta}};b-m)}_{\ast}(\mathcal{V}),\quad\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V})=\mathcal{P}^{({\boldsymbol{n}}-m{\boldsymbol{\delta}},{\boldsymbol{\alpha}};b-m)}_{\ast}(\mathcal{V}),\quad\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V})=\mathcal{P}^{({\boldsymbol{n}}-m{\boldsymbol{\delta}},{\boldsymbol{\alpha}};{\boldsymbol{b}}-m{\boldsymbol{\delta}})}.

The following holds:

𝒫((n1,n2),(α1,0);b)(𝒱)=𝒫((n1,n21);b)(𝒱),𝒫((n1,n2),(0,α2);b)(𝒱)=𝒫((n11,n2);b)(𝒱).\mathcal{P}^{((n_{1},n_{2}),(\alpha_{1},0);b)}_{\ast}(\mathcal{V})=\mathcal{P}^{((n_{1},n_{2}-1);b)}_{\ast}(\mathcal{V}),\quad\mathcal{P}^{((n_{1},n_{2}),(0,\alpha_{2});b)}_{\ast}(\mathcal{V})=\mathcal{P}^{((n_{1}-1,n_{2});b)}_{\ast}(\mathcal{V}).

The following holds:

𝒫((n1,n2),(α1,0);(b1,b2))(𝒱)=𝒫((n1,n21),(0,α2);(b2,b11))(𝒱).\mathcal{P}^{((n_{1},n_{2}),(\alpha_{1},0);(b_{1},b_{2}))}_{\ast}(\mathcal{V})=\mathcal{P}^{((n_{1},n_{2}-1),(0,\alpha_{2});(b_{2},b_{1}-1))}_{\ast}(\mathcal{V}).

We also have 𝒫(𝐧,𝛂;b)(𝒱)=𝒫(𝐧,γ𝛂;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V})=\mathcal{P}^{({\boldsymbol{n}},\gamma{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V}) and 𝒫(𝐧,𝛂;𝐛)(𝒱)=𝒫(𝐧,γ𝛂;𝐛)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V})=\mathcal{P}^{({\boldsymbol{n}},\gamma{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}) for any γ\gamma\in{\mathbb{C}}^{\ast}.  

Proposition 4.25

Let 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) be a filtered bundle such that (𝒫𝒱,θ)(\mathcal{P}_{\ast}\mathcal{V},\theta) is a regular filtered Higgs bundle. Then, 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) equals one of 𝒫(𝐧;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}};b)}_{\ast}(\mathcal{V}), 𝒫(𝐧,𝛂;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V}) or 𝒫(𝐧,𝛂;𝐛)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}).

Proof   If there exists 1<b0-1<b\leq 0 such that Grc𝒫(𝒱)=0\mathop{\rm Gr}\nolimits^{\mathcal{P}}_{c}(\mathcal{V})=0 unless cbc-b\in{\mathbb{Z}}, 𝒫(𝒱)\mathcal{P}_{\ast}(\mathcal{V}) equals either 𝒫(𝒏;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}};b)}_{\ast}(\mathcal{V}) or 𝒫(𝒏,𝜶;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V}). If not, there exists 𝒃2{\boldsymbol{b}}\in{\mathbb{R}}^{2} such that (i) b11<b2<b1b_{1}-1<b_{2}<b_{1}, (ii) Grc𝒫(𝒱)=0\mathop{\rm Gr}\nolimits^{\mathcal{P}}_{c}(\mathcal{V})=0 if cbic-b_{i}\not\in{\mathbb{Z}} (i=1,2)(i=1,2). By Lemma 4.23, one of 𝒫bi(𝒱)\mathcal{P}_{b_{i}}(\mathcal{V}) is of the form 𝒱(𝒏)\mathcal{V}^{({\boldsymbol{n}})} for some 𝒏2{\boldsymbol{n}}\in{\mathbb{Z}}^{2}. By exchanging (b1,b2)(b_{1},b_{2}) with (b2,b11)(b_{2},b_{1}-1) if necessary, we may assume that 𝒫b1(𝒱)=𝒱(𝒏)\mathcal{P}_{b_{1}}(\mathcal{V})=\mathcal{V}^{({\boldsymbol{n}})}. We obtain 𝒫b2(𝒗)=𝒱(𝒏,𝜶)\mathcal{P}_{b_{2}}({\boldsymbol{v}})=\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})} for some 𝜶2{(0,0)}{\boldsymbol{\alpha}}\in{\mathbb{C}}^{2}\setminus\{(0,0)\}. Then, we obtain 𝒫(𝒱)=𝒫(𝒏,𝜶;𝒃)(𝒱)\mathcal{P}_{\ast}(\mathcal{V})=\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}).  

