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

Counting Non-abelian Coverings of Algebraic Curve

Peisheng Yu
Abstract

In this article, we study the etale coverings of an algebraic curve CC with Galois group a semi-direct product /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}. Especially, for a given etale cyclic nn-covering DCD\to C, we determine how many curves EE are there, satisfying EDE\to D is an etale cyclic mm-covering and ECE\to C is Galois with non-abelian Galois group, under the assumption gcd(m,n)=1gcd(m,n)=1.

1 Introduction

Cyclic covering between algebraic varieties is frequently encountered in algebraic geometry. In the case of algebraic curve, the unramified cyclic coverings are encoded by the torsion points of Jacobian. The goal of this short article is to have a glance at the more general coverings between algebraic curves: we study the unramified Galois covering with Galois group equals a semi-direct product /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}. Troughout this article, the curves we consider are smooth and projective, defined over complex number \mathbb{C}; m,n2m,n\geq 2 are positive integers. For genus g=0g=0, 1\mathbb{P}^{1} does not have unramified covering other that 1\mathbb{P}^{1} itself; For genus g=1g=1, the fundamental group is isomorphic to \mathbb{Z}\oplus\mathbb{Z}, which is abelian. Thus, genus one curve does not have non-abelian unramified covering. For these reasons, the curve CC appeared in this article is assumed having genus g2g\geq 2, smooth and projective defined over \mathbb{C}. We describe how to construct such coverings ECE\to C in section 4. Especially, in the case of gcd(m,n)=1gcd(m,n)=1, we give an explicit formula counting the number of such coverings in section 4 and section 5.

2 Background: Unramified Cyclic Covering

Cyclic nn-coverings of an algebraic variety is well-understood. Roughly speaking, they correspond to the nn-th roots of a certain line bundle over that variety. For the sake of completeness, we record the well-known facts ([BHPVdV15]) concerning the unramified case as in the following subsections.

2.1 Construct Etale Coverings

Let XX be a smooth projective variety, and assume that there is a free action of /n\mathbb{Z}/n\mathbb{Z} on XX. Then the quotient Y=X/(/n)Y=X/(\mathbb{Z}/n\mathbb{Z}) is also smooth and projective. Thus, π:XY\pi:X\to Y is an etale cyclic covering of degree nn.

The pushforward π𝒪X\pi_{*}\mathcal{O}_{X} has a structure of 𝒪Y\mathcal{O}_{Y} algebra. Under the action of /n\mathbb{Z}/n\mathbb{Z}, π𝒪X\pi_{*}\mathcal{O}_{X} decomposes into irreducible representations of /n\mathbb{Z}/n\mathbb{Z}. Namely, we have:

π𝒪X𝒪Y1n+1\pi_{*}\mathcal{O}_{X}\cong\mathcal{O}_{Y}\oplus\mathcal{L}^{-1}\oplus\cdots\oplus\mathcal{L}^{-n+1}

where LPic(Y)L\in Pic(Y) is a nn-torsion line bundle, which means n𝒪Y\mathcal{L}^{\otimes n}\cong\mathcal{O}_{Y}.

Conversely, take a nn-torsion line bundle \mathcal{L} over YY, we can construct such an XX explicitly: fixing an isomorphism n𝒪Y\mathcal{L}^{\otimes n}\cong\mathcal{O}_{Y}, we can endow the sheaf i=0n1i\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i} with the structure of a 𝒪Y\mathcal{O}_{Y}-algebra. Let X=Spec(i=0n1i)X=Spec(\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i}), then the 𝒪Y\mathcal{O}_{Y}-structure map 𝒪Yi=0n1i\mathcal{O}_{Y}\to\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i} gives a surjection XYX\to Y, which is an etale cyclic nn-covering. Note that under this construction, in order to guarantee XX to be connected, we require nn to be minimum: i.e. n=ord()n=ord(\mathcal{L}) in the group Pic(Y)Pic(Y). Moreover, if we replace \mathcal{L} by any k\mathcal{L}^{k} with gcd(k,n)=1gcd(k,n)=1, the vairety XX^{\prime} we obtain is isomorphic to XX constructed above, because i=0n1ii=0n1ki\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i}\cong\bigoplus_{i=0}^{n-1}\mathcal{L}^{-ki} as 𝒪Y\mathcal{O}_{Y}-algebra. Therefore, if one wants to count XX up to isomorphism, such equvilance k\mathcal{L}\sim\mathcal{L}^{k} needs to be modulo out.

More geometrically, suppose we have a nn-torsion line bundle \mathcal{L} on YY, and the constant section 1Γ(Y,𝒪Y)1\in\Gamma(Y,\mathcal{O}_{Y}). We denote 𝕃\mathbb{L} the total space of \mathcal{L}, and we let p:𝕃Yp:\mathbb{L}\to Y be the bundle projection. If tΓ(𝕃,p)t\in\Gamma(\mathbb{L},p^{*}\mathcal{L}) is the tautological section, the zero divisor of p1tnp^{*}1-t^{n} defines XX in 𝕃\mathbb{L}. π=p|X\pi=p|_{X} exhibits XX as an etale cyclic nn-covering of YY. For such XX, we can also verify that π𝒪Xi=0n1i\pi_{*}\mathcal{O}_{X}\cong\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i}. Under such description, let K=σ/nK=\left\langle\sigma\right\rangle\cong\mathbb{Z}/n\mathbb{Z}, and let χ\chi denote the generating character of KK, i.e. χ(σ)=ζn\chi(\sigma)=\zeta_{n}, a primitive nn-th root of unity. Then i\mathcal{L}^{-i} is the eigenspace of χi\chi^{i}.

2.2 Case of Curve

The above argument exhibits a bijection between nn-torsion points in Pic(Y)Pic(Y) and etale cyclic nn-covering of YY. In the case of a smooth projective curve CC of genus g2g\geq 2, nn-torsion points in Pic0(C)Pic(C)Pic^{0}(C)\subset Pic(C) is just JC[n]J_{C}[n], where JCJ_{C} is the Jacobian of CC. Over \mathbb{C}, we know that JC[n](/n)2gJ_{C}[n]\cong(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g}. If we denote S(/n)2gS\subset(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g} as the primitive elements in (/n)2g(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g}: i.e. S={v(/n)2g,ord(v)=n}S=\{v\in(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g},ord(v)=n\}, then up to isomorphisms, the connected etale cyclic nn-covering of CC is bijective to S/(/n)×S/(\mathbb{Z}/n\mathbb{Z})^{\times}.

3 Motivation

This study is motivated by a result considering dihedral covering of 1\mathbb{P}^{1}. Roughly speaking, this result shows that all the etale cyclic nn-coverings of a hyperelliptical curve give rise to dihedral coverings of 1\mathbb{P}^{1}, as is indicated in the following theorem:

Theorem 3.1: Suppose HH is a hyperelliptical curve of genus g2g\geq 2, such that the 2-to-1 covering H1H\to\mathbb{P}^{1} ramified at 2g+22g+2 points. There is a hyperelliptical involution τ:HH\tau:H\to H interchanging the two sheets of HH. Consider π:XH\pi:X\to H, which is an etale cyclic nn-covering of HH. Then the involution τ\tau lifts to an involution τ~:XX\tilde{\tau}:X\to X. In other words, X1X\to\mathbb{P}^{1} is Galois, with Galois group D2nD_{2n}.

Proof: Suppose XX corresponds to a nn-torsion line bundle \mathcal{L} over HH, which by section 2.2 also corresponds to a nn-torsion point in JH[n](/n)2gJ_{H}[n]\cong(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g}. We need to determine the induced action of τ\tau on JHJ_{H}. Assume the affine piece of HH over A1A^{1} is given by equation y2=f(x)y^{2}=f(x), where f(x)f(x) is a degree 2g+22g+2 polynomial in xx. Then ωi=xi1dxy,i=1,2,,g\omega_{i}=\frac{x^{i-1}dx}{y},i=1,2,\cdots,g form a basis of H0(H,ωH)H^{0}(H,\omega_{H}). Since the hyperelliptical involution τ\tau sends (x,y)(x,y) to (x,y)(x,-y), it sends ωi\omega_{i} to ωi-\omega_{i}. Since JH=H0(H,ωH)/H1(H,)J_{H}=H^{0}(H,\omega_{H})^{*}/H_{1}(H,\mathbb{Z}), we see that τ\tau induces [1]:JHJH[-1]:J_{H}\to J_{H}. Combining with JH[n](/n)2gJ_{H}[n]\cong(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g}, the induced action of τ\tau sends v(/n)2gv\in(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g} to v-v. Thus, the torsion line bundle \mathcal{L} is sent to 1\mathcal{L}^{-1}. Note that X=Spec(i=0n1i)X=Spec(\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i}), we see τ~\tilde{\tau} sends XX to Spec(i=0n1i)XSpec(\bigoplus_{i=0}^{n-1}\mathcal{L}^{i})\cong X. Therefore, τ~\tilde{\tau} is an isomorphism of XX, satisfying τ~2=idX\tilde{\tau}^{2}=id_{X}. From this, we know X1X\to\mathbb{P}^{1} is Galois.
To determine the Galois group of this etale covering, let σ\sigma denotes the generator of the Galois group of the covering XHX\to H, which is isomorphic to /n\mathbb{Z}/n\mathbb{Z}. Recall that σ\sigma acts on i=0n1i\bigoplus_{i=0}^{n-1}\mathcal{L}^{-i}, thus acts on XX, in the following way: Each k\mathcal{L}^{k} is an eigenspace of σ\sigma. Suppose the eigenvalue of σ\sigma acting on \mathcal{L} is ζ\zeta, for some primitive nn-th root of unity ζ\zeta, then the eigenvalue of σ\sigma acting on k\mathcal{L}^{k} equals to ζk\zeta^{k}. Together with the fact that τ~\tilde{\tau} sends i\mathcal{L}^{i} to i\mathcal{L}^{-i}, one easily checks that τ~στ~=σ1\tilde{\tau}\sigma\tilde{\tau}=\sigma^{-1}. Therefore, the Galois group of X1X\to\mathbb{P}^{1} is σ,τ~|τ~στ~=σ1D2n\left\langle\sigma,\tilde{\tau}|\tilde{\tau}\sigma\tilde{\tau}=\sigma^{-1}\right\rangle\cong D_{2n}.

Non-abelian covering appears in the theorem above: the dihedral covering. Inspiring by that, we want to construct non-abelian covering for more general curves. Instead of H1H\to\mathbb{P}^{1}, our background setting will be DCD\to C, which is an etale cyclic nn-covering between smooth projective curves over \mathbb{C}, and the genus of CC is g2g\geq 2. The goal is to find all curves EE up to isomorphism, such that EDE\to D is a etale cyclic mm-covering, and the composition EDCE\to D\to C is Galois with non-abelian group /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}. We will describe how to construct such non-abelian coverings. In the case of gcd(m,n)=1gcd(m,n)=1, we will also give an explicit formula counting the number of such EE up to isomorphism.

4 The Main Result

Assume the curve CC has genus g2g\geq 2, then by Riemann-Hurwitz formula, the genus of the curve DD, which we denote by hh, is h=n(g1)+1h=n(g-1)+1. Let σ\sigma be a generator of the Galois group of the covering π:DC\pi:D\to C, σ/n\left\langle\sigma\right\rangle\cong\mathbb{Z}/n\mathbb{Z}. Suppose the etale covering DCD\to C corresponds to \mathcal{L}, which is a nn-torsion line bundle over curve CC. In order to search for etale cyclic mm-covering of DD, we should look for the mm-torsion points in the Jacobian of DD: JD[m]J_{D}[m]. Over \mathbb{C}, JD[m](/m)2hJ_{D}[m]\cong(\mathbb{Z}/m\mathbb{Z})^{\oplus 2h}. However, not all of these mm-torsion points in JDJ_{D} would give rise to the covering EDE\to D we are looking for. In fact, in order to understand what the composition EDCE\to D\to C is, we need to understand the induced action of σ\sigma on JD[m]J_{D}[m].

Since σ\sigma is an isomorphism of curve DD, the induced action of σ\sigma on the Jacobian JDJ_{D} is also an isomorphism of JDJ_{D}, which we denoted as σ\sigma as well. In particular, it preserves mm-torsion points in JDJ_{D}, thus σ\sigma is also an isomorphism on JD[m]J_{D}[m]. By definition, JD=H0(D,ωD)/H1(D,)J_{D}=H^{0}(D,\omega_{D})^{*}/H_{1}(D,\mathbb{Z}). We need first understand the action of σ\sigma on the canonical sheaf of DD. Actually, since the covering π:DC\pi:D\to C is etale, we have πωC=ωD\pi^{*}\omega_{C}=\omega_{D}. Also recall that:

π𝒪D𝒪C1n+1\pi_{*}\mathcal{O}_{D}\cong\mathcal{O}_{C}\oplus\mathcal{L}^{-1}\oplus\cdots\oplus\mathcal{L}^{-n+1} (4.1)

Combine this with projection formula, we get:

πωD=ππωCωCπ𝒪Di=0n1ωCi\pi_{*}\omega_{D}=\pi_{*}\pi^{*}\omega_{C}\cong\omega_{C}\otimes\pi_{*}\mathcal{O}_{D}\cong\bigoplus_{i=0}^{n-1}\omega_{C}\otimes\mathcal{L}^{-i} (4.2)

Taking global sections, we obtain:

H0(D,ωD)i=0n1H0(C,ωCi)H^{0}(D,\omega_{D})\cong\bigoplus_{i=0}^{n-1}H^{0}(C,\omega_{C}\otimes\mathcal{L}^{-i}) (4.3)

And the right hand side of (4.3) is exactly the decomposition into σ\sigma-representations. More precisely, each H0(C,ωCi)H^{0}(C,\omega_{C}\otimes\mathcal{L}^{-i}) is an eigenspace of σ\sigma, with eigenvalue ζi\zeta^{i}, where ζ\zeta is a primitive nn-th root of unity. As one can see from (4.3), the trivial subspace of σ\sigma inside H0(D,ωD)H^{0}(D,\omega_{D}) is just H0(C,ωC)H^{0}(C,\omega_{C}), and we single out the non-trivial representation:

R:=i=1n1H0(C,ωCi)R:=\bigoplus_{i=1}^{n-1}H^{0}(C,\omega_{C}\otimes\mathcal{L}^{-i}) (4.4)

By Riemann-Roch, the dimension:

dimH0(C,ωC)=gdim_{\mathbb{C}}H^{0}(C,\omega_{C})=g (4.5)

and for i0i\neq 0 :

dimH0(C,ωCi)=g1dim_{\mathbb{C}}H^{0}(C,\omega_{C}\otimes\mathcal{L}^{-i})=g-1 (4.6)

Thus we get:

dimR=(n1)(g1)dim_{\mathbb{C}}R=(n-1)(g-1) (4.7)

Here, the space RR is nothing but the tangent space of the Prym variety of the covering π:DC\pi:D\to C. By definition, the Prym variety PP of the covering π\pi is the principle connected component of the kernel of the induced norm map Nm(π):JDJCNm(\pi):J_{D}\to J_{C} [Ago20]. Alternatively, P=Im(1σ)=ker(1+σ++σn1)0P=Im(1-\sigma)=ker(1+\sigma+\cdots+\sigma^{n-1})^{0}.

Theorem 4.1: Suppose v(/m)2hJD[m]v\in(\mathbb{Z}/m\mathbb{Z})^{\oplus 2h}\cong J_{D}[m] is primitive, satisfying σv=λv\sigma v=\lambda v for some λ(/m)×\lambda\in(\mathbb{Z}/m\mathbb{Z})^{\times} and λ1\lambda\neq 1. (i.e. vv is an eigenvector of σ:JD[m]JD[m]\sigma:J_{D}[m]\to J_{D}[m], with eigenvalue λ\lambda) Then the etale cyclic mm-covering EDE\to D corresponding to vv will gives a connected non-abelian Galois covering of CC after composing with π:DC\pi:D\to C. Conversely, every connected etale Galois covering of CC with Galois group /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z} takes such form.

Proof: The requirement that vv is primitive is to guarantee the order of vv is exactly mm, which is equivalent to the requirement that the curve EE in covering EDE\to D is connected. In order to check the covering ECE\to C is Galois, we again need to look at the action of σ\sigma on vv. As in our assumption, σv=λv\sigma v=\lambda v for some λ(/m)×\lambda\in(\mathbb{Z}/m\mathbb{Z})^{\times}, and recall that vv and λv\lambda v give rise to the same curve EE, we see σ\sigma actually lifts to an automorphism of curve EE. Therefore, the covering ECE\to C is Galois. To determine the Galois group, let θ\theta be a generator of the covering EDE\to D corresponds to eigenvector vv, θ/m\left\langle\theta\right\rangle\cong\mathbb{Z}/m\mathbb{Z}. Suppose vv is represented by a mm-torsion line bundle η\eta, and

E=Spec(𝒪Dη1ηm+1)E=Spec(\mathcal{O}_{D}\oplus\mathcal{\eta}^{-1}\oplus\cdots\oplus\mathcal{\eta}^{-m+1}) (4.8)

Then the action of σ\sigma sends η\eta to ηλ\eta^{\lambda}. On each ηi\eta^{-i}, the action of θ\theta is multiply by ϵi\epsilon^{i}, where ϵ\epsilon is a primitive mm-th root of unity. From these, we can see that each ηk\eta^{k} is an eigen-subbundle for σθσ1\sigma\theta\sigma^{-1}, whose action is multiplying by ϵλ1k\epsilon^{\lambda^{-1}k}. Therefore, we deduce that σθσ1=θλ1\sigma\theta\sigma^{-1}=\theta^{\lambda^{-1}}. Hence, the Galois group of the covering ECE\to C is:

σ,θ|σθσ1=θλ1,σn=θm=1\left\langle\sigma,\theta|\sigma\theta\sigma^{-1}=\theta^{\lambda^{-1}},\sigma^{n}=\theta^{m}=1\right\rangle (4.9)

Once λ1\lambda\neq 1, the group above is non-abelian, which is nothing but a semi-direct product /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}.

For the converse, given a etale Galois covering ECE\to C with Galois group /m/n\cong\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}, we consider the intermediate curve D:=E/(/m)D:=E/(\mathbb{Z}/m\mathbb{Z}). Since /m\mathbb{Z}/m\mathbb{Z} is a normal subgroup of its Galois group, the covering DCD\to C is Galois and etale, with Galois group σ/n\left\langle\sigma\right\rangle\cong\mathbb{Z}/n\mathbb{Z}. Therefore, DCD\to C is determined by a nn-torsion point in JCJ_{C}, and EDE\to D is determined by a mm-torsion point in vJDv\in J_{D}. Since EE maps to itself under the action of σ\sigma, we deduce that vv must map to λv\lambda v under the action of σ\sigma, for some λ(/m)×\lambda\in(\mathbb{Z}/m\mathbb{Z})^{\times}. In order to make the covering non-abelian, it requires λ1\lambda\neq 1.

Moreover, we wish to count the number of such curve EE, given DCD\to C. Using theorem 4.1, this is equivalent to counting the number of eigen-directions (eigenvectors up to a scalar) of σ\sigma acting on (/m)2h(\mathbb{Z}/m\mathbb{Z})^{\oplus 2h}, with eigenvalue 1\neq 1. This turns out to be a little tricky. To ease the computation, we only consider the case gcd(m,n)=1gcd(m,n)=1. We will give two methods of counting such EE, in the section 4.1 and 4.2.

4.1 First Method of Counting

The idea of the first method is using similarity between matrices. To begin with, recall that the Jacobian JD=H0(D,ωD)/H1(D,)J_{D}=H^{0}(D,\omega_{D})^{*}/H_{1}(D,\mathbb{Z}). Since σ\sigma is an automorphism of JDJ_{D}, it is also an isomorphism of the lattice H1(D,)2hH_{1}(D,\mathbb{Z})\cong\mathbb{Z}^{\oplus 2h}. In this way, σ\sigma becomes an element in GL2h()GL_{2h}(\mathbb{Z}), satisfying σn=id\sigma^{n}=id. Since we know the eigenspace decomposition of H0(D,ωD)H^{0}(D,\omega_{D}) under the action of σ\sigma, we can easily deduce what matrix is σ\sigma similar to over \mathbb{C}. To see this, the complexification of σ\sigma, namely σ\sigma\otimes\mathbb{C}, acts on H1(D,)H_{1}(D,\mathbb{Z})\otimes\mathbb{C}, which is H1(D,)H_{1}(D,\mathbb{C}). Since H1(D,)=H1,0(D)H0,1(D)H^{1}(D,\mathbb{C})=H^{1,0}(D)\oplus H^{0,1}(D), taking dual, we obtain H1(D,)H0(D,ωD)H0(D,ωD)H_{1}(D,\mathbb{C})\cong H^{0}(D,\omega_{D})\oplus H^{0}(D,\omega_{D})^{*}. Further recall that: H0(D,ωD)H0(C,ωC)i=1n1H0(C,ωCi)H^{0}(D,\omega_{D})\cong H^{0}(C,\omega_{C})\oplus\bigoplus_{i=1}^{n-1}H^{0}(C,\omega_{C}\otimes\mathcal{L}^{-i}) gives the eigenspace decomposition of the action of σ\sigma, we obtain:

σ=I2gζI2g2ζn1I2g2\sigma\otimes\mathbb{C}=I_{2g}\oplus\zeta I_{2g-2}\oplus\cdots\oplus\zeta^{n-1}I_{2g-2} (4.10)

If we denote J=diag(ζ,ζ2,,ζn1)J=diag(\zeta,\zeta^{2},\cdots,\zeta^{n-1}), then σ\sigma is similar to the matrix I2g(JI2g2)I_{2g}\oplus(J\otimes I_{2g-2}) over \mathbb{C}.

Define the (n1)×(n1)(n-1)\times(n-1) matrix G=G=: (001101011)\begin{pmatrix}0&\cdots&0&-1\\ 1&\cdots&0&-1\\ \vdots&\ddots&\vdots&\vdots\\ 0&\cdots&1&-1\end{pmatrix} It is easy to see GG is similar to JJ over \mathbb{C}. Thus, σ\sigma is similar to I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) over \mathbb{C}. Note that both σ\sigma and I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) over \mathbb{C} is defined over \mathbb{Z}, thus over \mathbb{Q}, we deduce that σ\sigma and I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) is similar over \mathbb{Q}. The difficulty here we cannot obtain similarity over \mathbb{Z}. To make the argument work, we must require gcd(m,n)=1gcd(m,n)=1 in the following.

The goal is to obtain the similarity type of σ\sigma after reduction modulo mm. Once we get that information, we will know how σ\sigma acts on 1mH1(D,)/H1(D,)\frac{1}{m}H_{1}(D,\mathbb{Z})/H_{1}(D,\mathbb{Z}), which is just the mm-torsion points in JDJ_{D}. We achieve our goal in several steps:

Step 1:
Take any prime p|mp|m, consider the similarity type of σ(p):=σ\sigma(p):=\sigma mod pp. Here we view σ(p)\sigma(p) as an element in GL2h(𝔽p)GL_{2h}(\mathbb{F}_{p}). Because σ\sigma is similar to I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) over \mathbb{Q}, they have the same characteristic polynomial and minimal polynomial over \mathbb{Q}. Therefore, the characteristic polynomial of σ\sigma equals to (x1)2g(1+x++xn1)2g2(x-1)^{2g}(1+x+\cdots+x^{n-1})^{2g-2}, and the minimal polynomial of σ\sigma equals to xn1x^{n}-1. The characteristic polynomial of σ(p)\sigma(p) is just (x1)2g(1+x++xn1)2g2(x-1)^{2g}(1+x+\cdots+x^{n-1})^{2g-2} mod pp, and the minimal polynomial of σ(p)\sigma(p), say q(x)q(x), must dividing xn1x^{n}-1 mod pp. Since gcd(p,n)=1gcd(p,n)=1, xn1x^{n}-1 has distinct roots over 𝔽¯p\overline{\mathbb{F}}_{p}. Assume q(x)xn1q(x)\neq x^{n}-1 mod pp, then some factors of xn1x^{n}-1 must be left out by q(x)q(x). As a result, such factors cannot appear in the characteristic polynomial (x1)2g(1+x++xn1)2g2(x-1)^{2g}(1+x+\cdots+x^{n-1})^{2g-2} mod pp. However, (x1)2g(1+x++xn1)2g2(x-1)^{2g}(1+x+\cdots+x^{n-1})^{2g-2} mod pp has the same factor with xn1x^{n}-1 mod pp, leading to a contradiction. Thus, q(x)=xn1q(x)=x^{n}-1 mod pp.

Therefore, σ(p)\sigma(p) and I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) mod pp have the same characteristic and minimal polynomial. Note that the minimal polynomial xn1x^{n}-1 does not have multiple root over 𝔽¯p\overline{\mathbb{F}}_{p}, they are both diagonalizable over 𝔽¯p\overline{\mathbb{F}}_{p}. Same characteristic polynomial gives the same multiplicity of eigenvalues. Thus, over 𝔽¯p\overline{\mathbb{F}}_{p}, σ(p)\sigma(p) is similar to I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}). Notice both matrices are defined over 𝔽p\mathbb{F}_{p}, they are actually similar over 𝔽p\mathbb{F}_{p}.

Step 2:
In order to lift the similarity relation from over 𝔽p\mathbb{F}_{p} to over /m\mathbb{Z}/m\mathbb{Z}, we need some results of J.Pomfret [Pom73]:

Lemma 4.2: Let RR be a finite local ring with maximal ideal MM, and R/M=𝔽pf=kR/M=\mathbb{F}_{p^{f}}=k. Let α\alpha, β\beta be elements of GLn(R)GL_{n}(R) with gcd(|α|,p)=gcd(|β|,p)=1gcd(|\left\langle\alpha\right\rangle|,p)=gcd(|\left\langle\beta\right\rangle|,p)=1 (where |x||\left\langle x\right\rangle| means the order of xx). Then α\alpha is similar to β\beta if and only if α\alpha is similar to β\beta modulo MM.

If RR is a finite commutative ring with identity, then RR is uniquely isomorphic to a product of finite local rings: Ri=1tRiR\cong\prod_{i=1}^{t}R_{i}. Suppose RiR_{i} is a finite local ring with maximal ideal MiM_{i}, and Ri/Mi=kiR_{i}/M_{i}=k_{i}, we have the epimorphisms:

GLn(R)i=1tGLn(Ri)πiGLn(Ri)GLn(ki)GL_{n}(R)\cong\prod_{i=1}^{t}GL_{n}(R_{i})\stackrel{{\scriptstyle\pi_{i}}}{{\longrightarrow}}GL_{n}(R_{i})\to GL_{n}(k_{i}) (4.11)

Combining lemma 4.2 and (4.11), we get:

Lemma 4.3: Let RR be a finite commutative ring with identity and let the cardinality of RR be mm. Two elements α\alpha and β\beta of GLn(R)GL_{n}(R) satisfying gcd(|α|,m)=gcd(|β|,m)=1gcd(|\left\langle\alpha\right\rangle|,m)=gcd(|\left\langle\beta\right\rangle|,m)=1 are similar if and only if they are similar over each residue field kik_{i}.

In step 1, we see σ(p)\sigma(p) and I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) are similar over 𝔽p\mathbb{F}_{p} for any p|mp|m. Also note that 𝔽p\mathbb{F}_{p}’s are all the residue fields of /m\mathbb{Z}/m\mathbb{Z}. Applying lemma 4.3 to our case: R=/mR=\mathbb{Z}/m\mathbb{Z}, α=σ(m)\alpha=\sigma(m) (=σ=\sigma mod mm), and β=I2g(GI2g2)\beta=I_{2g}\oplus(G\otimes I_{2g-2}), we see that σ(m)\sigma(m) is similar to I2g(GI2g2)I_{2g}\oplus(G\otimes I_{2g-2}) over /m\mathbb{Z}/m\mathbb{Z}.

Step 3:
Denote A=GI2g2A=G\otimes I_{2g-2}, viewed as an element in GL(2g2)(n1)(/m)GL_{(2g-2)(n-1)}(\mathbb{Z}/m\mathbb{Z}). The part I2gI_{2g} corresponds to the eigenspace with eigenvalue 11. Since we only want the eigenspace with eigenvalue λ1\lambda\neq 1, it is equivalent to count the eigen-directions that AA have acting on (/m)(2g2)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(2g-2)(n-1)}.

We need to understand (/m)(2g2)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(2g-2)(n-1)} as a (/m)[A](\mathbb{Z}/m\mathbb{Z})[A] module. Since A=GI2g2A=G\otimes I_{2g-2}, we only need to study the (/m)[G](\mathbb{Z}/m\mathbb{Z})[G]-mod structure on (/m)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(n-1)}, because as a (/m)[A](\mathbb{Z}/m\mathbb{Z})[A] module, (/m)(2g2)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(2g-2)(n-1)} is just isomorphic to the direct sum of (2g2)(2g-2)-many copies of (/m)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(n-1)}, with each summand carrying the same (/m)[G](\mathbb{Z}/m\mathbb{Z})[G]-mod structure. The minimal polynomial and the characteristic polynomial of GG both equals to f(x)=1+x+x2++xn1f(x)=1+x+x^{2}+\cdots+x^{n-1}. Therefore, as a (/m)[G](\mathbb{Z}/m\mathbb{Z})[G]-mod, (/m)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(n-1)} is isomorphic to (/m)[G](\mathbb{Z}/m\mathbb{Z})[G].

Now we compute (/m)[G](\mathbb{Z}/m\mathbb{Z})[G]. We have the decomposition:

(/m)[G]p|m(/pe)[G](\mathbb{Z}/m\mathbb{Z})[G]\cong\prod_{p|m}(\mathbb{Z}/p^{e}\mathbb{Z})[G] (4.12)

For a set of eigenvectors: {vp(/pe)(n1)}p|m\{v_{p}\in(\mathbb{Z}/p^{e}\mathbb{Z})^{\oplus(n-1)}\}_{p|m}, satisfying Gvp=λpvpGv_{p}=\lambda_{p}v_{p} mod pep^{e}, we can find a unique vector v(/m)(n1)v\in(\mathbb{Z}/m\mathbb{Z})^{\oplus(n-1)}, satifying Gv=λvGv=\lambda v, and vvpv\equiv v_{p} mod pep^{e}, λλp\lambda\equiv\lambda_{p} mod pep^{e}. In this way, we reduce the computation to (/pe)[G](\mathbb{Z}/p^{e}\mathbb{Z})[G].

Note that (/pe)[G](/pe)[x]/f(x)(\mathbb{Z}/p^{e}\mathbb{Z})[G]\cong(\mathbb{Z}/p^{e}\mathbb{Z})[x]/f(x), and f(x)=d|n,d1Φd(x)f(x)=\prod_{d|n,d\neq 1}\Phi_{d}(x). It suffices for us to consider the behavior of decomposition of each Φd(x)\Phi_{d}(x) mod pep^{e}. We have the following lemma:

Lemma 4.4: For dp1d\nmid p-1, Φd(x)\Phi_{d}(x) has no degree one factor modulo pep^{e}. For dp1d\mid p-1, Φd(x)\Phi_{d}(x) decompose completely into ϕ(d)\phi(d)-many degree one factors modulo pep^{e}.

Proof: Passing to the residue field of /pe\mathbb{Z}/p^{e}\mathbb{Z}, which is just 𝔽p\mathbb{F}_{p}, we first the decomposition of Φd(x)\Phi_{d}(x) in 𝔽p[x]\mathbb{F}_{p}[x]. If Φd(x)\Phi_{d}(x) mod pep^{e} has any linear term (xλ)(x-\lambda), then so does Φd(x)\Phi_{d}(x) mod pp, because λ(/pe)×\lambda\in(\mathbb{Z}/p^{e}\mathbb{Z})^{\times} implies λ0\lambda\neq 0 mod pp. This shows that if Φd(x)\Phi_{d}(x) mod pp does not have any linear factor, then so does Φd(x)\Phi_{d}(x) mod pep^{e}. The only case that Φd(x)\Phi_{d}(x) has root over 𝔽p\mathbb{F}_{p} is that: the order of pp in /d\mathbb{Z}/d\mathbb{Z} is exactly one, i.e. dp1d\mid p-1. (In this case, primitive dd-th roots of unity will be fixed under Frobenius, thus lie in 𝔽p\mathbb{F}_{p}).
Assuming dp1d\mid p-1, then Φd(x)\Phi_{d}(x) splits completely in 𝔽p[x]\mathbb{F}_{p}[x]:

Φd(x)(xa1)(xa2)(xaϕ(d))\Phi_{d}(x)\equiv(x-a_{1})(x-a_{2})\cdots(x-a_{\phi(d)}) (4.13)

Since gcd(d,p)=1gcd(d,p)=1, roots aia_{i} and aja_{j} are distinct. By Hensel’s lemma, we can lift the factorization into /pe\mathbb{Z}/p^{e}\mathbb{Z}, namely:

Φd(x)(xa~1)(xa~2)(xa~ϕ(d))\Phi_{d}(x)\equiv(x-\tilde{a}_{1})(x-\tilde{a}_{2})\cdots(x-\tilde{a}_{\phi(d)}) (4.14)

with a~i(/pe)×\tilde{a}_{i}\in(\mathbb{Z}/p^{e}\mathbb{Z})^{\times} distinct. The lemma is proven.

With the aid of lemma 4.4, let us count how many linear factors can f(x)f(x) factor out in (/pe)[x](\mathbb{Z}/p^{e}\mathbb{Z})[x]. Only for those dd, satisfying: dnd\mid n, d1d\neq 1, and dp1d\mid p-1, Φd(x)\Phi_{d}(x) will contribute ϕ(d)\phi(d)-many linear factors. Therefore, the total number of linear factors in (/pe)[x]/f(x)(\mathbb{Z}/p^{e}\mathbb{Z})[x]/f(x) equals to:

dgcd(n,p1),d1ϕ(d)=gcd(n,p1)1\sum_{d\mid gcd(n,p-1),d\neq 1}\phi(d)=gcd(n,p-1)-1 (4.15)

Tensoring by I2g2I_{2g-2}, we pass from (/pe)[G](\mathbb{Z}/p^{e}\mathbb{Z})[G] to (/pe)[A](\mathbb{Z}/p^{e}\mathbb{Z})[A]. By doing so, the one dimensional eigenspace with eigenvalue a~i\tilde{a}_{i} (corresponding to the linear factor (xa~i)(x-\tilde{a}_{i})) is enlarged into a (2g2)(2g-2)-dimensional eigenspace still with eigenvalue a~i\tilde{a}_{i}. Since the roots of f(x)f(x) are distinct, all the eigenvectors come from the union of such (2g2)(2g-2)-dimensional eigenspaces. We first count the number of primitive points in those eigenspaces:

Take a NN dimensional space, V=(/pe)NV=(\mathbb{Z}/p^{e}\mathbb{Z})^{\oplus N}, the primitive vectors (those elements of order pep^{e}) in VV can be viewed as (/pe)N(p/pe)N(\mathbb{Z}/p^{e}\mathbb{Z})^{\oplus N}-(p\mathbb{Z}/p^{e}\mathbb{Z})^{\oplus N}. Therefore, the number of primitive vectors in VV equals to peN(1pN)p^{eN}(1-p^{-N}). Furthermore, suppose /mp|m/pe\mathbb{Z}/m\mathbb{Z}\cong\prod_{p|m}\mathbb{Z}/p^{e}\mathbb{Z}, and suppose XX is a NN dimensional space over /m\mathbb{Z}/m\mathbb{Z}, say X=(/m)NX=(\mathbb{Z}/m\mathbb{Z})^{\oplus N}, then a vector vXv\in X is primitive if and only if vp=vv_{p}=v mod pep^{e} is primitive in Xp=(/pe)NX_{p}=(\mathbb{Z}/p^{e}\mathbb{Z})^{\oplus N} for every pmp\mid m.

Combining all the arguments above together, we obtain the number of primitive eigenvector for AA in (/m)(2g2)(n1)(\mathbb{Z}/m\mathbb{Z})^{\oplus(2g-2)(n-1)} equals to:

pm(gcd(n,p1)1)pe(2g2)(1p2g+2)\prod_{p\mid m}(gcd(n,p-1)-1)\cdot p^{e(2g-2)}(1-p^{-2g+2}) (4.16)

Since what we need to count is the number of eigen-directions, we need to modulo thus primitive eigenvectors by scalars in (/m)×(\mathbb{Z}/m\mathbb{Z})^{\times}. Thus, we need to divide (4.16) by ϕ(m)\phi(m):

Tg(m,n)=pm(gcd(n,p1)1)pe(2g2)(1p2g+2)/pe(1p1)T_{g}(m,n)=\prod_{p\mid m}(gcd(n,p-1)-1)\cdot p^{e(2g-2)}(1-p^{-2g+2})/p^{e}(1-p^{-1}) (4.17)

Where Tg(m,n)T_{g}(m,n) is just the number we are looking for. Reduce (4.17), we obtain:

Tg(m,n)=m2g3pm(gcd(n,p1)1)(1+p1++p2g+3)T_{g}(m,n)=m^{2g-3}\prod_{p\mid m}(gcd(n,p-1)-1)\cdot(1+p^{-1}+\cdots+p^{-2g+3}) (4.18)

Summing up, we get the following counting result:

Theorem 4.5: Let CC be a smooth projective curve of genus g2g\geq 2, and suppose mm and nn are coprime integers. For a given etale cyclic nn-covering π:DC\pi:D\to C, up to isomorphism, there are Tg(m,n)T_{g}(m,n)-many curves EE, such that EDE\to D is an etale cyclic mm-covering, and the composition ECE\to C is Galois with non-abelian Galois group /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}.

4.2 Second Method of Counting

Here we will give an alternative method to count the number of such curves EE. Recall that the mm-torsion points in JD[m]J_{D}[m] are nothing but 1mH1(D,)/H1(D,)\frac{1}{m}H_{1}(D,\mathbb{Z})/H_{1}(D,\mathbb{Z}). If we can spell out the action of σ\sigma on H1(D,)H_{1}(D,\mathbb{Z}), we will get an explicit matrix representation of σGL2h()\sigma\in GL_{2h}(\mathbb{Z}), up to conjugation over \mathbb{Z}. That will be enough for us to compute the eigenvectors of σ(m)=σ\sigma(m)=\sigma mod mm. The key point is a topological lemma below:

Lemma 4.6: Suppose Σg\Sigma_{g} is a closed topological surface of genus g2g\geq 2, and Σh\Sigma_{h}, where h=n(g1)+1h=n(g-1)+1, is a normal covering surface of Σg\Sigma_{g}: ΣhΣg\Sigma_{h}\to\Sigma_{g}, with group of deck transformation /n\mathbb{Z}/n\mathbb{Z}. Then such covering is equivalent to the ’standard’ cyclic covering between surfaces, as illustrated by Figure 1 [MER03] and Figure 2 [LM18].

Refer to caption
Figure 1: Standard cyclic covering for /5\mathbb{Z}/5\mathbb{Z}
Refer to caption
Figure 2: Standard cyclic covering for /3\mathbb{Z}/3\mathbb{Z}

We first need to translate the word standard into more concrete definition. Let π=π1(Σg)\pi=\pi_{1}(\Sigma_{g}) be the fundamental group of genus gg closed surface, and let a1,b1,,ag,bga_{1},b_{1},\cdots,a_{g},b_{g} be a basis of π\pi satisfying i=1g[ai,bi]=e\prod_{i=1}^{g}[a_{i},b_{i}]=e. Then the standard /n\mathbb{Z}/n\mathbb{Z}-covering corresponds to the epimorphism π/n1\pi\to\mathbb{Z}/n\mathbb{Z}\to 1, sending a1a_{1} to a generator of /n\mathbb{Z}/n\mathbb{Z} and b1,a2,b2,ag,bgb_{1},a_{2},b_{2},\cdots a_{g},b_{g} to 0. In this case, we have a short exact sequence:

1Kπ/n11\to K\to\pi\to\mathbb{Z}/n\mathbb{Z}\to 1 (4.19)

and KK is nothing but the fundamental group of Σh\Sigma_{h}, with h=n(g1)+1h=n(g-1)+1. It is easy to see that such covering ΣhΣg\Sigma_{h}\to\Sigma_{g} has the geometric figure that looks like rotating by /n\mathbb{Z}/n\mathbb{Z} along a hole, which we refer to as standard cyclic covering. Therefore, to prove Lemma 4.6, it suffices to prove the following result:

Lemma 4.7: Let π\pi be the fundamental group of Σg\Sigma_{g}. For every epimorphism p:π/np:\pi\to\mathbb{Z}/n\mathbb{Z}, there is a basis of π\pi, namely x~1,,x~g,y~1,,y~g\tilde{x}_{1},\cdots,\tilde{x}_{g},\tilde{y}_{1},\cdots,\tilde{y}_{g}, such that under pp, x~1\tilde{x}_{1} maps to generator of /n\mathbb{Z}/n\mathbb{Z}, and x~2,,x~g,y~1,,y~g\tilde{x}_{2},\cdots,\tilde{x}_{g},\tilde{y}_{1},\cdots,\tilde{y}_{g} map to 0. Moreover, i=1g[x~i,y~i]=e\prod_{i=1}^{g}[\tilde{x}_{i},\tilde{y}_{i}]=e.

Proof: Consider the following diagram:

   \textstyle{\mathbb{Z}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}modn\scriptstyle{mod\quad n}   π/[π,π]2g\textstyle{\pi/[\pi,\pi]\cong\mathbb{Z}^{2g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ\scriptstyle{\quad\delta}δ~\scriptstyle{\tilde{\delta}}/n\textstyle{\mathbb{Z}/n\mathbb{Z}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}   0\textstyle{0}   

Since /n\mathbb{Z}/n\mathbb{Z} is abelian, p:π/np:\pi\to\mathbb{Z}/n\mathbb{Z} factors through π/[π,π]=H1(Σg,)\pi/[\pi,\pi]=H_{1}(\Sigma_{g},\mathbb{Z}), which is isomorphic to 2g\mathbb{Z}^{2g}. Since 2g\mathbb{Z}^{2g} is a free \mathbb{Z}-mod, thus projective, δ\delta lifts to δ~:2g\tilde{\delta}:\mathbb{Z}^{2g}\to\mathbb{Z}. There exists v2gv\in\mathbb{Z}^{2g}, such that δ~(v)\tilde{\delta}(v) mod nn = δ(v)\delta(v) generates /n\mathbb{Z}/n\mathbb{Z}. If vv is not primitive, we can divide vv by kk\in\mathbb{Z}, such that v/kv/k is primitive and δ~(v)\tilde{\delta}(v)\in\mathbb{Z}. After modulo nn, δ(v/k)\delta(v/k) still generates /n\mathbb{Z}/n\mathbb{Z}. Therefore, we can find a primitive vector v2gv\in\mathbb{Z}^{2g}, satisfying δ(v)(/n)×\delta(v)\in(\mathbb{Z}/n\mathbb{Z})^{\times}. Now since vv is primitive, 2g={v}ker(δ~)\mathbb{Z}^{2g}=\mathbb{Z}\{v\}\oplus ker(\tilde{\delta}), and ker(δ~)2g1ker(\tilde{\delta})\cong\mathbb{Z}^{2g-1}.

Consider the intersection form on H1(Σg,)H_{1}(\Sigma_{g},\mathbb{Z}), i.e. BB, a skew-symmetric, bilinear, unimodular form on 2g\mathbb{Z}^{2g}. Note that B(v,v)=0B(v,v)=0. Consider the linear function B(v,):ker(δ~)B(v,\cdot):ker(\tilde{\delta})\to\mathbb{Z}. It must be surjective, otherwise det(B)±1det(B)\neq\pm 1. Therefore, exists y1ker(δ~)y_{1}\in ker(\tilde{\delta}), primitive, such that {v,y1}\{v,y_{1}\} form a hyperbolic basis. ker(δ)~2g1={y1}Uker(\tilde{\delta)}\cong\mathbb{Z}^{2g-1}=\mathbb{Z}\{y_{1}\}\oplus U, with 2g2U=ker(δ~)\mathbb{Z}^{2g-2}\cong U=ker(\tilde{\delta}).

Further consider the linear function B(y1,):UB(y_{1},\cdot):U\to\mathbb{Z}. If ker(B(y1,))ker(B(y_{1},\cdot)) is the whole UU, then BB takes diagonal form under {v,y1}U\mathbb{Z}\{v,y_{1}\}\oplus U, and the restriction of BB to UU is also unimodular and skew-symmetric, and we can apply induction to get a standard basis. If B(y1,)B(y_{1},\cdot) is non-trivial, let WUW\subset U to be the kernel of B(y1,)B(y_{1},\cdot). Note that W2g3W\cong\mathbb{Z}^{2g-3}, and BB restricts to WW is integral and skew-symmetric. Thus, there is a basis {w1,w2,,w2g3}\{w_{1},w_{2},\cdots,w_{2g-3}\} for WW. We can assume that BB takes the form diag(S1,S2,,Sg2,0)diag(S_{1},S_{2},\cdots,S_{g-2},0) on WW, where SiS_{i}’s are integral skew-symmetric 2×22\times 2 matrices, say Si=(0didi0)S_{i}=\begin{pmatrix}0&d_{i}\\ -d_{i}&0\end{pmatrix} Assume U={t}WU=\mathbb{Z}\{t\}\oplus W, then under the basis {v,y1,w1,,w2g3,t}\{v,y_{1},w_{1},\cdots,w_{2g-3},t\}, BB takes the form:

(01000010000p2g200S100p2g3000S20p2g40000Sg2p200000p10p2g2p2g3p2g4p10)\begin{pmatrix}0&1&0&0&\cdots&0&0\\ -1&0&0&0&\cdots&0&p_{2g-2}\\ 0&0&S_{1}&0&\cdots&0&p_{2g-3}\\ 0&0&0&S_{2}&\cdots&0&p_{2g-4}\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&0&0&\cdots&S_{g-2}&p_{2}\\ 0&0&0&0&\cdots&0&p_{1}\\ 0&-p_{2g-2}&-p_{2g-3}&-p_{2g-4}&\cdots&-p_{1}&0\end{pmatrix} (4.20)

Computing det(B)det(B) from (4.20), we get:

det(B)=d12d22dg22p12det(B)=d_{1}^{2}d_{2}^{2}\cdots d_{g-2}^{2}p_{1}^{2}

But BB is unimodular, so d12d22dg22p12=1d_{1}^{2}d_{2}^{2}\cdots d_{g-2}^{2}p_{1}^{2}=1, which implies di=±1d_{i}=\pm 1 and p1=±1p_{1}=\pm 1. As a result, we are able to perform changing basis within sublattice 2g1={y1,w1,,w2g3,t}=ker(δ~)\mathbb{Z}^{2g-1}=\mathbb{Z}\{y_{1},w_{1},\cdots,w_{2g-3},t\}=ker(\tilde{\delta}), to make BB into the standard form:

Bdiag{(0110),,(0110)}B\sim diag\{\begin{pmatrix}0&1\\ -1&0\end{pmatrix},\cdots,\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\}

Summing up, we can find a symplectic basis {x1,,xg,y1,,yg}\{x_{1},\cdots,x_{g},y_{1},\cdots,y_{g}\} of 2g\mathbb{Z}^{2g}, such that the intersection form BB is represented by:

B=x1y1++xgygB=x_{1}^{*}\wedge y_{1}^{*}+\cdots+x_{g}^{*}\wedge y_{g}^{*}

Moreover, δ(x1)\delta(x_{1}) generates /n\mathbb{Z}/n\mathbb{Z}, and δ(x2),,δ(xg),δ(y1),,δ(yg)=0\delta(x_{2}),\cdots,\delta(x_{g}),\delta(y_{1}),\cdots,\delta(y_{g})=0. By realization theorem, we can also find xi~,yi~π=π1(Σg)\tilde{x_{i}},\tilde{y_{i}}\in\pi=\pi_{1}(\Sigma_{g}), such that xi~xi\tilde{x_{i}}\to x_{i}, and yi~yi\tilde{y_{i}}\to y_{i}, under the Hurewicz map π1(Σg)H1(Σg,)\pi_{1}(\Sigma_{g})\to H_{1}(\Sigma_{g},\mathbb{Z}), and the geometric intersection number #(xi~,yj~)=δij\#(\tilde{x_{i}},\tilde{y_{j}})=\delta_{ij}.

Let {ai,bj}\{a_{i},b_{j}\} be the standard basis in H1(Σg,)H_{1}(\Sigma_{g},\mathbb{Z}), such that exist {αi,βj}π1(Σg){ai,bj}\{\alpha_{i},\beta_{j}\}\subset\pi_{1}(\Sigma_{g})\to\{a_{i},b_{j}\} with relation i=1g[αi,βj]=e\prod_{i=1}^{g}[\alpha_{i},\beta_{j}]=e. Then the basis {xi,yj}\{x_{i},y_{j}\} can be obtained from {ai,bj}\{a_{i},b_{j}\} by a symplectic transformation. Since the mapping class group ModgMod_{g} maps surjectively to the symplective group Sp2g()Sp_{2g}(\mathbb{Z}) [Put], [LM18], we can find ff, a homeomorphism of Σg\Sigma_{g}, such that f(ai)=xif_{*}(a_{i})=x_{i}, and f(bj)=yjf_{*}(b_{j})=y_{j} (here f:H1(Σg,)H1(Σg,)f_{*}:H_{1}(\Sigma_{g},\mathbb{Z})\to H_{1}(\Sigma_{g},\mathbb{Z})). Let xi~=f(αi)\tilde{x_{i}}=f_{*}(\alpha_{i}), and yj~=f(βj)\tilde{y_{j}}=f_{*}(\beta_{j}) (here f:ππf_{*}:\pi\to\pi, and xi~,yj~π\tilde{x_{i}},\tilde{y_{j}}\in\pi). Then we have i=1g[xi~,yi~]=e\prod_{i=1}^{g}[\tilde{x_{i}},\tilde{y_{i}}]=e, and xi~xi,yj~yj\tilde{x_{i}}\to x_{i},\tilde{y_{j}}\to y_{j} in H1(Σg,)H_{1}(\Sigma_{g},\mathbb{Z}). So these {xi~,yj~}\{\tilde{x_{i}},\tilde{y_{j}}\} is just the basis we wanted, proving the lemma.

Under the standard cyclic covering, the matrix for the deck transformation is I2(1In1)I2g2I_{2}\oplus\begin{pmatrix}\quad&1\\ I_{n-1}&\quad\end{pmatrix}\otimes I_{2g-2}, acting on H1(Σh,)H_{1}(\Sigma_{h},\mathbb{Z}). Applying this to our case π:DC\pi:D\to C, we obtain:

Corollary 4.7: σ\sigma is similar to I2(1In1)I2g2I_{2}\oplus\begin{pmatrix}\quad&1\\ I_{n-1}&\quad\end{pmatrix}\otimes I_{2g-2} over \mathbb{Z}.

Since we obtain similarity type over \mathbb{Z}, we can go further than we do in the first method. In fact, for any integer mm, not necessarily coprime with nn, by Corollary 4.7, we have σ(m)\sigma(m) is similar to I2(1In1)I2g2I_{2}\oplus\begin{pmatrix}\quad&1\\ I_{n-1}&\quad\end{pmatrix}\otimes I_{2g-2} over /m\mathbb{Z}/m\mathbb{Z}. From here on, the computation is almost the same as the third step of the first method. However, if gcd(m,n)>1gcd(m,n)>1, the case will get more complicated, and we won’t do the computation for it here.

5 Some Other Results

In this last section, we use the main result in section 4 to deduce some further corollaries.

We keep the assumption gcd(m,n)=1gcd(m,n)=1 throughout this section. To begin with, note that in the semi-direct product /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z}, the subgroup /m\mathbb{Z}/m\mathbb{Z} is unique. Therefore, for the covering ECE\to C, there is a unique intermidiate curve D:=E/(/m)D:=E/(\mathbb{Z}/m\mathbb{Z}), such that DCD\to C is an etale cyclic nn-covering. In other words, if we start our construction from non-isomorphic DD, (DCD\to C is etale nn-cyclic), we will obtain non-isomorphic curve EE.

It is easy for us to count: given a smooth projective curve CC, the number of curve DD up to isomorphism, such that DCD\to C is etale cyclic nn-covering. From section 2.2, we know it equals the number of primitive points in (/n)2g(\mathbb{Z}/n\mathbb{Z})^{\oplus 2g}, modulo scalars in (/n)×(\mathbb{Z}/n\mathbb{Z})^{\times}. Again, assume that /n=qn/qf\mathbb{Z}/n\mathbb{Z}=\prod_{q\mid n}\mathbb{Z}/q^{f}\mathbb{Z}. As in section 4.1, the number we are looking for is:

qnq2gf(1q2g)/qf(1q1)\prod_{q\mid n}q^{2gf}(1-q^{-2g})/q^{f}(1-q^{-1}) (5.1)

(5.1) equals to:

n2g1qn(1+q1++q2g+1)n^{2g-1}\prod_{q\mid n}(1+q^{-1}+\cdots+q^{-2g+1}) (5.2)

Combining these together, we obtain:

Corollary 5.1: For a given smooth projective curve CC of genus g2g\geq 2, the number of curve EE up to isomorphism, such that ECE\to C is etale non-abelian Galois with Galois group a semi-direct product /m/n\mathbb{Z}/m\mathbb{Z}\rtimes\mathbb{Z}/n\mathbb{Z} equals to Cg(m,n)C_{g}(m,n):

Cg(m,n)=m2g3n2g1pm,qn(gcd(n,p1)1)(1+p1++p2g+3)(1+q1++q2g+1)C_{g}(m,n)=m^{2g-3}n^{2g-1}\prod_{p\mid m,q\mid n}(gcd(n,p-1)-1)(1+p^{-1}+\cdots+p^{-2g+3})(1+q^{-1}+\cdots+q^{-2g+1})


The most special case of Corollary 5.1 might be m=pm=p, n=qn=q, with pqp\neq q are distinct prime numbers. In the case of qp1q\nmid p-1, Cg(p,q)=0C_{g}(p,q)=0; If qp1q\mid p-1, then Cg(p,q)=(q2g1)(1+p++p2g3)C_{g}(p,q)=(q^{2g}-1)(1+p+\cdots+p^{2g-3}). We get:

Corollary 5.2: Given curve CC as above. If qp1q\mid p-1, there are (q2g1)(1+p++p2g3)(q^{2g}-1)(1+p+\cdots+p^{2g-3}) many curve EE up to isomorphism such that ECE\to C is etale Galois with Galois group /p/q\mathbb{Z}/p\mathbb{Z}\rtimes\mathbb{Z}/q\mathbb{Z}.

References

  • [Ago20] Daniele Agostini. On the prym map for cyclic covers of genus two curves. Journal of Pure and Applied Algebra, 224(10):106384, 2020.
  • [BHPVdV15] Wolf Barth, Klaus Hulek, Chris Peters, and Antonius Van de Ven. Compact complex surfaces, volume 4. Springer, 2015.
  • [LM18] Justin Lanier and Dan Margalit. Normal generators for mapping class groups are abundant. arXiv preprint arXiv:1805.03666, 2018.
  • [MER03] S MERKULOV. Hatcher, a. algebraic topology (cambridge university press, 2002), 556 pp., 0 521 79540 0 (softback),£ 20.95, 0 521 79160 x (hardback),£ 60. Proceedings of the Edinburgh Mathematical Society, 46(2):511–512, 2003.
  • [Pom73] J Pomfret. Similarity of matrices over finite rings. Proceedings of the American Mathematical Society, 37(2):421–422, 1973.
  • [Put] Andrew Putman. The symplectic representation of the mapping class group is surjective.