In this paper, we investigate the minimality of biharmonic hypersurfaces with some recurrent operators in a pseudo-Euclidean space.
Citation: Li Du, Xiaoqin Yuan. The minimality of biharmonic hypersurfaces in pseudo-Euclidean spaces[J]. Electronic Research Archive, 2023, 31(3): 1587-1595. doi: 10.3934/era.2023081
[1] | Tongzhu Li, Ruiyang Lin . Classification of Möbius homogeneous curves in R4. AIMS Mathematics, 2024, 9(8): 23027-23046. doi: 10.3934/math.20241119 |
[2] | Fawaz Alharbi, Yanlin Li . Vector fields on bifurcation diagrams of quasi singularities. AIMS Mathematics, 2024, 9(12): 36047-36068. doi: 10.3934/math.20241710 |
[3] | Roa Makki . Some characterizations of non-null rectifying curves in dual Lorentzian 3-space D31. AIMS Mathematics, 2021, 6(3): 2114-2131. doi: 10.3934/math.2021129 |
[4] | Young Joon Ahn . An approximation method for convolution curves of regular curves and ellipses. AIMS Mathematics, 2024, 9(12): 34606-34617. doi: 10.3934/math.20241648 |
[5] | Fumei Ye, Xiaoling Han . Global bifurcation result and nodal solutions for Kirchhoff-type equation. AIMS Mathematics, 2021, 6(8): 8331-8341. doi: 10.3934/math.2021482 |
[6] | Muhsin Incesu . LS (3)-equivalence conditions of control points and application to spatial Bézier curves and surfaces. AIMS Mathematics, 2020, 5(2): 1216-1246. doi: 10.3934/math.2020084 |
[7] | Young Joon Ahn . G2/C1 Hermite interpolation of offset curves of parametric regular curves. AIMS Mathematics, 2023, 8(12): 31008-31021. doi: 10.3934/math.20231587 |
[8] | Shudi Yang, Zheng-An Yao . Weight distributions for projective binary linear codes from Weil sums. AIMS Mathematics, 2021, 6(8): 8600-8610. doi: 10.3934/math.2021499 |
[9] | Khudija Bibi . Solutions of initial and boundary value problems using invariant curves. AIMS Mathematics, 2024, 9(8): 22057-22066. doi: 10.3934/math.20241072 |
[10] | Yan Ma, Di Han . On the high-th mean of one special character sums modulo a prime. AIMS Mathematics, 2023, 8(11): 25804-25814. doi: 10.3934/math.20231316 |
In this paper, we investigate the minimality of biharmonic hypersurfaces with some recurrent operators in a pseudo-Euclidean space.
Let X be a connected projective smooth curve over an algebraically closed field k and G⊆Aut(X) be a finite subgroup. Then G acts in a natural way on the space of the holomorphic differential forms on X, and thus we obtain a k-linear representation G→GL(H0(X,ΩX)). In other words, H0(X,ΩX) is a k[G]-module.
If we want to study the k[G]-module structure on H0(X,ΩX), a basic problem is to determine the multiplicity of each indecomposable representation on it. The group algebra k[G] is reductive (or we say semisimple) if and only if the characteristic char(k) is either zero or coprime with the order of G; in this case the indecomposable representations of G are exactly its irreducible representations.
The Chevalley-Weil formula tells us that the multiplicity of each irreducible representation can be characterized by the genus of the quotient curve X/G and the ramification data in the quotient map π:X→X/G, which is a branched (or ramified) cover.
The study of the Chevalley-Weil formula can be traced back to the late 19th century when Hurwitz [8] studied finite cyclic automorphism groups on compact Riemann surfaces (k=C). Later, in the 1930s, Chevalley and Weil [2] calculated the irreducible multiplicities for a general finite group G in the unramified case of π. Shortly after, Weil independently resolved the ramified case [18]. This result is therefore named the Chevalley-Weil formula, and the formula also holds for the characteristic p case with k[G] semisimple[9].
When char(k)=p∣#G, the structure of H0(X,ΩX) becomes more complicated, so people usually impose some restrictions on the cover π and the group G. There has been some research on the case where the ramified cover is tamely ramified [9,16] or weakly ramified [10]. In addition, there has been research on some special automorphism groups, such as cyclic groups [17], finite p-groups [5], or groups with cyclic Sylow subgroups [1]. There are also some studies over perfect fields [12].
For general theory, in 1980, Ellingsrud and Lønsted [4] used the language of equivariant K-theory to study the Lefschetz trace of coherent G-sheaves on projective varieties for k[G] semisimple. They obtained precise results of the k[G]-module structure on H0(X,L) where L is an invertible G-sheaf and X is a smooth projective curve [4, Theorem 3.8], which is a generalization of the Chevalley-Weil formula.
For higher dimensions, Nakajima gave some basic results in the 1980s. Consider a finite étale Galois cover f:X→Y between two projective algebraic varieties over a field k with G=Gal(X/Y). Then for any coherent G-sheaf F, there exists a finitely generated free k[G]-module complex
L∙:0→L0→L1→⋯→Lm→0, |
such that Hi(X,F)≃Hi(L∙) as k[G]-modules[15]. The approach basically follows Mumford's method in [14, II.5 Lemma 1].
When the étale condition on the cover f is replaced by the requirement that f be tamely ramified, the result will be weaker: the above finitely generated free k[G]-module complex is just a finitely generated projective k[G]-module complex [16]. It leads to a corollary that Hn(X,F) is k[G]-projective when Hi(X,F)=0 for all other i≠n. Now for a non-empty finite G-invariant set S in the curve X, we have H1(X,ΩX(S))=0 (and therefore Hi(X,ΩX(S))=0 for all i>0). As a direct consequence of the above result, when π is tamely ramified, the logarithmic differential space H0(X,ΩX(S)) is a projective k[G]-module. It should be noted that the field above need not be algebraically closed.
Almost at the same time, Kani described H0(X,ΩX) through the study of logarithmic differential space H0(X,ΩX(S)), also obtaining that H0(X,ΩX(S)) is projective [9]. However, at the end of his paper, Kani also pointed out that most of his results were covered by Nakajima's work. Nevertheless, Kani's proof process is more precise and also provides valuable tools for the main results of this paper.
For smooth curves, the Chevalley-Weil formula was well understood by now, but no attempt has been made for singular curves. The case of a nodal curve is a good frame in which to generalize the Chevalley-Weil formula.
Now let X be a connected projective nodal curve over an algebraically closed field k, G⊆Aut(X) is a finite subgroup of order n with k[G] semisimple, and π:X→Y=X/G is the quotient map. Then Y is also a nodal curve.
In Section 2.1, we demonstrate that the regular differentials of nodal curves are suitable generalizations of the holomorphic differentials. A regular differential φ∈H0(X,ωX) is essentially a log differential in H0(ˆX,ΩˆX(ˆSX)) that satisfies certain residue relations, where ˆSX is the preimage of the singular points under the normalization ˆX→X. Therefore G acts (right) naturally on the regular differential space H0(X,ωX). Here ωX is the dualizing sheaf of X.
In this paper, we set the goal to calculate the multiplicity of every character χ:G→k×, that is, the dimension of the eigenspace:
H0(X,ωX)χ:={φ∈H0(X,ωX)∣σφ=χ(σ)φ,∀σ∈G}. |
Remark 1.1. Every 1-dimensional representation is its own character. Although characters are the simplest irreducible representations, it still contains a lot of information that we are concerned about. There are mainly two reasons why we study it:
● When G is cyclic (or abelian), all irreducible representations of G are 1-dimensional;
● The space of G-invariant regular differentials is just the eigenspace of trivial character 1G:G→k×, namely H0(X,ωX)G=H0(X,ωX)1G.
Recall that when X is smooth, every G-invariant holomorphic differential on X is the pullback of a holomorphic differential on the quotient curve Y=X/G (see Proposition 3.2). Consequently,
dimkH0(X,ΩX)G=gY, |
where gY denotes the (geometric) genus of Y.
When X is a nodal curve, we establish that
H0(X,ωX)G=π∗H0(Y,ωY), |
proven for the irreducible case in Proposition 3.5 and extended to the general case (where X has multiple irreducible components) in Proposition 4.3. This leads to
dimkH0(X,ωX)G=pa(Y), |
where pa(Y) is the arithmetic genus of Y. This result generalizes the smooth case mentioned earlier.
For an arbitrary character χ:G→k×, when X is smooth, we have:
dimkH0(X,ΩX)χ=gY−1+mχ+⟨χ,1G⟩. |
In this expression:
● gY is the genus of Y.
● The term mχ is defined as
mχ=∑Q∈Y⌊eQ−1eQ+1n⟨χ,RG,Q⟩G⌋−1n⟨χ,RG⟩G. |
Here:
– The sum runs over all points Q∈Y.
– eQ:=eP is the ramification index at any P∈π−1(Q). Note that eQ>1 if and only if Q is a branch point, with RG,Q=0 at all other points; therefore the sum is finite.
● RG denotes the ramification module of π and RG,Q is the ramification module at the point Q (see Section 2.4).
● The notation ⟨χ,V⟩G:=dimkHomk[G](W,V) represents the multiplicity of χ in a G-representation V, where W is the irreducible k[G]-module corresponding to χ.
● ⌊a⌋ denotes the greatest integer less than or equal to a.
When X is a nodal curve, we require the quotient curve Y=X/G to be smooth* (see Remark 3.9).
*Note that the smoothness assumption is imposed for a general character χ, while in the case of the trivial character χ=1G, the quotient Y=X/G is permitted to be nodal.
For the irreducible case, we have
H0(X,ωX)χ∼→H0(ˆX,ΩˆX(ˆSχX))χ, |
where ˆX→X is the normalization. We present the Chevalley-Weil formula for irreducible nodal curves in Theorem 3.13:
dimkH0(X,ωX)χ=gY−1+mχ(ˆSχX)+⟨χ,1G⟩. |
In this expression:
● ˆSχX⊂ˆSX is the singular χ-set of π (see Proposition 3.7).
● The term mχ(ˆSχX) is an integer defined as
mχ(ˆSχX)=#ˆπ(ˆSχX)+∑Q∉ˆπ(ˆSχX)⌊eQ−1eQ+1n⟨χ,RG,Q⟩G⌋−1n⟨χ,RG⟩G. |
Here:
– ˆπ:ˆX→Y is the projection map after normalization.
– #ˆπ(ˆSχX) denotes the number of points in the image of ˆSχX under ˆπ.
– eQ and RG,RG,Q are defined as before, but for ˆπ this time.
Remark 1.2. Compared to the smooth case, the difference in dimkH0(X,ωX)χ arises between mχ and mχ(ˆSχX). The latter incorporates additional ramification data at singularities (the singular χ-set). Note that when ˆSχX=∅, we have
H0(X,ωX)χ∼→H0(ˆX,ΩˆX)χ, |
and mχ(ˆSχX)=mχ, just as in the smooth case.
In particular, when X is smooth, we have ˆX=X, so this formula is a direct generalization of the smooth one.
In the general case, let X=∪di=1Xi be the decomposition into irreducible components. We can reduce the computation to the irreducible case.
Let α1:ˆX1→X1 be the normalization of X1, and then ˆX1ˆπ1→Y∼=X1/G1 is the induced covering map of smooth curves and we have
H0(X,ωX)χ∼→H0(X1,ωX1(Iχ1))χ1∼→H0(ˆX1,ΩˆX1[ˆSχ1X1⊔Iχ1])χ1. |
Here:
● G1:={σ∈G∣σ(X1)=X1} is the stabilizer subgroup of G on the component X1.
● χ1:G1→k× is the restriction of χ to G1.
● ˆSχ1X1 is the singular χ1-set of ˆπ1.
● I1 denotes the intersection locus in X (comprising all intersection points of the irreducible components) restricted to X1 and
Iχ1={P∈I1∣∃τ∈GP−G1 such that χ(τ)=−1}. |
Here GP is the stabilizer subgroup.
● ˆSχ1X1∩Iχ1=∅, since the points in I1 are all nodes in X, but smooth points in X1.
By applying the same argument as in the irreducible case, we obtain the Chevalley-Weil formula on connected nodal curves in Theorem 4.5:
dimkH0(X,ωX)χ=gY−1+mχ1(ˆSχ1X1⊔Iχ1)+δχ. |
● See the term mχ1(ˆSχ1X1⊔Iχ1) in (4.14).
● The final term δχ=0 or 1 where δχ=1 if and only if Iχ1=∅ and χ1=1G1.
Remark 1.3. Compared to the H0(X1,ωX1)χ1, the difference lies between mχ1(ˆSχ1X1) and mχ1(ˆSχ1X1⊔Iχ1). The latter includes additional ramification data at intersection points. Note that when d=1, we have X=X1 and Iχ1=∅. Therefore,
H0(X,ωX)χ∼→H0(X1,ωX1)χ1. |
Hence, this formula directly generalizes the irreducible case mentioned above.
In this paper, we consider a finite group G acting faithfully on a connected nodal curve X over an algebraically closed field k. Let #G=n and suppose that either char(k)=p∤n or char(k)=0 so that k[G] is semisimple. A curve means an equidimensional reduced projective scheme of finite type of dimension 1 over k.
Let X be a connected nodal curve. Since X is a projective variety, the dualizing sheaf ωX always exists and can be explicitly described. Generally speaking, if X⊆PNk with codimension r, then the dualizing sheaf of X is given by ωX=ExtrPNk(OX,ωPNk) [7, Ