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

Generic triviality of automorphism groups of complete intersections

Renjie Lyu School of Mathematical Sciences, Xiamen University, Xiamen 361005, China. [email protected]  and  Dingxin Zhang YMSC, Tsinghua University, Beijing 100086, China. [email protected]
Abstract.

We prove in most cases that a general smooth complete intersection in the projective space has no non-trivial automorphisms.

Introduction

Let kk be an algebraically closed field, and X𝐏knX\subset\mathbf{P}^{n}_{k} be a smooth closed subvariety. A linear automorphism is an automorphism of XX that extends to an automorphism of 𝐏n\mathbf{P}^{n}. We denote by AutL(X)\mathrm{Aut}_{L}(X) the group of linear automorphisms of XX.

The primary aim of this note is to demonstrate that the linear automorphism group of a general smooth complete intersection in a projective space is trivial with a few undetermined cases. For hypersurfaces, Matsumura and Monsky [10] showed that when n3n\geq 3 and d3d\geq 3, the group of linear automorphisms for a generic degree dd hypersurface in 𝐏n\mathbf{P}^{n} is trivial. The cases for n=2,d3n=2,d\geq 3 can be found in [8, § 10] and [12]. Benoist [1] proved that the group of linear automorphisms of a smooth complete intersection is finite, except for hyperquadrics. Building upon this result, Chen et al. [2] further derived the following.

Theorem 0.1 ([2, Thm. 1.3]).

Let (d1,,dc;n)(d_{1},\ldots,d_{c};n) be a sequence of natural numbers, with di2d_{i}\geq 2 and nc2n-c\geq 2. Let XX be a general smooth complete intersection in 𝐏kn\mathbf{P}^{n}_{k} of multidegree (d1,,dc)(2,2)(d_{1},\ldots,d_{c})\neq(2,2).

  1. (1)

    if kk has characteristic 0, then AutL(X)={1}\mathrm{Aut}_{L}(X)=\{1\}.

  2. (2)

    if kk has characteristic p>0p>0, then there exists r0r\geq 0, such that

    CardAutL(X)=pr.\mathrm{Card}\mathrm{Aut}_{L}(X)=p^{r}.

The assertion (2)(2) in Theorem 0.1 does not determine the number rr explicitly in terms of (d1,,dc;n)(d_{1},\ldots,d_{c};n). In this exposition, we present an alternative method for studying the automorphism group of a general complete intersection, and show that r=0r=0 in many cases. The main result is:

Theorem 0.2 (= Theorem 2.1).

Let kk be an algebraically closed field of characteristic p2p\neq 2. Let X𝐏knX\subset\mathbf{P}^{n}_{k} be a general smooth complete intersection of multidegree (d1,,dc)(d_{1},\ldots,d_{c}) with c<nc<n. Assume that (d1,,dc)(d_{1},\ldots,d_{c}) satisfies one of the following conditions

  • 3d1d2dc3\leq d_{1}\leq d_{2}\leq\cdots\leq d_{c} and pd1p\nmid d_{1};

  • c3c\geq 3 and 2=d1=d2=d3dc2=d_{1}=d_{2}=d_{3}\leq\cdots\leq d_{c}.

Then we have AutL(X)={1}\mathrm{Aut}_{L}(X)=\{1\} if (d1,,dc;n)(3;2)(d_{1},\ldots,d_{c};n)\neq(3;2).

We note that Javanpeykar and Loughran  [7, Lemma 2.13-2.14] obatined similar conclusion for complex complete intersections of multidegree (d1,,dc)(d_{1},\ldots,d_{c}) being (2,2,2)(2,2,2) or satisfying 3d1<d23\leq d_{1}<d_{2}\leq\cdots.

Our proof of Theorem 0.2 is inspired by Katz and Sarnak [8, §10-11]. They showed that the monodromy group of the universal family of smooth curves or smooth projective hypersurfaces is sufficiently big in a sense. As a consequence, it implies that a general member within the family has no automorphisms. We refine their method in general settings, and apply it to complete intersections.

Consider a smooth projective family of subvarieties of 𝐏n\mathbf{P}^{n} parameterized by a smooth kk-variety BB. The family is represented in the diagram

𝒳{\mathcal{X}}𝐏N×B{\mathbf{P}^{N}\times B}B.{B.}i\scriptstyle{i}π\scriptstyle{\pi}

Let :=i𝒪𝐏N(1)\mathcal{L}:=i^{*}\mathcal{O}_{\mathbf{P}^{N}}(1) be the natural polarization. In Section 1, we demonstrate the following criterion of triviality of linear automorphisms for a general member in the family.

Theorem 0.3 (= Theorem 1.2).

Assume that the following three conditions hold for the family π:𝒳B\pi:\mathcal{X}\to B.

  • The relative polarized automorphism group scheme AutB(𝒳,)\operatorname{\mathrm{Aut}}_{B}(\mathcal{X},\mathcal{L}) is finite and unramifed over BB.

  • There exists a closed point bBb\in B such that AutL(Xb)\mathrm{Aut}_{L}(X_{b}) acts faithfully on the \ell-adic cohomology He´tm(Xb;𝐐)\mathrm{H}^{m}_{\mathrm{\acute{e}t}}(X_{b};\mathbf{Q}_{\ell}) with dimXb=m\dim X_{b}=m.

  • The monodromy group associated to the lisse 𝐐\mathbf{Q}_{\ell}-sheaf (Rmπ𝐐)prim(R^{m}\pi_{\ast}\mathbf{Q}_{\ell})_{\mathrm{prim}} is sufficiently big (see Theorem 1.2(3) for the precise statement).

Then the linear automorphism group AutL(Xb)\mathrm{Aut}_{L}(X_{b}) associated to a general point bBb\in B is either trivial or isomorphic to 𝐙/2𝐙\mathbf{Z}/2\mathbf{Z}.

In Section 2, we show that the three conditions in Theorem 0.3 hold for most complete intersections. Suppose that π:𝒳B\pi\colon\mathcal{X}\to B is a family of smooth complete intersections of a given type. The group scheme AutB(𝒳,)\operatorname{\mathrm{Aut}}_{B}(\mathcal{X},\mathcal{L}) is finite over BB if the corresponding moduli stack is separated, which was explored by Benoist [1]. In cases the canonical divisor is ample or trivial, the separation isca direct consequence due to Matsusaka and Mumford [11]. The sufficient bigness of the monodromy group for complete intersections will be obtained by considering the vanishing cycles of Lefschetz pencils [3, §4-5]. Verifying the cohomological action by the automorphism group is faithful requires substantial effort and occupies the main part of this section. The advantage is that we need only verify the faithfullness property for a single member in the family. In our circumstance, we look into a special complete intersection of Fermat types (2.4). Lastly, we rule out the possibility of Aut(Xb)𝐙/2𝐙\mathrm{Aut}(X_{b})\simeq\mathbf{Z}/2\mathbf{Z} through a direct argument, given the assumption char(k)2\mathrm{char}(k)\neq 2.

It is likely that conclusions in Theorem 0.2 can be strengthened. The assumption pd1p\nmid d_{1} is indispensible for considering complete intersections of Fermat type. One may check the faithfulness property for other special complete intersections. For the cases of complete intersections that not covered in Theorem 0.2, see Remark 2.3.

The article [2] also confirmed the faithfulness of the cohomological action. The main difference is: they directly proved that a general complex complete intersection has no non-trivial automorphisms, then use this result to deduce all complex complete intersections have the faithfulness property; In contrast, we gain the generic triviality result by verifying the faithfulness property for a single smooth complete intersection.

The cohomological method may find applications that are not limited to complete intersections in projective spaces. One example is offered below:

Theorem 0.4 (= Theorem 1.4).

Let PP be a smooth, complex projective variety. Let ωP\omega_{P} be the canonical sheaf on PP, and let \mathscr{L} be a very ample invertible sheaf such that ωP\mathscr{L}\otimes\omega_{P} remains very ample. Then, a general smooth section of \mathscr{L} has no non-trivial automorphisms.

Acknowledgements

We are grateful to Wenfei Liu for communications on this work. We would like to thank Xi Chen for comments on the manuscript.

1. A criterion of generic triviality of automorphisms

Situation 1.1.

Let kk be an algebraically closed field. Consider a smooth projective family of algebraic varieties

𝒳{\mathcal{X}}𝐏N×B{\mathbf{P}^{N}\times B}B{B}π\scriptstyle{\pi}i\scriptstyle{i}

parameterized by a smooth, irreducible kk-variety BB. Let :=i𝒪𝐏N(1)\mathcal{L}:=i^{*}\mathcal{O}_{\mathbf{P}^{N}}(1) be the natural polarization. Denote AutB(𝒳,)\operatorname{\mathrm{Aut}}_{B}(\mathcal{X},\mathcal{L}) as the group scheme of automorphisms relative to the family π:𝒳B\pi:\mathcal{X}\to B, preserving \mathcal{L}. Let 𝒢\mathcal{G} be a closed subgroup scheme of AutB(𝒳,)\operatorname{\mathrm{Aut}}_{B}(\mathcal{X},\mathcal{L}). Let XbX_{b} be the fiber π1(b)\pi^{-1}(b) over a closed point bBb\in B, and GbG_{b} be the fiber of 𝒢\mathcal{G} over bb.

Fix a prime number \ell different from the characteristic of kk. Let He´tm(Xb;𝐐)\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell}) be the \ell-adic cohomology group. The cup-product on He´tm(Xb;𝐐)\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell}) gives a non-degenerate symmetric (resp. anti-symmetric) form ψ\psi if mm is even (resp. odd). In this paper, the primitive cohomology He´tm(Xb;𝐐)prim\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell})_{\textrm{prim}} is defined to be the orthogonal complement of the image of the restriction map He´tm(𝐏N;𝐐)He´tm(Xb;𝐐)\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(\mathbf{P}^{N};\mathbf{Q}_{\ell})\to\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell}). Let π1(B,b)\pi_{1}(B,b) denote the étale fundamental group for of BB at the point bb.

Theorem 1.2.

Consider the notation in Situation 1.1. Suppose that the following hypotheses hold.

  1. (1)

    The group scheme 𝒢B\mathcal{G}\to B is finite and unramifed over BB.

  2. (2)

    There exists a closed point bBb\in B such that the map GbAut(He´tm(Xb;𝐐))G_{b}\to\operatorname{Aut}(\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell})) is injective.

  3. (3)

    The geometric monodromy group, that is, the Zariski closure of the image of the monodromy representation

    ρ:π1(B,b)Aut(He´tm(Xb;𝐐)prim,ψ),\rho\colon\pi_{1}(B,b)\to\operatorname{Aut}(\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell})_{\mathrm{prim}},\psi),

    equals the full symplectic group Sp(H,ψ)\mathrm{Sp}(\mathrm{H},\psi) if mm is odd or the orthogonal group O(H,ψ)\mathrm{O}(\mathrm{H},\psi) if mm is even, where H\mathrm{H} represents the 𝐐\mathbf{Q}_{\ell}-vector space He´tm(Xb;𝐐)prim\mathrm{H}_{\mathrm{\acute{e}t}}^{m}(X_{b};\mathbf{Q}_{\ell})_{\mathrm{prim}}.

Then GbG_{b} is either a trivial group or isomorphic to 𝐙/2𝐙\mathbf{Z}/2\mathbf{Z} for a general bBb\in B.

Note that Hypotheses 1.2(13) are not sufficient to deduce the triviality of GbG_{b}. For instance, let BB be the space of 66 distinct ordered points on 𝐏1\mathbf{P}^{1}, 𝒳B\mathcal{X}\to B is the family of genus 22 curves branching over the given points, and 𝒢=AutB(𝒳)\mathcal{G}=\operatorname{Aut}_{B}(\mathcal{X}). Then every XbX_{b} has an automorphism of order 22, although the finiteness of automorphism groups, the faithfulness of Aut(Xb)\operatorname{Aut}(X_{b}) acting on H1(Xb)\mathrm{H}^{1}(X_{b}), and the bigness of the geometric monodromy group are all satisfied.

Proof of Theorem 1.2.

The morphism pp is finite and unramified. The fiber group GbG_{b} over any bBb\in B is thus finite and discrete, and the order |Gb||G_{b}| is equal to the rank of the finite 𝒪B\mathcal{O}_{B}-module p𝒪𝒢p_{*}\mathcal{O}_{\mathcal{G}} at the stalk bb. It follows that the function b|Gb|b\mapsto|G_{b}| is upper semi-continuous. Then there is a non-empty open subset on which the number |Gb||G_{b}| reaches the minimum. Consider the closed subgroup scheme

𝒵:=Ker(𝒢GL(Rmπ𝐐))\mathcal{Z}\colon=\operatorname{Ker}(\mathcal{G}\to\mathrm{GL}(R^{m}\pi_{*}\mathbf{Q}_{\ell}))

of automorphisms acting trivially on the cohomology. Then 𝒵B\mathcal{Z}\to B remains a finite and unramified morphism. By the previous argument, let BBB^{\prime}\subset B be the open subset on which the order function b|Zb|b\mapsto|Z_{b}| reaches the minimum. By Hypothesis 1.2(2), the minimum must be 11, i.e. 𝒵|B\mathcal{Z}|_{B^{\prime}} is the identity group scheme.

By the generic flatness, there exists a non-empty open subset UBU\subset B such that morphism p:𝒢Bp\colon\mathcal{G}\to B is flat over UU. Then pp is a finite étale morphism over UU. Since BB is irreducible, UU must meet with BB^{\prime}. In conclusion, we can obtain a Zarski open dense subset UBU\subset B such that 𝒢\mathcal{G} is a finite étale cover over UU, and the fiber group GbG_{b} for any bUb\in U acts faithfully on He´tm(Xb;𝐐)\mathrm{H}^{m}_{\mathrm{\acute{e}t}}(X_{b};\mathbf{Q}_{\ell}).

Let bUb\in U be a closed point. The induced map π1(U,b)π1(B,b)\pi_{1}(U,b)\to\pi_{1}(B,b) shall be surjective. By Hypothesis 1.2(3), the geometric monodromy group of π1(U,b)\pi_{1}(U,b) equals to Sp(H,ψ)\mathrm{Sp}(\mathrm{H},\psi) or O(H,ψ)\mathrm{O}(\mathrm{H},\psi). Let VUV\to U be a finite étale base change such that the pullback 𝒢|V\mathcal{G}|_{V} becomes a constant group scheme over VV. Under a finite base change, the identity component of the geometric monodromy group does not change. Since the symplectic group or the orthogonal group is connected, the geometric monodromy group of π1(V,b)\pi_{1}(V,b) is still Sp(H,ψ)\mathrm{Sp}(\mathrm{H},\psi) or O(H,ψ)\mathrm{O}(\mathrm{H},\psi). Here we abusively denote bb as a closed point in VV lying over bUb\in U.

As 𝒢|V\mathcal{G}|_{V} is a constant group scheme, the monodromy action of π1(V,b)\pi_{1}(V,b) on GbG_{b} is invariant. Viewing GbG_{b} as a subgroup of GL(H)\mathrm{GL}(\mathrm{H}), we have

ρ(γ)gρ(γ)1=g,gGb,γπ1(V,b)\rho(\gamma)\cdot g\cdot\rho(\gamma)^{-1}=g,~{}\forall g\in G_{b},~{}\forall\gamma\in\pi_{1}(V,b)

where ρ\rho is the monodromy action of π1(V,b)\pi_{1}(V,b). A linear automorphism on H\mathrm{H} that commutes with all symplectic or orthogonal matrices must be the scalar action by some λ𝐐\lambda\in\mathbf{Q}^{*}_{\ell}. Assume that the degree of the finite étale covering 𝒢U\mathcal{G}\to U is rr, that is, the order of GbG_{b} is rr. Then λ\lambda is a rr-th root of unity in 𝐐\mathbf{Q}_{\ell}.

By Hensel’s lemma, we know that any root of unity in 𝐐\mathbf{Q}_{\ell} is an (1)(\ell-1)-th root of unity for \ell odd or ±1\pm 1 for =2\ell=2. In the case of char(k)2\mathrm{char}(k)\neq 2, let =2\ell=2. Then λ=±1\lambda=\pm 1. When char(k)=2\mathrm{char}(k)=2 and rr is odd, we can select an odd prime \ell such that 1\ell-1 is coprime to rr. Then λ1=1\lambda^{\ell-1}=1 implies λ=1\lambda=1 since (1,r)=1(\ell-1,r)=1. When char(k)=2\mathrm{char}(k)=2 and rr is even, choose the odd prime \ell such that 12\frac{\ell-1}{2} is coprime to rr. Then we have λ12=±1\lambda^{\frac{\ell-1}{2}}=\pm 1, which implies λ=±1\lambda=\pm 1 since (12,r)=1(\frac{\ell-1}{2},r)=1. In conclusion, the order of the group GbG_{b} for bUb\in U is bounded by 22. ∎

As a simple application of Theorem 1.2, we show that a sufficiently ample, sufficiently general hypersurface in any smooth complex projective variety is free of automorphisms. To begin with, we need a lemma.

Lemma 1.3.

Let XX be an m–dimensional, smooth, complex projective variety with very ample canonical bundle. Then the natural map Aut(X)Aut(Hm(X;𝐐))\operatorname{Aut}(X)\to\operatorname{Aut}(\mathrm{H}^{m}(X;\mathbf{Q})) is injective.

Proof.

Since the canonical bundle ωX\omega_{X} is very ample, we consider the canonical embedding X𝐏(H0(X,ωX))X\hookrightarrow\mathbf{P}(\mathrm{H}^{0}(X,\omega_{X})^{\vee}). Let ff be an automorphism of XX such that ff^{\ast} is the identity. By the Hodge decomposition ff^{\ast} acts identically on H0(X,ωX)\mathrm{H}^{0}(X,\omega_{X}), which induces the idenity on 𝐏(H0(X,ωX))\mathbf{P}(\mathrm{H}^{0}(X,\omega_{X})^{\vee}). Via the canonical embedding, we have f=IdXf=\operatorname{Id}_{X}. ∎

Theorem 1.4.

Let PP be a smooth, complex projective variety. Let ωP\omega_{P} be the canonical sheaf on PP, and let \mathscr{L} be a very ample invertible sheaf such that ωP\mathscr{L}\otimes\omega_{P} remains very ample. Then a general smooth section of \mathscr{L} has no non-trivial automorphisms.

Proof.

We need to check Hypotheses  1.2(13) are satisfied. Let XX be the zero locus of any smooth transverse section of \mathscr{L}. By the adjunction formula and our assumption, the canonical divisor ωXωP\omega_{X}\cong\mathscr{L}\otimes\omega_{P} is very ample.

Let Hm(X;𝐂)p\mathrm{H}^{m}(X;\mathbf{C})_{p} be the kernel of the cup-product operator

ωX:Hm(X;𝐂)Hm+2(X;𝐂)\cup\omega_{X}\colon\mathrm{H}^{m}(X;\mathbf{C})\to\mathrm{H}^{m+2}(X;\mathbf{C})

with m=dimXm=\dim X. Let QQ be the intersection form. The Hodge-Riemann bilinear relation shows that the Hermitian form h:=p+q=mipqQ(,¯)h:=\oplus_{p+q=m}i^{p-q}Q(\cdot,\bar{\cdot}) is positive definite on

Hm(X;𝐂)p=p+q=mHp,qHm(X;𝐂)p.\mathrm{H}^{m}(X;\mathbf{C})_{p}=\bigoplus_{p+q=m}\mathrm{H}^{p,q}\cap\mathrm{H}^{m}(X;\mathbf{C})_{p}.

Any automorphism of XX preserves ωX\omega_{X} and hh. By Lemma 1.3, Aut(X)\operatorname{Aut}(X) can be regarded as a subgroup of the unitary group U(Hm(X;𝐂)p,h)\mathrm{U}(\mathrm{H}^{m}(X;\mathbf{C})_{p},h). Additionally, Aut(X)\operatorname{Aut}(X) acts on the integral cohomology Hn(X;𝐙)\mathrm{H}^{n}(X;\mathbf{Z}). Therefore, Aut(X)\operatorname{\mathrm{Aut}}(X) is a discrete subgroup in a compact group, which is necessarily finite.

Let 𝒳B\mathcal{X}\to B be the family of smooth sections of \mathscr{L}. The above discussion implies that the relative automorphism group scheme

AutB(𝒳)B\mathrm{Aut}_{B}(\mathcal{X})\to B

is a quasi-finite morphism. The hypersurface XX is not uniruled since ωX\omega_{X} is ample. Matsusaka-Mumford’s theorem [11] ensures that the morphism Aut(𝒳/B)B\mathrm{Aut}(\mathcal{X}/B)\to B is proper and hence finite. It is unramifed since a group scheme over the complex number is always smooth. Thus, Hypothesis 1.2(1) is verified.

The \ell-adic cohomology of a complex variety is isomorphic to the Betti cohomology of the associated analytic space. In the following, when we apply Theorem 1.2, we will use Betti cohomology in place of the \ell-adic cohomology. Then Lemma 1.3 affirms Hypothesis 1.2(2)

To show the monodromy group is as big as possible, we consider a Lefschetz pencil DD of hyperplane sections |||\mathscr{L}| on PP such that DD passing through [X]||[X]\in|\mathscr{L}|. Vanishing cycles for the Lefschetz pencil DD are conjugate under the monodromy action, see [3] or [16, §2] for the details.

If mm is odd, Kazhdan-Margulis theorem [3, Théorème 5.10] shows that the geometric monodromy group associated to DD is the full sympelctic group Sp(H,ψ)\mathrm{Sp}(\mathrm{H},\psi) where HHprimm(X;𝐐)\mathrm{H}\subset\mathrm{H}^{m}_{\textrm{prim}}(X;\mathbf{Q}) stands for the subspace generated by vanishing cycles. If m=2pm=2p is even, the geometric monodromy group is either a finite group or the full orthogonal group O(H,ψ)\mathrm{O}(\mathrm{H},\psi), see [4, Théorème 4.4.2]. The case of finite group occurs when the intersection form ψ\psi is definite. By the Hodge index theorem, it deduces that Hp,q(X)=0\mathrm{H}^{p,q}(X)=0 unless p=qp=q. However, in our case Hm,0(X)\mathrm{H}^{m,0}(X) is non-trivial since the canonical divisor ωX\omega_{X} is very ample. Hence monodromy group is sufficently big.

As a consequence of Theorem 1.2, a general section XX of \mathscr{L} either has no non-trivial automorphism, or admits an involution that acts by 1-1 on Hm(X)\mathrm{H}^{m}(X). The second possibility is readily ruled out by the same argument in the proof of Lemma 1.3. ∎

2. Automorphism of complete intersection in projective spaces.

Let kk be an algebraically closed field. Let (d1,,dc;n)(d_{1},\ldots,d_{c};n) be a sequence of positive integers with 2didi+12\leq d_{i}\leq d_{i+1}, n>cn>c. We say a complete intersection X𝐏knX\subset\mathbf{P}^{n}_{k} is of type (d1,,dc;n)(d_{1},\ldots,d_{c};n) if the ideal of XX is generated by cc homogeneous polynomials with the prescribed degrees {d1,,dc}\{d_{1},\ldots,d_{c}\}. The goal of this section is proving the following:

Theorem 2.1.

Let kk be an algebraically closed field of characteristic p2p\neq 2. Let X𝐏knX\subset\mathbf{P}^{n}_{k} be a general smooth complete intersection of multidegree (d1,,dc)(d_{1},\ldots,d_{c}) with c<nc<n. Assume that (d1,,dc)(d_{1},\ldots,d_{c}) satisfies one of the following conditions

  • 3d1d2dc3\leq d_{1}\leq d_{2}\leq\cdots\leq d_{c} and pd1p\nmid d_{1};

  • c3c\geq 3 and 2=d1=d2=d3dc2=d_{1}=d_{2}=d_{3}\leq\cdots\leq d_{c}.

Then we have AutL(X)={1}\mathrm{Aut}_{L}(X)=\{1\} if (d1,,dc;n)(3;2)(d_{1},\ldots,d_{c};n)\neq(3;2).

We have the following simple observation.

Observation 2.2.

Let XX be a general complete intersection of multidegree (d1,,dc)(d_{1},\ldots,d_{c}), and F1,,FcF_{1},\ldots,F_{c} denote the defining polynomials of XX with degFi=di\deg F_{i}=d_{i}. Suppose that rr is the maximal number such that d1==dr<dr+1d_{1}=\cdots=d_{r}<d_{r+1}. Then the action of any linear automorphism of XX preserves the ideal (F1,,Fr)(F_{1},\ldots,F_{r}). Regard XX as a subscheme of the complete intersection YY defined by (F1,,Fr)(F_{1},\ldots,F_{r}), then we have AutL(X)AutL(Y)\operatorname{\mathrm{Aut}}_{L}(X)\subset\operatorname{\mathrm{Aut}}_{L}(Y). Therefore, to prove AutL(X)={1}\operatorname{\mathrm{Aut}}_{L}(X)=\{1\}, it suffices to prove that a general complete intersection with equal multidegrees has no non-trivial automorphisms.

Remark 2.3.

The cases of complete intersections of type (d1,,dc)(d_{1},\ldots,d_{c}) with 2=d1d2<d3dc2=d_{1}\leq d_{2}<d_{3}\leq\cdots\leq d_{c} are not covered in Theorem 2.1. According to Observation  2.2, such cases correspond to general hyperquadrics or complete intersections of two quadrics, which admit non-trivial linear automorphisms, see [13]. Hence, the proof strategy in Observation 2.2 does not work directly. Instead, one may attempt to characterize the autmorphism group for complete intersections of type (2,d,,d;n)(2,d,\ldots,d;n) and (2,2,d,,d;n)(2,2,d,\ldots,d;n) with d>2d>2.

To apply Theorem 1.2 to complete intersections with equal multidegrees, we will verify Hypothesis  1.2(2) for complete intersection of Fermat type. In the following, we will describe their automorphism group and cohomology group, see Proposition 2.5 and Theorem 2.10, and prove that the automorphism group acts faithfully on the cohomology group in Proposition 2.11.

Automorphism of complete intersections of Fermat type

Let kk be an algebraically closed field. Fix natural numbers n3,r2n\geq 3,r\geq 2, and d2d\geq 2, where dd is prime to char(k)\mathrm{char}(k). Let X:=Xn,r,d𝐏knX:=X_{n,r,d}\subset\mathbf{P}^{n}_{k} be the complete intersection defined by the following Fermat equations:

(2.4) {x0d++xnd=0,λ0x0d++λnxnd=0,λ0r1x0d++λnr1xnd=0,\left\{\begin{matrix}x_{0}^{d}+\cdots+x_{n}^{d}=0,\\ \lambda_{0}x_{0}^{d}+\cdots+\lambda_{n}x_{n}^{d}=0,\\ \vdots\\ \lambda_{0}^{r-1}x_{0}^{d}+\cdots+\lambda_{n}^{r-1}x_{n}^{d}=0,\end{matrix}\right.

where {λ0,,λn}\{\lambda_{0},\ldots,\lambda_{n}\} are pairwise distinct elements in kk^{*}. It is known that XX is non-singular [15, Prop. 2.4.1]. Let μd\mu_{d} be the group of dd-th roots of unity in kk, and μdn+1\mu_{d}^{n+1} be the (n+1)(n+1)-th product of μd\mu_{d}. Denote by GndG_{n}^{d} the quotient group μdn+1/Δ(μd)\mu_{d}^{n+1}/\Delta(\mu_{d}), where Δ:μdμdn+1\Delta:\mu_{d}\to\mu_{d}^{n+1} is the diagonal embedding. The group GndG_{n}^{d} acts on the variety XX as follows:

[ξ0,,ξn](x0::xn)(ξ0x0::ξnxn),ξiμd[\xi_{0},\ldots,\xi_{n}]\cdot(x_{0}:\ldots:x_{n})\mapsto(\xi_{0}x_{0}:\ldots:\xi_{n}x_{n}),~{}\xi_{i}\in\mu_{d}
Proposition 2.5.

Let X:=Xn,r,d𝐏knX:=X_{n,r,d}\subset\mathbf{P}^{n}_{k} be the Fermat complete intersection defined by the equations above, and let AutL(X)\operatorname{\mathrm{Aut}}_{L}(X) be the group of linear automorphisms of XX. Then AutL(X)\operatorname{\mathrm{Aut}}_{L}(X) fits into an exact sequence

1GndAutL(X)𝔖n+1,1\to G^{d}_{n}\to\operatorname{\mathrm{Aut}}_{L}(X)\to\mathfrak{S}_{n+1},

where 𝔖n+1\mathfrak{S}_{n+1} is the permutation group of n+1n+1 elements. Moreover, if the coefficients λ0,,λn\lambda_{0},\ldots,\lambda_{n} correspond to a general point in the kk-vector space kn+1k^{n+1}, then Gnd=AutL(X)G^{d}_{n}=\operatorname{\mathrm{Aut}}_{L}(X).

Proof.

Let (x0::xn)(x_{0}\colon\dots\colon x_{n}) be a coordinate in 𝐏n\mathbf{P}^{n}. The action of a linear automorphism gAutL(X)g\in\operatorname{\mathrm{Aut}}_{L}(X) on xix_{i} transforms xix_{i} to a linear form

gi:=g(xi)=j=0naijxj,aijk.g_{i}:=g^{*}(x_{i})=\sum_{j=0}^{n}a_{ij}x_{j},~{}a_{ij}\in k.

Since the ideal of XX is generated by the Fermat equations (2.4), we observe

(2.6) g(x0d++xnd)=i=0ngid=i=0ncixid, for some cik.g^{*}(x_{0}^{d}+\cdots+x_{n}^{d})=\sum_{i=0}^{n}g_{i}^{d}=\sum_{i=0}^{n}c_{i}x_{i}^{d},\text{~{}for some~{}}c_{i}\in k.

Differentiating both sides of (2.6) with respect to xj\frac{\partial}{\partial x_{j}}, we obtain

di=0ngid1gixj=di=0ngid1aij=dcjxjd1.d\sum_{i=0}^{n}g^{d-1}_{i}\cdot\frac{\partial g_{i}}{\partial x_{j}}=d\sum_{i=0}^{n}g^{d-1}_{i}a_{ij}=d\cdot c_{j}\cdot x_{j}^{d-1}.

Focusing on the coefficients of the monomial x0d1x_{0}^{d-1} on both sides. There is the relation

di=0naijai0d1={0,j0;dc0,j=0.d\cdot\sum_{i=0}^{n}a_{ij}a_{i0}^{d-1}=\begin{cases}0,&j\neq 0;\\ d\cdot c_{0},&j=0.\end{cases}

Let AA represent the matrix (aij)(a_{ij}). Since dd is invertible in the field kk, we have

(a00d1,,an0d1)A=(c0,0,,0).(a_{00}^{d-1},\ldots,a_{n0}^{d-1})\cdot A=(c_{0},0,\ldots,0).

Similarly,

g(λ0x0d++λnxnd)=i=0nλigid=i=0ncixid, for some cik.g^{*}(\lambda_{0}x_{0}^{d}+\cdots+\lambda_{n}x_{n}^{d})=\sum_{i=0}^{n}\lambda_{i}g_{i}^{d}=\sum_{i=0}^{n}c_{i}^{\prime}x_{i}^{d},\text{~{}for some~{}}c^{\prime}_{i}\in k.

Following the same discussion as above, we have

(λ0a00d1,,λnan0d1)A=(c0,0,,0).(\lambda_{0}a_{00}^{d-1},\ldots,\lambda_{n}a_{n0}^{d-1})\cdot A=(c_{0}^{\prime},0,\ldots,0).

Note that the matrix AA is invertible, and c0,c0c_{0},c^{\prime}_{0} are non-zero. Hence (a00d1,,an0d1)(a_{00}^{d-1},\ldots,a_{n0}^{d-1}) is proportional to (λ0a00d1,,λnan0d1)(\lambda_{0}a_{00}^{d-1},\ldots,\lambda_{n}a_{n0}^{d-1}). Since λ0,λ1,,λn\lambda_{0},\lambda_{1},\ldots,\lambda_{n} are pairwise distinct, only one element among {a00,,an0}\{a_{00},\ldots,a_{n0}\} is non-zero. Considering the coefficients of xjd1x_{j}^{d-1} for each jj, the same conclusion holds for each tuple (a0j,,anj)(a_{0j},\ldots,a_{nj}). Therefore, gg determines a permutation σg𝔖n+1\sigma_{g}\in\mathfrak{S}_{n+1} such that g(xi)=aσg(i)ixig^{*}(x_{i})=a_{\sigma_{g}(i)i}x_{i}. The assignment gσgg\mapsto\sigma_{g} thus establishes a group homomorphism

σ:AutL(X)𝔖n+1.\sigma:\operatorname{\mathrm{Aut}}_{L}(X)\to\mathfrak{S}_{n+1}.

In the following, we demonstrate that the kernel of σ\sigma is GndG^{d}_{n}. Suppose that gKerσg\in\operatorname{\mathrm{Ker}}\sigma. Then gg is represented by a diagonal matrix (aii)0in(a_{ii})_{0\leq i\leq n}. The action of gg on the Fermat polynomials {i=0nλijxid|0jr1}\{\sum\limits_{i=0}^{n}\lambda_{i}^{j}x_{i}^{d}~{}|~{}0\leq j\leq r-1\} yields an r×rr\times r matrix BB satisfying the relationship

(2.7) B(11λ0λnλ0r1λnr1)=(11λ0λnλ0r1λnr1)(a00dannd)B\cdot\begin{pmatrix}1&\cdots&1\\ \lambda_{0}&\cdots&\lambda_{n}\\ \vdots&&\vdots\\ \lambda_{0}^{r-1}&\cdots&\lambda_{n}^{r-1}\end{pmatrix}=\begin{pmatrix}1&\cdots&1\\ \lambda_{0}&\cdots&\lambda_{n}\\ \vdots&&\vdots\\ \lambda_{0}^{r-1}&\cdots&\lambda_{n}^{r-1}\end{pmatrix}\cdot\begin{pmatrix}a^{d}_{00}&&\\ &\ddots&\\ &&a^{d}_{nn}\end{pmatrix}

Let us view each column(1,λi,,λir1)𝖳(1,\lambda_{i},\ldots,\lambda_{i}^{r-1})^{\mathsf{T}} as a vector viv_{i} in the rr-dimensional vector space krk^{r}. Then BB represents a linear map with eigenvectors {v0,,vn}\{v_{0},\ldots,v_{n}\} and eigenvalues {a00d,,annd}\{a_{00}^{d},\ldots,a_{nn}^{d}\}. Note that any rr elements among {v0,,vn}\{v_{0},\ldots,v_{n}\} form a basis of krk^{r} since the basis corresponds to a non-degenerate Vandermonde matrix. It implies that aiid=ajjda_{ii}^{d}=a_{jj}^{d} for all i,ji,j. Consequently, the matrix (aii)(a_{ii}) can be represented by

(ξ0a00ξna00),for someξiμd.\begin{pmatrix}\xi_{0}\cdot a_{00}&&\\ &\ddots&\\ &&\xi_{n}\cdot a_{00}\end{pmatrix},~{}\textrm{for some}~{}\xi_{i}\in\mu_{d}.

Hence, the action of gg on XX is equivalent to the action of [ξ0,,ξn]Gnd[\xi_{0},\cdots,\xi_{n}]\in G_{n}^{d}.

Now let us prove the final assertion. Given a linear automorphism gg, let σ\sigma denote the permutation σ(g)𝔖n+1\sigma(g)\in\mathfrak{S}_{n+1}, and (aσ(i)i)(a_{\sigma(i)i}) be the matrix representing gg. Again, the action of gg^{*} on the Fermat polynomials (2.4) yields an r×rr\times r matrix B:=(bij)0i,jr1B:=(b_{ij})_{0\leq i,j\leq r-1} such that

B(11λ0λnλ0r1λnr1)=(11λ0λnλ0r1λnr1)(aσ(i)i).B\cdot\begin{pmatrix}1&\cdots&1\\ \lambda_{0}&\cdots&\lambda_{n}\\ \vdots&&\vdots\\ \lambda_{0}^{r-1}&\cdots&\lambda_{n}^{r-1}\end{pmatrix}=\begin{pmatrix}1&\cdots&1\\ \lambda_{0}&\cdots&\lambda_{n}\\ \vdots&&\vdots\\ \lambda_{0}^{r-1}&\cdots&\lambda_{n}^{r-1}\end{pmatrix}\cdot(a_{\sigma(i)i}).

By this relation, the first two rows of the matrix BB correspond to two polynomials

q(t)=j=0r1b0jtj,p(t)=j=0r1b1jtjq(t)=\sum_{j=0}^{r-1}b_{0j}t^{j},~{}p(t)=\sum_{j=0}^{r-1}b_{1j}t^{j}

satisfying the interpolation data

q(λi)=aσ(i)i,p(λi)=λσ(i)aσ(i)i, for 0in.q(\lambda_{i})=a_{\sigma(i)i},~{}p(\lambda_{i})=\lambda_{\sigma(i)}a_{\sigma(i)i},\textrm{~{}for~{}}0\leq i\leq n.

Consider the Lagrangian polynomial

L(t):=k=0naσ(k)k0jn,jktλjλkλjL(t):=\sum_{k=0}^{n}a_{\sigma(k)k}\prod_{0\leq j\leq n,j\neq k}\frac{t-\lambda_{j}}{\lambda_{k}-\lambda_{j}}

which interpolates the given data (λi,ασ(i)i)(\lambda_{i},\alpha_{\sigma(i)i}). Assume that the constants {ασ(i)i}0in\{\alpha_{\sigma(i)i}\}_{0\leq i\leq n} are not all equal. It is known that, for a generic choice of the tuple (λ0,,λn)(\lambda_{0},\ldots,\lambda_{n}) in kn+1k^{n+1}, the degree of L(t)L(t) is nn. Hence L(t)=q(t)L(t)=q(t) since L(t)q(t)L(t)-q(t) contains n+1n+1 distinct roots. However, we have degq(t)=r1<n\deg q(t)=r-1<n as a contradiction. Therefore, all aσ(i)ia_{\sigma(i)i} must be equal, and q(t)q(t) is the constant polynomial b00b_{00}. It follows that the second interpolation data becomes

p(λi)=b00λσ(i).p(\lambda_{i})=b_{00}\lambda_{\sigma(i)}.

Applying the Lemma 2.8 below, such a polynomial p(t)p(t) exists for a generic choice of (λ0,,λn)(\lambda_{0},\ldots,\lambda_{n}) in kn+1k^{n+1} only if σ=Id\sigma=\operatorname{Id}. Thus, the last assertion follows. ∎

Lemma 2.8.

Let σ𝔖n+1\sigma\in\mathfrak{S}_{n+1} be a non-trivial permutation, and let rr be an integer with 1<r<n1<r<n. Suppose that (λ0,,λn)kn+1(\lambda_{0},\ldots,\lambda_{n})\in k^{n+1} is a general point. Then there exists no polynomial function p(t)k[t]p(t)\in k[t] with degp(t)r1\deg p(t)\leq r-1 that interpolates the n+1n+1 data points (λ0,λσ(0)),,(λn,λσ(n))(\lambda_{0},\lambda_{\sigma(0)}),\ldots,(\lambda_{n},\lambda_{\sigma(n)}), i.e.,

p(λi)=λσ(i),0in.p(\lambda_{i})=\lambda_{\sigma(i)},0\leq i\leq n.
Proof.

We identify the space of one-variable polynomials of degree at most r1r-1 with the rr-dimensional vector space krk^{r}. We consider the incidence subvariety

Γσ:={(p(t),(y0,,yn))kr×kn+1|p(yi)yσ(i)=0,0in}.\Gamma_{\sigma}:=\{(p(t),(y_{0},\ldots,y_{n}))\in k^{r}\times k^{n+1}~{}|~{}p(y_{i})-y_{\sigma(i)}=0,~{}0\leq i\leq n\}.

Let π1:Γσkr\pi_{1}:\Gamma_{\sigma}\to k^{r} and π2:Γσkn+1\pi_{2}:\Gamma_{\sigma}\to k^{n+1} be natural projections. Suppose that (p(t),(λ0,,λn))(p(t),(\lambda_{0},\ldots,\lambda_{n})) is a point in Γσ\Gamma_{\sigma}, and NN is the order of σ\sigma, then

pN(λi)=λσN(i)=λi,0in.p^{\circ N}(\lambda_{i})=\lambda_{\sigma^{N}(i)}=\lambda_{i},~{}\forall~{}0\leq i\leq n.

Case 1. The polynoimal pN(t)tp^{\circ N}(t)-t is non-zero. Then pN(t)tp^{\circ N}(t)-t has finite roots. Hence the possiblity of {λ0,,λn}\{\lambda_{0},\ldots,\lambda_{n}\} are finitely many. It follows that π1\pi_{1} is a quasi-finite map, and the dimension of Γσ\Gamma_{\sigma} is rr. Then π2(Γσ)\pi_{2}(\Gamma_{\sigma}) is a proper subset in kn+1k^{n+1}. Therefore, for a generic (λ0,,λn)kn+1(\lambda_{0},\ldots,\lambda_{n})\in k^{n+1}, there exists no p(t)krp(t)\in k^{r} satisfying the condition p(λi)=λσ(i),0inp(\lambda_{i})=\lambda_{\sigma(i)},0\leq i\leq n.

Case 2. The polynoimal pN(t)tp^{\circ N}(t)-t is zero. This situation occurs only if p(t)p(t) is a linear form α+βt\alpha+\beta t with βN=1\beta^{N}=1. All such linear forms consitutes a one-dimensional subset Ξkr\Xi\subset k^{r}. We claim that dimπ11(Ξ)<n+1\dim\pi_{1}^{-1}(\Xi)<n+1. The fiber of π1\pi_{1} over a linear form α+βt\alpha+\beta t is a subset in kn+1k^{n+1} defined by n+1n+1 equations

α+βyi=yσ(i),0in.\alpha+\beta y_{i}=y_{\sigma(i)},0\leq i\leq n.

Since n>r2n>r\geq 2 and σId\sigma\neq\operatorname{Id}, these equations impose at least two constraints on kn+1k^{n+1}. Therefore, the dimension of π11(Ξ)\pi_{1}^{-1}(\Xi) is less than n+1n+1. In conclusion, for a generic choice of (λ0,,λn)kn+1(\lambda_{0},\ldots,\lambda_{n})\in k^{n+1}, no linear form can interpolate the data (λi,λσ(i))0in(\lambda_{i},\lambda_{\sigma(i)})_{0\leq i\leq n}. ∎

The exposition regarding the cohomology of the complete intersection of Fermat type is attributed to Terasoma [15]. It is characterized by abelian covers of the projective line, as outlined in the following.

Cohomology of the complete intersection of Fermat type

Let λ0,,λn\lambda_{0},\ldots,\lambda_{n} be the n+1n+1 distinct elements in kk. Let π1(𝐏1{λ0,,λn})\pi_{1}(\mathbf{P}^{1}-\{\lambda_{0},\ldots,\lambda_{n}\}) be the fundamental group of the punctured sphere 𝐏1{λ0,,λn}\mathbf{P}^{1}-\{\lambda_{0},\ldots,\lambda_{n}\}. We denote GndG^{d}_{n} as the quotient group (𝐙/d𝐙)n+1/Δ(𝐙/d𝐙)(\mathbf{Z}/d\mathbf{Z})^{n+1}/\Delta(\mathbf{Z}/d\mathbf{Z}), where Δ:𝐙/d𝐙(𝐙/d𝐙)n+1\Delta:\mathbf{Z}/d\mathbf{Z}\to(\mathbf{Z}/d\mathbf{Z})^{n+1} represents the diagonal embedding.

Let γiπ1(𝐏1{λ0,,λn})\gamma_{i}\in\pi_{1}(\mathbf{P}^{1}-\{\lambda_{0},\ldots,\lambda_{n}\}) be the loop winding counter-clockwise around the point λi\lambda_{i}. There is a surjective map

π1(𝐏1{λ0,,λn})Gnd\pi_{1}(\mathbf{P}^{1}-\{\lambda_{0},\ldots,\lambda_{n}\})\to G^{d}_{n}

by assigning each γi\gamma_{i} to the ii-th generator (0,,1,,0)Gnd(0,\ldots,1,\ldots,0)\in G^{d}_{n}. It corresponds to a nonsingular curve DD with function field

k(D):=k(x,f1d,,fnd),fi=xλixλ0.k(D):=k(x,\sqrt[d]{f_{1}},\ldots,\sqrt[d]{f_{n}}),~{}f_{i}=\frac{x-\lambda_{i}}{x-\lambda_{0}}.

The covering map D𝐏𝟏D\to\mathbf{P^{1}} is unramified over 𝐏1{λ0,,λn}\mathbf{P}^{1}-\{\lambda_{0},\ldots,\lambda_{n}\}, and the Galois group Gal(D/𝐏1)\mathrm{Gal}(D/\mathbf{P}^{1}) is isomorphic to GndG^{d}_{n}, cf. [14, §2.1].

Let 𝔖nr\mathfrak{S}_{n-r} be the permutation group of (nr)(n-r) elements. The group 𝔖nr\mathfrak{S}_{n-r} acts on the product group (Gnd)nr(G^{d}_{n})^{n-r} by permuting its components. The product DnrD^{n-r} of the curve DD is naturally endowed with a group action by the semi-direct product (Gnd)nr𝔖nr(G^{d}_{n})^{n-r}\rtimes\mathfrak{S}_{n-r}. Define NN as the kernel of the homomorphism

Σ:(Gnd)nr𝔖nrGnd,(g1,,gnr;σ)i=1nrgσ(i).\Sigma:(G^{d}_{n})^{n-r}\rtimes\mathfrak{S}_{n-r}\to G^{d}_{n},~{}(g_{1},\ldots,g_{n-r};\sigma)\mapsto\sum_{i=1}^{n-r}g_{\sigma(i)}.

It has benn shown by Terasoma [15, Thm. 2.4.2] that the complete intersection Xn,r,dX_{n,r,d} of Fermat type is isomorphic to the quotient space Dnr/ND^{n-r}/N.

Let ζdμd\zeta_{d}\in\mu_{d} be a primitive dd-th root of unity. A character χ:Gndμd\chi:G^{d}_{n}\to\mu_{d} can be represented by a tuple (a0,,an)(𝐙/d𝐙)n+1(a_{0},\ldots,a_{n})\in(\mathbf{Z}/d\mathbf{Z})^{\oplus n+1} with ai0modd\sum a_{i}\equiv 0\ \mathrm{mod}\ d, which allows us to express χ([k0,,kn])=ζda0k0++ankn\chi([k_{0},\ldots,k_{n}])=\zeta_{d}^{a_{0}k_{0}+\cdots+a_{n}k_{n}} for any [k0,,kn]Gnd[k_{0},\ldots,k_{n}]\in G^{d}_{n}. By Kummer theory, the character χ\chi corresponds to a nonsingular curve CχC_{\chi} as the group quotient of DD by the action of Ker(χ)\operatorname{Ker}(\chi). To be more explicit, CχC_{\chi} is the cyclic cover of 𝐏1\mathbf{P}^{1} determined by the equation

(2.9) ye=(xλ0)b0(xλn)bn,bi=eaidy^{e}=(x-\lambda_{0})^{b_{0}}\cdots(x-\lambda_{n})^{b_{n}},~{}b_{i}=\frac{ea_{i}}{d}

where e=#Im(χ)e=\#\operatorname{Im}(\chi) and de=gcd(a0,,an)\frac{d}{e}=\textrm{gcd}(a_{0},\ldots,a_{n}).

Let KK be a field extension of 𝐐\mathbf{Q}_{\ell} containing dd-th roots of unity, such as K=𝐐(ζd)K=\mathbf{Q}_{\ell}(\zeta_{d}). For a GndG^{d}_{n}-representation VV over KK and a character χ\chi of GndG^{d}_{n}, the χ\chi-eigenspace is

Vχ:={vV|g(v)=χ(g)v for all gGnd}.V_{\chi}:=\{v\in V~{}|~{}g(v)=\chi(g)v\text{~{}for all~{}}g\in G^{d}_{n}\}.

The group GndG^{d}_{n} naturally acts on the cohomology space H1(D,K)\mathrm{H}^{1}(D,K). By the description of CχC_{\chi}, the χ\chi-eignespace H1(D,K)χ\mathrm{H}^{1}(D,K)_{\chi} is isomorphic to H1(Cχ,K)\mathrm{H}^{1}(C_{\chi},K). Recall that the primitive part Hprimi(Xn,r,d,K)\mathrm{H}^{i}_{\mathrm{prim}}(X_{n,r,d},K) is the orthogonal complement of the image of the restriction map Hi(𝐏n,K)Hi(Xn,r,d,K)\mathrm{H}^{i}(\mathbf{P}^{n},K)\to\mathrm{H}^{i}(X_{n,r,d},K).

Theorem 2.10 ([15, Thm. 2.5.1]).

Let notaions be as above. The primtivie cohomology of Xn,r,dX_{n,r,d} decomposes as follows

Hprimi(Xn,r,d,K){0inr;χG^ndnrH1(D,K)χi=nr.\mathrm{H}^{i}_{\mathrm{prim}}(X_{n,r,d},K)\cong\begin{cases}0&i\neq n-r;\\ \bigoplus_{\chi\in\hat{G}^{d}_{n}}\wedge^{n-r}\mathrm{H}^{1}(D,K)_{\chi}&i=n-r.\end{cases}

The cohomology of Xn,r,dX_{n,r,d} is therefore expressed in terms of the cyclic covers CχC_{\chi}. This decomposition is essential for proving the faithfulness of the cohomological action by the automorphism group of Xn,r,dX_{n,r,d}. We first examine the cases of complete intersections of quadrics.

Complete intersections of two or more quadrics

Let Xn,r,2𝐏knX_{n,r,2}\subset\mathbf{P}^{n}_{k} be the complete intersection of quadrics of Fermat type defined as in (2.4), and kk is an algebraically closed field of char(k)2\mathrm{char}(k)\neq 2. Let χG^n2\chi\in\hat{G}^{2}_{n} be the character represented by (a0,,an)(𝐙/2𝐙)n+1(a_{0},\ldots,a_{n})\in(\mathbf{Z}/2\mathbf{Z})^{\oplus n+1}, and gg be the integer 12#{ai=1}\frac{1}{2}\#\{a_{i}=1\}. The corresponding cyclic cover CχC_{\chi} is a hyperelliptic curve ramified at 2g2g points. By Hurwitz’s formula, the genus of CχC_{\chi} is equal to g1g-1. As CχC_{\chi} is complete and nonsingular, it follows that H1(Cχ,𝐐)𝐐2g2\mathrm{H}^{1}(C_{\chi},\mathbf{Q}_{\ell})\cong\mathbf{Q}_{\ell}^{2g-2} for any odd prime \ell.

  • n+1n+1 even, r=2r=2. In this case, the space n2H1(Cχ,𝐐)\wedge^{n-2}\mathrm{H}^{1}(C_{\chi},\mathbf{Q}_{\ell}) is non-zero if and only if 2g=n+12g=n+1, i.e., χ=(1,,1)\chi=(1,\ldots,1). Hence Hprimn2(Xn,r,2,𝐐)\mathrm{H}^{n-2}_{\mathrm{prim}}(X_{n,r,2},\mathbf{Q}_{\ell}) is isomorphic to the single χ\chi-eigenspace n2H1(Cχ,𝐐)\wedge^{n-2}\mathrm{H}^{1}(C_{\chi},\mathbf{Q}_{\ell}). Let eiGn2e_{i}\in G^{2}_{n} be the action

    (x0::xi::xn)(x0::xi::xn).(x_{0}:\cdots:x_{i}:\cdots:x_{n})\mapsto(x_{0}:\cdots:-x_{i}:\cdots:x_{n}).

    The eigenvalue of the induced action eie_{i}^{*} on n2H1(Cχ)\wedge^{n-2}\mathrm{H}^{1}(C_{\chi}) is χ(ei)=1\chi(e_{i})=-1. We can see that eiej=1e^{*}_{i}\circ e^{*}_{j}=1 but ejeiIde_{j}\circ e_{i}\neq\operatorname{Id}. Thus the action of Gn2G^{2}_{n} on the cohomology of XX is not faithful.

  • n+1n+1 even, r3r\geq 3. Consider the characters

    χi,j:=(1,ai,,aj,,1),ai=aj=0.\chi_{i,j}:=(1\ldots,a_{i},\ldots,a_{j},\ldots,1),~{}a_{i}=a_{j}=0.

    The genus of Cχi,jC_{\chi_{i,j}} is n32\frac{n-3}{2}, and H1(Cχi,j,𝐐)𝐐n3\mathrm{H}^{1}(C_{\chi_{i,j}},\mathbf{Q}_{\ell})\cong\mathbf{Q}_{\ell}^{n-3}. Hence nrH1(Cχi,j,𝐐)\wedge^{n-r}\mathrm{H}^{1}(C_{\chi_{i,j}},\mathbf{Q}_{\ell}) is non-zero. For an automorphism σ:=(b0,,bn)Gn2\sigma:=(b_{0},\ldots,b_{n})\in G^{2}_{n}, we have

    χi,j(σ)χi,j(σ)=(1)bjbj.\frac{\chi_{i,j}(\sigma)}{\chi_{i,j^{\prime}}(\sigma)}=(-1)^{b_{j^{\prime}}-b_{j}}.

    Assuming that σ\sigma acts trivially on the cohomology, then χi,j(σ)=1\chi_{i,j}(\sigma)=1 for all i,ji,j. It implies that

    bjbjmod2,j,j,b_{j}\equiv b_{j^{\prime}}\mathrm{~{}mod~{}}2,~{}\forall j^{\prime},j,

    and σ=(b0,,bn)\sigma=(b_{0},\ldots,b_{n}) is a diagonal element in Gn2G^{2}_{n}, which represents the identity map on XX. Thus the action of Gn2G^{2}_{n} on the primitive cohomology of Xn,r,2X_{n,r,2} is faithful.

  • n+1n+1 odd, r2r\geq 2. Let us consider the characters

    χi=(1,,ai,,1),ai=0.\chi_{i}=(1,\ldots,a_{i},\ldots,1),~{}a_{i}=0.

    The genus of CχiC_{\chi_{i}} is n22\frac{n-2}{2}, and H1(Cχi,𝐐)𝐐n2\mathrm{H}^{1}(C_{\chi_{i}},\mathbf{Q}_{\ell})\cong\mathbf{Q}_{\ell}^{n-2}. Then nrH1(Cχi,𝐐)\wedge^{n-r}\mathrm{H}^{1}(C_{\chi_{i}},\mathbf{Q}_{\ell}) is non-trivial for all 0in0\leq i\leq n. Let σ:=(b0,,bn)Gn2\sigma:=(b_{0},\ldots,b_{n})\in G^{2}_{n}. We have

    χi(σ)χj(σ)=(1)bi+bj.\frac{\chi_{i}(\sigma)}{\chi_{j}(\sigma)}=(-1)^{-b_{i}+b_{j}}.

    Assuming that σ\sigma acts trivially on the cohomology, then χi(σ)=1\chi_{i}(\sigma)=1 for all ii. It implies that

    bibjmod2,i,j.b_{i}\equiv b_{j}\mathrm{~{}mod~{}}2,~{}\forall i,j.

    By the same reason as above, Gn2G^{2}_{n} acts faithfully on the primitive cohomology of Xn,r,2X_{n,r,2}.

Proposition 2.11.

Let X:=Xn,r,d𝐏knX:=X_{n,r,d}\subset\mathbf{P}^{n}_{k} be the complete intersection of Fermat type defined in (2.4). Fix a prime char(k)\ell\neq\mathrm{char}(k). Then the group AutL(X)\operatorname{\mathrm{Aut}}_{L}(X) of linear automorphisms acts faithfully on the primitive cohomology Hprimnr(X,𝐐)\mathrm{H}^{n-r}_{\mathrm{prim}}(X,\mathbf{Q}_{\ell}) unless d=r=2d=r=2 and n+1n+1 is even.

Proof.

We begin by proving the faithfulness of the action of the subgroup GndG^{d}_{n} on Hprimnr(X,𝐐)\mathrm{H}^{n-r}_{\mathrm{prim}}(X,\mathbf{Q}_{\ell}).

Consider a field extension KK of 𝐐\mathbf{Q}_{\ell} containing the primitive dd-th root of unity, e.g., K=𝐐(ζd)K=\mathbf{Q}_{\ell}(\zeta_{d}). By Theorem 2.10, there is the eigenspace decomposition

Hprimnr(X,K)=χG^ndnrH1(Cχ,K),\mathrm{H}^{n-r}_{\mathrm{prim}}(X,K)=\bigoplus_{\chi\in\hat{G}^{d}_{n}}\wedge^{n-r}\mathrm{H}^{1}(C_{\chi},K),

where CχC_{\chi} is the cyclic over of 𝐏1\mathbf{P}^{1} defined by (2.9).

Suppose that a linear automorphism gAutL(X)g\in\operatorname{\mathrm{Aut}}_{L}(X) acts trivially on Hprimnr(X,𝐐)\mathrm{H}^{n-r}_{\mathrm{prim}}(X,\mathbf{Q}_{\ell}). By considering the base change of coefficients and the above decomposition, gg also acts trivially on each summand nrH1(Cχ,K)\wedge^{n-r}\mathrm{H}^{1}(C_{\chi},K). Since

v=g(v)=χ(g)v,vnrH1(Cχ,K),v=g^{*}(v)=\chi(g)v,~{}\forall v\in\wedge^{n-r}\mathrm{H}^{1}(C_{\chi},K),

it follows that χ(g)=1\chi(g)=1 provided that nrH1(Cχ,K)0\wedge^{n-r}\mathrm{H}^{1}(C_{\chi},K)\neq 0. To prove g=Idg=\operatorname{Id}, we aim to select a subset SS of characters of GndG^{d}_{n} satisfying

  • nrH1(Cχ,K)0,χS\wedge^{n-r}\mathrm{H}^{1}(C_{\chi},K)\neq 0,\forall\chi\in S;

  • {gGnd|χ(g)=1,χS}={Id}.\{g\in G^{d}_{n}~{}|~{}\chi(g)=1,\forall\chi\in S\}=\{\operatorname{Id}\}.

Choose integers k,s,t𝐙0k,s,t\in\mathbf{Z}_{\geq 0} satisfying

dk+n,ds+t+n1 and (n+t,d)=1.d\mid k+n,~{}d\mid s+t+n-1\text{~{}and~{}}(n+t,d)=1.

Consider the following characters

χk\displaystyle\chi_{k} :=(1,,1,k);\displaystyle:=(1,\ldots,1,k);
χ(s,t,i)\displaystyle\chi_{(s,t,i)} :=(1,,s,,t),1in,\displaystyle:=(1,\ldots,s,\ldots,t),1\leq i\leq n,

where χ(s,t,i)\chi_{(s,t,i)} means the ii-th component is ss, the (n+1)(n+1)-th component is tt, and 11 otherwise. As a result of Lemma 2.11 below, we can assert from

χk(g)=χ(s,t,i)(g)=1,1in,\chi_{k}(g)=\chi_{(s,t,i)}(g)=1,~{}1\leq i\leq n,

that gg is the identity element in GndG^{d}_{n}. The remaining is to verify

nrH1(Cχ,K)0\wedge^{n-r}\mathrm{H}^{1}(C_{\chi},K)\neq 0

for all χ=χk\chi=\chi_{k} or χ(s,t,i)\chi_{(s,t,i)}.

  • For χ=χk\chi=\chi_{k}, the cyclic covering π:Cχ𝐏1\pi:C_{\chi}\to\mathbf{P}^{1} is determined by the affine equation

    yd=i=0n1(xλi)(xλn)k.y^{d}=\prod_{i=0}^{n-1}(x-\lambda_{i})(x-\lambda_{n})^{k}.

    The branch points {λ0,,λn1}\{\lambda_{0},\ldots,\lambda_{n-1}\} have a common ramification index dd. The ramification index over the point λn\lambda_{n} is dgcd(k,d)\frac{d}{\operatorname{gcd}(k,d)}. The point at infinity 𝐏1\infty\in\mathbf{P}^{1} is unbranched because k+n0moddk+n\equiv 0\ \mathrm{mod}\ d. By Hurwitz’s formula

    2g(Cχ)2=2d+0in1(d1)+dgcd(k,d),2g(C_{\chi})-2=-2d+\sum_{0\leq i\leq n-1}(d-1)+\frac{d}{\operatorname{gcd}(k,d)},

    we obtain

    dimKH1(Cχ,K)=2g(Cχ)(d1)(n2).\dim_{K}\mathrm{H}^{1}(C_{\chi},K)=2g(C_{\chi})\geq(d-1)(n-2).

    Since n3n\geq 3 and r2r\geq 2 in our case, it follows that nrH1(Cχk,K)0\wedge^{n-r}\mathrm{H}^{1}(C_{\scalebox{0.8}{$\chi$}_{k}},K)\neq 0 for all possible kk.

  • For χ=χ(s,t,i)\chi=\chi_{(s,t,i)}, the cyclic covering Cχ𝐏1C_{\chi}\to\mathbf{P}^{1} is determined by the affine equation

    yd=0jn1,ji(xλj)(xλi)s(xλn)ty^{d}=\prod_{0\leq j\leq n-1,j\neq i}(x-\lambda_{j})(x-\lambda_{i})^{s}(x-\lambda_{n})^{t}

    Again by Hurwitz’s formula

    2g(Cχ)2=2d+0jin1(d1)+dgcd(s,d)+dgcd(t,d),2g(C_{\chi})-2=-2d+\sum_{0\leq j\neq i\leq n-1}(d-1)+\frac{d}{\operatorname{gcd}(s,d)}+\frac{d}{\operatorname{gcd}(t,d)},

    we have

    dimH1(Cχ,K)=2g(Cχ)(d1)(n3).\dim\mathrm{H}^{1}(C_{\chi},K)=2g(C_{\chi})\geq(d-1)(n-3).

    Let us discuss following cases.

    1. (1)

      n>3,d>2n>3,d>2 and r2r\geq 2. Then (d1)(n3)nr(d-1)(n-3)\geq n-r. Therefore nrH1(Cχ(s,t,i))0\wedge^{n-r}\mathrm{H}^{1}(C_{\chi_{(s,t,i)}})\neq 0 for all possible ss and tt.

    2. (2)

      n=3,d>2n=3,d>2 and r=2r=2. Note that st0s\cdot t\neq 0, otherwise the condition ()(\star) implies d2d\mid 2 as a contradiction. Hence we can assume s0s\neq 0 and dgcd(s,d)>0\frac{d}{\operatorname{gcd}(s,d)}>0. Then Hurwitz’s formula gives the inequality 2g(Cχ)(d1)(n3)+1n22g(C_{\chi})\geq(d-1)(n-3)+1\geq n-2, which implies that n2H1(Cχ(s,t,i))0\wedge^{n-2}\mathrm{H}^{1}(C_{\chi_{(s,t,i)}})\neq 0 for all possible ss and tt.

    3. (3)

      For d=2d=2, this case has been exhibited in the previous discussion for complete intersections of quadrics. The faithfulness only fails when d=r=2d=r=2 and n+1n+1 is even.

In conclusion, the action of GndG^{d}_{n} on Hprimnr(X,K)\mathrm{H}^{n-r}_{\mathrm{prim}}(X,K), as well as on Hprimnr(X,𝐐)\mathrm{H}^{n-r}_{\mathrm{prim}}(X,\mathbf{Q}_{\ell}), is faithful, except for XX being an odd-dimensional complete intersection of two quadrics. Then the assertion of this proposition holds for a generic complete intersection XX of Fermat type as Proposition 2.5 ensures AutL(X)=Gnd\operatorname{\mathrm{Aut}}_{L}(X)=G^{d}_{n}.

Now we consider the cases Gnd<AutL(X)G^{d}_{n}<\operatorname{\mathrm{Aut}}_{L}(X). For a character χG^nd\chi\in\hat{G}^{d}_{n} and an automorphism σAutL(X)\sigma\in\operatorname{\mathrm{Aut}}_{L}(X), we define the character χσ\chi^{\sigma} as

χσ(g):=χ(σgσ1).\chi^{\sigma}(g):=\chi(\sigma g\sigma^{-1}).

Note that σgσ1Gnd\sigma g\sigma^{-1}\in G^{d}_{n} since GndG^{d}_{n} is a normal subgroup. The action σ\sigma^{*} transfers a χ\chi-eigenspace to a χσ\chi^{\sigma}-eigenspace if the χ\chi-eigenspace is nontrivial. Specifically, for any vHprimnr(X,K)χv\in\mathrm{H}^{n-r}_{\mathrm{prim}}(X,K)_{\chi}

gσ(v)=σ(σgσ1)(v)=σ(χ(σgσ1)v)=χσ(g)σ(v).g^{*}\sigma^{*}(v)=\sigma^{*}(\sigma g\sigma^{-1})^{*}(v)=\sigma^{*}(\chi(\sigma g\sigma^{-1})v)=\chi^{\sigma}(g)\sigma^{*}(v).

Suppose that σ\sigma^{*} acts trivially on Hprimnr(X,K)\mathrm{H}^{n-r}_{\mathrm{prim}}(X,K). Then χ=χσ\chi=\chi^{\sigma} for any χS\chi\in S. Hence

χ(g)=χ(σgσ1),gGnd.\chi(g)=\chi(\sigma g\sigma^{-1}),~{}\forall g\in G^{d}_{n}.

From the preceding discussion, it follows that g=σgσ1g=\sigma g\sigma^{-1} for all gGndg\in G^{d}_{n}, i.e., σ\sigma is a commutator of the subgroup GndG^{d}_{n}. Let gg be a diagonal matrix with distinct entries. The relation σg=gσ\sigma g=g\sigma implies σ\sigma must be a diagonal matrix, thus σ\sigma is contained in GndG^{d}_{n}. Given that the action of GndG^{d}_{n} on the cohomology is faithful, we conclude that σ\sigma is the identity map. ∎

Lemma 2.11.

Let dd and nn be positive integers. Choose positive integers k,t,s𝐙k,t,s\in\mathbf{Z} such that

dk+n,(n+t,d)=1 and s+tk+1modd.d\mid k+n,~{}(n+t,d)=1\textrm{~{}and~{}}s+t\equiv k+1\ \textrm{mod}\ d.

Consider the (n+1)×(n+1)(n+1)\times(n+1)-matrix

(2.12) A:=(s1t1st11k)Mn+1(𝐙/d𝐙)A:=\begin{pmatrix}s&\ldots&1&t\\ \vdots&\ddots&\vdots&\vdots\\ 1&\ldots&s&t\\ 1&\ldots&1&k\end{pmatrix}\in M_{n+1}(\mathbf{Z}/d\mathbf{Z})

where the ii-th row (1,,s,,t)(1,\ldots,s,\ldots,t) has ss at place ii, tt at place n+1n+1, and 11 otherwise. Then the set of solutions {x|Ax=0}\{\vec{x}~{}|~{}A\cdot\vec{x}=0\} in (𝐙/d𝐙)n+1(\mathbf{Z}/d\mathbf{Z})^{\oplus n+1} consists of the diagonal elements {(a,,a)|a𝐙/d𝐙}\{(a,\cdots,a)~{}|~{}a\in\mathbf{Z}/d\mathbf{Z}\}.

Proof.

Suppose that x=(b1,,bn+1)\vec{x}=(b_{1},\ldots,b_{n+1}) is a solution of Ax=0A\cdot\vec{x}=0. Then we have

ji,1jnbj+sbi+tbn+10modd,j=1nbj+kbn+10modd.\sum_{j\neq i,1\leq j\leq n}b_{j}+sb_{i}+tb_{n+1}\equiv 0\ \textrm{mod}\ d,~{}\sum_{j=1}^{n}b_{j}+kb_{n+1}\equiv 0\ \textrm{mod}\ d.

It follows that

(2.13) (s1)bi+(tk)bn+10modd,1in.(s-1)b_{i}+(t-k)b_{n+1}\equiv 0\ \textrm{mod}\ d,~{}\forall 1\leq i\leq n.

Given the assumption on the integers k,t,sk,t,s, we have

(s1)(kt)modd,(tk)(n+t)modd.(s-1)\equiv(k-t)\ \textrm{mod}\ d,~{}(t-k)\equiv(n+t)\ \textrm{mod}\ d.

Since n+tn+t is coprime to dd, it follows that ktk-t and s1s-1 are invertible in 𝐙/d𝐙\mathbf{Z}/d\mathbf{Z}. Then the equation (2.13) implies

bibn+1modd,1in.b_{i}\equiv b_{n+1}\ \textrm{mod}\ d,~{}\forall{1\leq i\leq n}.

Thus our assertion follows. ∎

Now let us prove Theorem 2.1.

Proof of Theorem 2.1.

By Observation 2.2, the complete intersection X𝐏nX\subset\mathbf{P}^{n} of type (d1,,dc)(d_{1},\ldots,d_{c}), satisfying the assumptions in Theorem 2.1, can be regarded as a subscheme in the complete intersection defined by the first rr polynomials with minimal degree d:=d1d:=d_{1} where d3d\geq 3, r1r\geq 1, or d=2d=2, r3r\geq 3. Thus we shall prove that a general codimension rr complete intersection Y𝐏nY\subset\mathbf{P}^{n} of hypersurfaces with equal degree dd has no non-trivial linear automorphisms.

(i) Suppose that d3d\geq 3, r=1r=1. This reduces to the cases of hypersurfaces. A general smooth hypersurface Y𝐏nY\subset\mathbf{P}^{n} of degY3\deg Y\geq 3 has no linear automorphisms unless (d;n)=(3;2)(d;n)=(3;2), see [10, 12].

(ii) Suppose that d3d\geq 3, r2r\geq 2 or d=2d=2, r3r\geq 3. Let us verify Hypotheses (1)–(3) in Theorem 1.2.

  1. (1)

    Set π:𝒳B\pi:\mathcal{X}\to B as the universal family of smooth complete intersections of type (d1,,dc;n)(d_{1},\ldots,d_{c};n). The relative automorphism schemes AutB(𝒳)B\operatorname{\mathrm{Aut}}_{B}(\mathcal{X})\to B is a finite group scheme over BB if the moduli stack of smooth complete intersections is a separated and Delgine-Mumford stack, see [9, Lem. 7.7] or [7, Lem. 2.3]. Benoist affirmed that the moduli stack is separated and Delgine-Mumford if (d1,,dc;n)(2;n)(d_{1},\ldots,d_{c};n)\neq(2;n), see [1]. Hence Hypothsis (1) holds.

  2. (2)

    Let Xn,r,d𝐏nX_{n,r,d}\subset\mathbf{P}^{n} be the complete intersection of Fermat type defined as (2.4) with d3,r2d\geq 3,r\geq 2 or d=2,r3d=2,r\geq 3. By Proposition 2.11 AutL(Xn,r,d)\operatorname{\mathrm{Aut}}_{L}(X_{n,r,d}) acts faithfully on Hprimnr(Xn,r,d)\mathrm{H}^{n-r}_{\mathrm{prim}}(X_{n,r,d}). Thus the action is also faithful on Hnr(Xn,r,d)\mathrm{H}^{n-r}(X_{n,r,d}). Then Hypothesis (2) is satisfied.

  3. (3)

    The proof of the bigness of the monodromy group for complete intersections is in line with the proof of Theorem  1.4

    Let XbX_{b} be the smooth complete intersection of multidegree (d1,,dc)(d_{1},\ldots,d_{c}) over a general point bBb\in B. Let X1,,XcX_{1},\ldots,X_{c} be the hypersurfaces such that Xb=X1XcX_{b}=X_{1}\cap\cdots\cap X_{c}. We may assume the complete intersection Y:=X2Xc1Y\colon=X_{2}\cap\cdots\cap X_{c-1} is smooth. Then XbX_{b} is a hyperplane section in YY. Let DD be a Lefschetz pencil of hyperplane sections in YY passing through a point [Xb]D[X_{b}]\in D. The set of hyperplane sections that admit ordinary double points is a finite subset SS in DD. Let UU be the open complement D SD\mathbin{\rule[1.99997pt]{6.69998pt}{1.19995pt}}S. The monodromy action of the fundamental group π1(U,0)\pi_{1}(U,0) on Hprimnc(Xb;𝐐)\mathrm{H}^{n-c}_{\mathrm{prim}}(X_{b};\mathbf{Q}_{\ell}) factors through the monodromy action of π1(B,b)\pi_{1}(B,b) via the natural inclusion (U,0)(B,b)(U,0)\hookrightarrow(B,b). Therefore it suffices to prove the monodromy group of π1(U,0)\pi_{1}(U,0) is as big as possible.

    Let H\mathrm{H} denote the cohomology space Hprimnc(Xb;𝐐)\mathrm{\mathrm{H}}^{n-c}_{\mathrm{prim}}(X_{b};\mathbf{Q}_{\ell}), ψ\psi the intersection form on H\mathrm{H}, and MAut(H,ψ)M\subset\operatorname{\mathrm{Aut}}(\mathrm{H},\psi) the geometric monodromy group of π1(U,0)\pi_{1}(U,0). If ncn-c is odd, MM is the symplectic group Sp(H,ψ)\mathrm{Sp}(\mathrm{H},\psi) [3, Théorèm 5.10]. If ncn-c is even, then MM is either the full orthogonal group O(H,ψ)\mathrm{O}(\mathrm{H},\psi) or a finite subgroup of O(H,ψ)\mathrm{O}(\mathrm{H},\psi) [4, Théorèm 4.4.1].

    If MM is finite, the pp-adic Newton polygon for XbX_{b} is a straight line [8, Thm. 11.4.9]. Illusie [6] proved that the Newton polygon of a general complete intersection conincides with its Hodge polygon. The Hodge polygon is a straight line if and only if the Hodge numbers hi,ncih^{i,n-c-i} of XbX_{b} all vanishes except for i=nc2i=\frac{n-c}{2}. By [5, Exposé XI], such siutation arises only when (d1,,dc;n)(d_{1},\ldots,d_{c};n) falls into one of the following

    • (2;n)(2;n), hyperquadrics;

    • (3;3)(3;3), cubic surfaces;

    • (2,2;n)(2,2;n) and nn is even, even-dimensional complete intersections of two quadrics.

    None of the three cases fits the conditions d3d\geq 3, r2r\geq 2, or d=2d=2, r3r\geq 3.

Therefore, Theorem 1.2 assures that AutL(Y)\operatorname{\mathrm{Aut}}_{L}(Y) is isomorphic to {1}\{1\} or 𝐙/2𝐙\mathbf{Z}/2\mathbf{Z}. The second possibility will be ruled out by Proposition 2.14. This completes the proof. ∎

Proposition 2.14.

Let kk be an algebraically closed field of char(k)2\mathrm{char}(k)\neq 2. Let X𝐏knX\subset\mathbf{P}^{n}_{k} be a complete intersection of multidegree (d,,d)r\underbrace{(d,\ldots,d)}_{r} with r<nr<n. Assmuing d3d\geq 3, or d=2,r3d=2,r\geq 3, then a generic XX has no linear automorphisms of order 22.

Proof.

Let 𝐏kn\mathbf{P}^{n}_{k} be the projective space, associated with a kk-linear space VV of dimension n+1n+1. The defining polynomials of the codimension rr complete intersection X𝐏knX\subset\mathbf{P}^{n}_{k} span an rr-dimensional subspace WXW_{X} in the kk-space SymdV\operatorname{\mathrm{Sym}}^{d}V^{*}. If σAutL(X)\sigma\in\operatorname{\mathrm{Aut}}_{L}(X) is a linear automorphism, then WXW_{X} is σ\sigma-invariant. Our goal is to show that a generic rr-dimensional subspace in SymdV\operatorname{\mathrm{Sym}}^{d}V^{*} is not stabilized by any involution σ±Id\sigma\neq\pm\operatorname{Id}.

Let 𝐆(r,SymdV)\mathbf{G}(r,\operatorname{\mathrm{Sym}}^{d}V^{*}) be the Grassmannian of rr-dimensional subspaces in SymdV\operatorname{\mathrm{Sym}}^{d}V^{*}. For an element gG:=GL(V)g\in G\colon=\operatorname{GL}(V), we define the subset

Gr(g):={[U]𝐆(r,SymdV)|gU=U}G_{r}(g):=\{[U]\in\mathbf{G}(r,\operatorname{\mathrm{Sym}}^{d}V^{*})~{}|~{}g\cdot U=U\}

of the gg-stabilized subspaces in SymdV\operatorname{\mathrm{Sym}}^{d}V^{*}. If two elements g,hGg,h\in G are conjuagte, then Gr(g)G_{r}(g) is isomorphic to Gr(h)G_{r}(h) respectively.

Let σG:=GL(V)\sigma\in G:=\operatorname{GL}(V) be an involution. Denote by ClG(σ)\operatorname{Cl}_{G}(\sigma) the conjugacy class of σ\sigma in GG. Define the incidence variety

σ:={(g,[U])ClG(σ)×𝐆(r,SymdV)|gU=U}.\mathcal{I}_{\sigma}:=\{(g,[U])\in\operatorname{Cl}_{G}(\sigma)\times\mathbf{G}(r,\operatorname{\mathrm{Sym}}^{d}V^{*})~{}|~{}g\cdot U=U\}.

The projection σClG(σ)\mathcal{I}_{\sigma}\to\operatorname{Cl}_{G}(\sigma) is a fibration, with each fiber isomorphic to Gr(σ)G_{r}(\sigma).

Given that char(k)2\mathrm{char}(k)\neq 2, involutions in GG are similar to the diagonal matrices of order 22. Therefore, to establish our assertion, it suffices to prove

dim𝐆(r,SymdV)>dimσ=dimClG(σ)+dimGr(σ)\dim\mathbf{G}(r,\operatorname{\mathrm{Sym}}^{d}V^{*})>\dim\mathcal{I}_{\sigma}=\dim\operatorname{Cl}_{G}(\sigma)+\dim G_{r}(\sigma)

for all diagonal matrices σ±Id\sigma\neq\pm\operatorname{Id} of order 22.

The action of σ\sigma on VV induces a decomposition VV+VV\cong V_{+}\oplus V_{-} with eigenvalues ±1\pm 1. Then the symmetric tensor SymdV\operatorname{\mathrm{Sym}}^{d}V^{*}, under the action of σ\sigma, decomposes into eigenspaces

S+=j evenSdjV+SjV,S=j oddSdjV+SjV.S_{+}=\bigoplus_{j\text{~{}even}}S^{d-j}V_{+}\otimes S^{j}V_{-},~{}S_{-}=\bigoplus_{j\text{~{}odd}}S^{d-j}V_{+}\otimes S^{j}V_{-}.

Let us set

dimV+=n1,dimV=n2,n1+n2=n+1,n1,n2>0,\displaystyle\dim V_{+}=n_{1},\dim V_{-}=n_{2},~{}n_{1}+n_{2}=n+1,n_{1},n_{2}>0,
dimS+=e1,dimS=e2,e1+e2=(n+dd).\displaystyle\dim S_{+}=e_{1},\dim S_{-}=e_{2},~{}e_{1}+e_{2}=\binom{n+d}{d}.

Suppose that WSymdVW\subset\operatorname{\mathrm{Sym}}^{d}V^{*} is a subspace stabilized by σ\sigma. Then WW admits a decomposition W+WW_{+}\oplus W_{-} where W+S+,WSW_{+}\subset S_{+},W_{-}\subset S_{-}. Hence the set Gr(σG_{r}(\sigma consists of pairs of subspaces in S+S_{+} and SS_{-}. Let f1:=dimW+f_{1}:=\dim W_{+}, f2:=dimWf_{2}:=\dim W_{-}. The dimension of Gr(σ)G_{r}(\sigma) is bounded by

max{f1(e1f1)+f2(e2f2)|0fiei,f1+f2=r}.\operatorname{max}\{f_{1}(e_{1}-f_{1})+f_{2}(e_{2}-f_{2})~{}|~{}0\leq f_{i}\leq e_{i},f_{1}+f_{2}=r\}.

Let ZG(σ)<GZ_{G}(\sigma)<G denote the center of σ\sigma. As a diagonal matrix σ\sigma with sign(σ)=(n1,n2)\mathrm{sign}(\sigma)=(n_{1},n_{2}), it is direct to see dimZG(σ)=n12+n22\dim Z_{G}(\sigma)=n_{1}^{2}+n_{2}^{2}. By the formula

dimClG(σ)+dimZG(σ)=dimG,\dim\operatorname{Cl}_{G}(\sigma)+\dim Z_{G}(\sigma)=\dim G,

we have dimClG(σ)=2n1n2\dim\operatorname{Cl}_{G}(\sigma)=2n_{1}n_{2}. Then dim𝐆(r,SymdV)dimσ\dim\mathbf{G}(r,\operatorname{\mathrm{Sym}}^{d}V^{*})-\dim\mathcal{I}_{\sigma} is equal to

min{f2(e1f1)+f1(e2f2)2n1n2|0fiei,f1+f2=r}.\operatorname{min}\{f_{2}(e_{1}-f_{1})+f_{1}(e_{2}-f_{2})-2n_{1}n_{2}~{}|~{}0\leq f_{i}\leq e_{i},f_{1}+f_{2}=r\}.

Let us show that f2(e1f1)+f1(e2f2)2n1n2f_{2}(e_{1}-f_{1})+f_{1}(e_{2}-f_{2})-2n_{1}n_{2} are positive numbers within the cases d3d\geq 3 and d=2,r3d=2,r\geq 3.

  • For d3d\geq 3, we consider the two subcases:

    (a) Assuming f1f2>0f_{1}\geq f_{2}>0, we have the inequality

    f2(e1f1)+f1(e2f2)2n1n2f2(e1+e22f1)2n1n2.f_{2}(e_{1}-f_{1})+f_{1}(e_{2}-f_{2})-2n_{1}n_{2}\geq f_{2}(e_{1}+e_{2}-2f_{1})-2n_{1}n_{2}.

    Observe that n1n2(n+12)2n_{1}n_{2}\leq(\frac{n+1}{2})^{2} and f1<r<nf_{1}<r<n. Therefore,

    f2(e1+e22f1)2n1n2\displaystyle f_{2}(e_{1}+e_{2}-2f_{1})-2n_{1}n_{2} >f2(e1+e22r)(n+1)22\displaystyle>f_{2}(e_{1}+e_{2}-2r)-\frac{(n+1)^{2}}{2}
    >f2((n+dd)2n)(n+1)22.\displaystyle>f_{2}(\binom{n+d}{d}-2n)-\frac{(n+1)^{2}}{2}.

    The term (n+dd)\binom{n+d}{d} increases with dd. For d=3d=3, it is easy to verify that

    (n+33)2n(n+1)22>0,n3.\binom{n+3}{3}-2n-\frac{(n+1)^{2}}{2}>0,\forall n\geq 3.

    (b) Assuming f2=0,f1=rf_{2}=0,~{}f_{1}=r, we have

    f2(e1f1)+f1(e2f2)2n1n2=re22n1n2.f_{2}(e_{1}-f_{1})+f_{1}(e_{2}-f_{2})-2n_{1}n_{2}=re_{2}-2n_{1}n_{2}.

    The dimension e2=dimSe_{2}=\dim S_{-}, which depends on the integers n1,n2n_{1},n_{2} and dd, also increases with dd. For d=3d=3, we get e2=(n2+23)+n2(n1+12)e_{2}=\binom{n_{2}+2}{3}+n_{2}\cdot\binom{n_{1}+1}{2}. This leads to

    re22n1n2\displaystyle re_{2}-2n_{1}n_{2} =r(n2+23)+rn2(n1+12)2n1n2\displaystyle=r\cdot\binom{n_{2}+2}{3}+rn_{2}\cdot\binom{n_{1}+1}{2}-2n_{1}n_{2}
    2(n2+23)+n2n1(n1+1)2n1n2.\displaystyle\geq 2\cdot\binom{n_{2}+2}{3}+n_{2}n_{1}(n_{1}+1)-2n_{1}n_{2}.

    Hence re22n1n2>0re_{2}-2n_{1}n_{2}>0 unless n1=1,n2=0n_{1}=1,n_{2}=0. However, thess values violates the condition n1+n2=n+1>2n_{1}+n_{2}=n+1>2. Swapping f1f_{1} and f2f_{2} does not affect our argument. Thus the assertion follows.

  • For d=2,r3d=2,r\geq 3, we get

    e1=(n1+12)+(n2+12),e2=n1n2.e_{1}=\binom{n_{1}+1}{2}+\binom{n_{2}+1}{2},~{}e_{2}=n_{1}n_{2}.

    Therefore the experssion f2(e1f1)+f1(e2f2)2n1n2f_{2}(e_{1}-f_{1})+f_{1}(e_{2}-f_{2})-2n_{1}n_{2} equals

    f2(e12f1)+(f12)n1n2.f_{2}(e_{1}-2f_{1})+(f_{1}-2)n_{1}n_{2}.

    We claim that e12f1>0e_{1}-2f_{1}>0. Since n1+n22=n1rf1n_{1}+n_{2}-2=n-1\geq r\geq f_{1}, it suffices to prove that e1>2(n1+n22)e_{1}>2(n_{1}+n_{2}-2). We have

    (n1+12)+(n2+12)2(n1+n22)\displaystyle\binom{n_{1}+1}{2}+\binom{n_{2}+1}{2}-2(n_{1}+n_{2}-2)
    =\displaystyle= 12(n12+n223(n1+n2)+8)\displaystyle\frac{1}{2}(n_{1}^{2}+n_{2}^{2}-3(n_{1}+n_{2})+8)
    =\displaystyle= 12((n132)2+(n232)2+72)>0.\displaystyle\frac{1}{2}((n_{1}-\frac{3}{2})^{2}+(n_{2}-\frac{3}{2})^{2}+\frac{7}{2})>0.

    Furthermore, e1>2(n1+n22)e_{1}>2(n_{1}+n_{2}-2) ensures e1>4e_{1}>4 since n1+n2=n+1>r+14n_{1}+n_{2}=n+1>r+1\geq 4. When f12f_{1}\geq 2, it follows that f2(e12f1)+(f12)n1n2f_{2}(e_{1}-2f_{1})+(f_{1}-2)n_{1}n_{2} is positive.

    Now consider the specific cases (f1,f2)=(0,r)(f_{1},f_{2})=(0,r) and (1,r1)(1,r-1), then the expression f2(e12f1)+(f12)n1n2f_{2}(e_{1}-2f_{1})+(f_{1}-2)n_{1}n_{2} becomes re12n1n2re_{1}-2n_{1}n_{2} and (r1)(e12)n1n2(r-1)(e_{1}-2)-n_{1}n_{2} repsectively. We claim that e12>n1n2e_{1}-2>n_{1}n_{2}. In fact, the condition n1+n2>4n_{1}+n_{2}>4 implies that

    (n1+12)+(n2+12)n1n22=12((n1n2)2+n1+n24)>0.\binom{n_{1}+1}{2}+\binom{n_{2}+1}{2}-n_{1}n_{2}-2=\frac{1}{2}((n_{1}-n_{2})^{2}+n_{1}+n_{2}-4)>0.

    Using r3r\geq 3, it is easy to see that both re12n1n2re_{1}-2n_{1}n_{2} and (r1)(e12)n1n2(r-1)(e_{1}-2)-n_{1}n_{2} are positive numbers.

References

  • [1] Olivier Benoist. Séparation et propriété de Deligne-Mumford des champs de modules d’intersections complètes lisses. J. Lond. Math. Soc. (2), 87(1):138–156, 2013.
  • [2] Xi Chen, Xuanyu Pan, and Dingxin Zhang. Automorphism and cohomology ii: Complete intersections. International Journal of Mathematics, 2024. DOI:http://doi.org/10.1142/S0129167X24500289.
  • [3] Pierre Deligne. La conjecture de Weil. I. Inst. Hautes Études Sci. Publ. Math., (43):273–307, 1974.
  • [4] Pierre Deligne. La conjecture de Weil. II. Inst. Hautes Études Sci. Publ. Math., (52):137–252, 1980.
  • [5] Pierre Deligne and Nicholas Katz. Groupes de monodromie en géométrie algébrique. II. Lecture Notes in Mathematics, Vol 340. Springer-Verlag, 1973. Séminaire de Géométrie Algébrique du Bois-Marie 1967–1969 (SGA 7, II).
  • [6] Luc Illusie. Ordinarité des intersections complètes générates, pages 375–405. Birkhäuser Boston, Boston, MA, 2007.
  • [7] A. Javanpeykar and D. Loughran. Complete intersections: moduli, Torelli, and good reduction. Math. Ann., 368(3-4):1191–1225, 2017.
  • [8] Nicholas M. Katz and Peter Sarnak. Random matrices, Frobenius eigenvalues, and monodromy, volume 45 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 1999.
  • [9] Gérard Laumon and Laurent Moret-Bailly. Champs algébriques, volume 39 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 2000.
  • [10] Hideyuki Matsumura and Paul Monsky. On the automorphisms of hypersurfaces. J. Math. Kyoto Univ., 3:347–361, 1963/1964.
  • [11] T. Matsusaka and D. Mumford. Two fundamental theorems on deformations of polarized varieties. Amer. J. Math., 86:668–684, 1964.
  • [12] Bjorn Poonen. Varieties without extra automorphisms. III. Hypersurfaces. Finite Fields Appl., 11(2):230–268, 2005.
  • [13] M.A. Reid. The Complete Intersection of Two Or More Quadrics. University of Cambridge, 1972.
  • [14] Jan Christian Rohde. Cyclic coverings, Calabi-Yau manifolds and complex multiplication, volume 1975 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2009.
  • [15] Tomohide Terasoma. Complete intersections of hypersurfaces—the Fermat case and the quadric case. Japan. J. Math. (N.S.), 14(2):309–384, 1988.
  • [16] Claire Voisin. Hodge Theory and Complex Algebraic Geometry II, volume 2 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2003.