4.4.3 Compatibility with the symmetric form

Proposition 4.26

Let 𝐧2{\boldsymbol{n}}\in{\mathbb{Z}}^{2} and bb\in{\mathbb{R}}.

(I)

𝒫(𝒏;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}};b)}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if n1m12=n2m22=bn_{1}-\frac{m_{1}}{2}=n_{2}-\frac{m_{2}}{2}=b.

(II)

If 𝜶()2{\boldsymbol{\alpha}}\in({\mathbb{C}}^{\ast})^{2}, 𝒫(𝒏,𝜶;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if n112m1=n212m2=b+12n_{1}-\frac{1}{2}m_{1}=n_{2}-\frac{1}{2}m_{2}=b+\frac{1}{2} and α12+α22=0\alpha_{1}^{2}+\alpha_{2}^{2}=0.

(III-1)

If 𝜶()2{\boldsymbol{\alpha}}\in({\mathbb{C}}^{\ast})^{2}, 𝒫(𝒏,𝜶;𝒃)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if n112m1=n212m2=12(b1+b2)n_{1}-\frac{1}{2}m_{1}=n_{2}-\frac{1}{2}m_{2}=\frac{1}{2}(b_{1}+b_{2}) and α12+α22=0\alpha_{1}^{2}+\alpha_{2}^{2}=0.

(III-2)

𝒫(𝒏,(0,α2);𝒃)(𝒱)\mathcal{P}^{({\boldsymbol{n}},(0,\alpha_{2});{\boldsymbol{b}})}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if n112m1=b1n_{1}-\frac{1}{2}m_{1}=b_{1} and n212m2=b2n_{2}-\frac{1}{2}m_{2}=b_{2}.

Proof   As a preliminary, let us study the dual lattices. Let e1,e2e_{1}^{\lor},e_{2}^{\lor} be the dual frame of 𝒱\mathcal{V}^{\lor}. For 𝒏2{\boldsymbol{n}}\in{\mathbb{Z}}^{2} and 𝜶2{(0,0)}{\boldsymbol{\alpha}}\in{\mathbb{C}}^{2}\setminus\{(0,0)\}, we set

(𝒱)(𝒏)=𝒪Uzn1e1𝒪Uzn2e2,(𝒱)(𝒏,𝜶)=z(𝒱)(𝒏)+𝒪U(α1zn1e1+α2zn2e2).(\mathcal{V}^{\lor})^{({\boldsymbol{n}})}=\mathcal{O}_{U}\cdot z^{-n_{1}}e_{1}^{\lor}\oplus\mathcal{O}_{U}\cdot z^{-n_{2}}e_{2}^{\lor},\quad\quad(\mathcal{V}^{\lor})^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}=z(\mathcal{V}^{\lor})^{({\boldsymbol{n}})}+\mathcal{O}_{U}\cdot\bigl{(}\alpha_{1}z^{-n_{1}}e_{1}^{\lor}+\alpha_{2}z^{-n_{2}}e_{2}^{\lor}\bigr{)}.

For any lattice 𝒰\mathcal{U} of 𝒱\mathcal{V}, we set 𝒰=om𝒪U(𝒰,𝒪U)\mathcal{U}^{\lor}={\mathcal{H}om}_{\mathcal{O}_{U}}(\mathcal{U},\mathcal{O}_{U}). We have (𝒱(𝒏))=(𝒱)(𝒏)(\mathcal{V}^{({\boldsymbol{n}})})^{\lor}=(\mathcal{V}^{\lor})^{(-{\boldsymbol{n}})}.

Lemma 4.27

For 𝛂2{(0,0)}{\boldsymbol{\alpha}}\in{\mathbb{C}}^{2}\setminus\{(0,0)\}, let 𝛃2{(0,0)}{\boldsymbol{\beta}}\in{\mathbb{C}}^{2}\setminus\{(0,0)\} such that α1β1+α2β2=0\alpha_{1}\beta_{1}+\alpha_{2}\beta_{2}=0. Then, we have

(𝒱(𝒏,𝜶))=(𝒱)(𝒏+𝜹,𝜷).(\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})})^{\lor}=(\mathcal{V}^{\lor})^{(-{\boldsymbol{n}}+{\boldsymbol{\delta}},{\boldsymbol{\beta}})}.

Proof   Because 𝒱(𝒏𝜹)𝒱(𝒏,𝜶)\mathcal{V}^{({\boldsymbol{n}}-{\boldsymbol{\delta}})}\subset\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})}, we obtain (𝒱(𝒏,𝜶))(𝒱(𝒏𝜹))=(𝒱)(𝒏+𝜹)(\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})})^{\lor}\subset(\mathcal{V}^{({\boldsymbol{n}}-{\boldsymbol{\delta}})})^{\lor}=(\mathcal{V}^{\lor})^{(-{\boldsymbol{n}}+{\boldsymbol{\delta}})}. Because

β1zn11e1+β2zn21e2,α1zn1e1+α2zn2e2=z1(α1β1+α2β2),\bigl{\langle}\beta_{1}z^{n_{1}-1}e_{1}^{\lor}+\beta_{2}z^{n_{2}-1}e_{2}^{\lor},\alpha_{1}z^{-n_{1}}e_{1}+\alpha_{2}z^{-n_{2}}e_{2}\bigr{\rangle}=z^{-1}\bigl{(}\alpha_{1}\beta_{1}+\alpha_{2}\beta_{2}\bigr{)},

β1zn11e1+β2zn21e2\beta_{1}z^{n_{1}-1}e_{1}^{\lor}+\beta_{2}z^{n_{2}-1}e_{2}^{\lor} is contained in (𝒱(𝒏,𝜶))(\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})})^{\lor} if and only if β1α1+β2α2=0\beta_{1}\alpha_{1}+\beta_{2}\alpha_{2}=0.  

Let us return to the proof of Proposition 4.26. Because ΨC𝒎(𝒱(𝒏))=(𝒱)(𝒏𝒎)\Psi_{C_{{\boldsymbol{m}}}}(\mathcal{V}^{({\boldsymbol{n}})})=(\mathcal{V}^{\lor})^{({\boldsymbol{n}}-{\boldsymbol{m}})}, 𝒫(𝒏;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}};b)}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if 2n1m1=2n2m2=2b2n_{1}-m_{1}=2n_{2}-m_{2}=2b, i.e., n112m1=n212m2=bn_{1}-\frac{1}{2}m_{1}=n_{2}-\frac{1}{2}m_{2}=b.

Let us consider the case of 𝒫(𝒏,𝜶;b)(𝒱)\mathcal{P}_{\ast}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};b)}(\mathcal{V}) for 𝜶()2{\boldsymbol{\alpha}}\in({\mathbb{C}}^{\ast})^{2}. Note that ΨC𝒎(𝒱(𝒏,𝜶))=(𝒱)(𝒏𝒎,𝜶)\Psi_{C_{{\boldsymbol{m}}}}(\mathcal{V}^{({\boldsymbol{n}},{\boldsymbol{\alpha}})})=(\mathcal{V}^{\lor})^{({\boldsymbol{n}}-{\boldsymbol{m}},{\boldsymbol{\alpha}})}. Because it cannot be (𝒱)(𝒑)(\mathcal{V}^{\lor})^{({\boldsymbol{p}})} for any 𝒑2{\boldsymbol{p}}\in{\mathbb{Z}}^{2}, it equals to (𝒱)(𝒏+𝜹,𝜷)(\mathcal{V}^{\lor})^{(-{\boldsymbol{n}}+\ell{\boldsymbol{\delta}},{\boldsymbol{\beta}})} for some \ell\in{\mathbb{Z}} and 𝜷()2{\boldsymbol{\beta}}\in({\mathbb{C}}^{\ast})^{2} such that α1β1+α2β2=0\alpha_{1}\beta_{1}+\alpha_{2}\beta_{2}=0 by Lemma 4.27. Hence, we obtain α12+α22=0\alpha_{1}^{2}+\alpha_{2}^{2}=0 and n112m1=n212m2n_{1}-\frac{1}{2}m_{1}=n_{2}-\frac{1}{2}m_{2}. Note the following equalities:

α1zn1e1+α2zn2e2,α1zn1+m1e1+α2zn2+m2e2=0,\bigl{\langle}\alpha_{1}z^{-n_{1}}e_{1}+\alpha_{2}z^{-n_{2}}e_{2},\,\,\alpha_{1}z^{-n_{1}+m_{1}}e_{1}^{\lor}+\alpha_{2}z^{-n_{2}+m_{2}}e_{2}^{\lor}\bigr{\rangle}=0,
α1zn1e1+α2zn2e2,α1zn1+1+m1e1α2zn2+1+m2e2=z2n1+1+m1(α12α22),\bigl{\langle}\alpha_{1}z^{-n_{1}}e_{1}+\alpha_{2}z^{-n_{2}}e_{2},\,\,\alpha_{1}z^{-n_{1}+1+m_{1}}e_{1}^{\lor}-\alpha_{2}z^{-n_{2}+1+m_{2}}e_{2}^{\lor}\bigr{\rangle}=z^{-2n_{1}+1+m_{1}}(\alpha_{1}^{2}-\alpha_{2}^{2}),
α1zn1+1e1α2zn2+1e2,α1zn1+m1e1+α2zn2+m2e2=z2n1+1+m1(α12α22),\bigl{\langle}\alpha_{1}z^{-n_{1}+1}e_{1}-\alpha_{2}z^{-n_{2}+1}e_{2},\,\,\alpha_{1}z^{-n_{1}+m_{1}}e_{1}^{\lor}+\alpha_{2}z^{-n_{2}+m_{2}}e_{2}^{\lor}\bigr{\rangle}=z^{-2n_{1}+1+m_{1}}(\alpha_{1}^{2}-\alpha_{2}^{2}),
α1zn1+1e1α2zn2+1e2,α1zn1+1+m1e1α2zn2+1+m2e2=0.\bigl{\langle}\alpha_{1}z^{-n_{1}+1}e_{1}-\alpha_{2}z^{-n_{2}+1}e_{2},\,\,\alpha_{1}z^{-n_{1}+1+m_{1}}e_{1}^{\lor}-\alpha_{2}z^{-n_{2}+1+m_{2}}e_{2}^{\lor}\bigr{\rangle}=0.

Hence, under the condition n112m1=n212m2n_{1}-\frac{1}{2}m_{1}=n_{2}-\frac{1}{2}m_{2} and α12+α22=0\alpha_{1}^{2}+\alpha_{2}^{2}=0, 𝒫(𝒏,α;b)(𝒱)\mathcal{P}^{({\boldsymbol{n}},\alpha;b)}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if 2b=2n11m12b=2n_{1}-1-m_{1}. Similarly, if 𝜶()2{\boldsymbol{\alpha}}\in({\mathbb{C}}^{\ast})^{2}, and if 𝒫(𝒏,𝜶;𝒃)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}) is compatible with C𝒎C_{{\boldsymbol{m}}}, we obtain n112m1=n212m2n_{1}-\frac{1}{2}m_{1}=n_{2}-\frac{1}{2}m_{2} and α12+α22=0\alpha_{1}^{2}+\alpha_{2}^{2}=0, and under these conditions, the compatibility condition is (b11)+b2=2n11m1(b_{1}-1)+b_{2}=2n_{1}-1-m_{1}, i.e., n112m1=12(b1+b2)n_{1}-\frac{1}{2}m_{1}=\frac{1}{2}(b_{1}+b_{2}). If α1=0\alpha_{1}=0, then the filtration 𝒫(𝒏,𝜶;𝒃)(𝒱)\mathcal{P}^{({\boldsymbol{n}},{\boldsymbol{\alpha}};{\boldsymbol{b}})}_{\ast}(\mathcal{V}) is compatible with the decomposition 𝒱=𝒪U(0)e1𝒪U(0)e2\mathcal{V}=\mathcal{O}_{U}(\ast 0)e_{1}\oplus\mathcal{O}_{U}(\ast 0)e_{2}. Hence, it is compatible with C𝒎C_{{\boldsymbol{m}}} if and only if it equals to the canonical filtered extension in §4.1.3 by Lemma 4.4. It is equivalent to ni12mi=bin_{i}-\frac{1}{2}m_{i}=b_{i}.  

Suppose m1=m2=:mm_{1}=m_{2}=:m. Then, the filtered bundle 𝒫((0,0);m/2)(𝒱)\mathcal{P}^{((0,0);-m/2)}_{\ast}(\mathcal{V}) of type I is unique by Lemma 4.24, which equals the canonical filtered extension. There are two filtered bundles 𝒫((0,0),𝜶;(1+m)/2)(𝒱)\mathcal{P}_{\ast}^{((0,0),{\boldsymbol{\alpha}};-(1+m)/2)}(\mathcal{V}) of type II corresponding to 𝜶=(1,±1){\boldsymbol{\alpha}}=(1,\pm\sqrt{-1}). The filtered bundles 𝒫((0,0),𝜶;(b1,b1m))(𝒱)\mathcal{P}^{((0,0),{\boldsymbol{\alpha}};(b_{1},-b_{1}-m))}_{\ast}(\mathcal{V}) of type (III-1) are parameterized by 𝜶=(1,±1){\boldsymbol{\alpha}}=(1,\pm\sqrt{-1}) and m+12<b1<m2-\frac{m+1}{2}<b_{1}<-\frac{m}{2}, where we set 𝒏=(0,0){\boldsymbol{n}}=(0,0). The filtered bundle of type (III-2) does not exist.

Suppose m1m2m_{1}\neq m_{2}. Then, there does not exist the filtered bundle of type I, II nor (III-1). There is a unique filtered bundle of type (III-2), which is equal to the canonical filtered extension.

Acknowledgements

The authors are grateful to the reviewer for his/her careful reading and valuable comments.

T.M. is grateful to Michael McBreen, Natsuo Miyatake, Franz Pedit Martin Traizet and Hitoshi Fujioka for interesting discussions. T.M. is partially supported by 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), and the Grant-in-Aid for Scientific Research (C) (No. 20K03609), Japan Society for the Promotion of Science. T.M. is also partially supported by the Research Institute for Mathematical Sciences, an International Joint Usage/Research Center located in Kyoto University.

Q.L. 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.

References

  • [1] A. Beauville, M. S. Narasimhan, S. Ramanan, Spectral curves and the generalised theta divisor, J. Reine Angew. Math. 398 (1989), 169–179.
  • [2] O. Biquard, P. Boalch Wild non-abelian Hodge theory on curves, Compos. Math. 140 (2004), 179–204.
  • [3] S. K. Donaldson, Boundary value problems for Yang-Mills fields, J. Geom. Phys. 8 (1992), 89–122.
  • [4] O. Forster, Lectures on Riemann surfaces, Graduate Texts in Mathematics, 81. Springer-Verlag, New York-Berlin, 1981.
  • [5] N. J. Hitchin, The self-duality equations on a Riemann surface, Proc. London Math. Soc. (3) 55 (1987), 59–126.
  • [6] N. J. Hitchin, Stable bundles and integrable systems, Duke Math. J. 54 (1987), 91–114.
  • [7] N. J. Hitchin, Lie groups and Teichmüller space, Topology 31 (1992), 449–473.
  • [8] Q. Li, An introduction to Higgs bundles via harmonic maps, SIGMA Symmetry, Integrability Geom. Methods and Appl. 15 (2019), Paper No. 035, 30 pp
  • [9] Q. Li, T. Mochizuki, Complete solutions of Toda equations and cyclic Higgs bundles over non-compact surfaces, arXiv:2010.05401
  • [10] Q. Li, T. Mochizuki, Isolated singularities of Toda equations and cyclic Higgs bundles, arXiv:2010.06129,
  • [11] Q. Li, T. Mochizuki, Higgs bundles in the Hitchin section over non-compact hyperbolic surfaces, arXiv:2307.03365
  • [12] T. Mochizuki, Wild harmonic bundles and wild pure twistor DD-modules, Astérisque 340, (2011)
  • [13] T. Mochizuki, Asymptotic behaviour of certain families of harmonic bundles on Riemann surfaces, J. Topol. 9 (2016), 1021–1073.
  • [14] T. Mochizuki, Good wild harmonic bundles and good filtered Higgs bundles, SIGMA Symmetry Integrability Geom. Methods Appl. 17 (2021), Paper No. 068, 66 pp
  • [15] T. Mochizuki, Asymptotic behaviour of the Hitchin metric on the moduli space of Higgs bundles, arXiv:2305.17638
  • [16] T. Mochizuki, Sz. Szabó, Asymptotic behaviour of large-scale solutions of Hitchin’s equations in higher rank, arXiv:2303.04913
  • [17] C. Simpson, Constructing variations of Hodge structure using Yang-Mills theory and application to uniformization, J. Amer. Math. Soc. 1 (1988), 867–918.
  • [18] C. Simpson, Harmonic bundles on non-compact curves, J. Amer. Math. Soc. 3 (1990), 713–770.
  • [19] C. Simpson, Higgs bundles and local systems, Publ. I.H.E.S., 75 (1992), 5–95.
  • [20] A. Tamburelli, M. Wolf, Planar minimal surfaces with polynomial growth in the 𝒮p(4,){\mathcal{S}p}(4,{\mathbb{R}})-symmetric spaces, arXiv:2002.07295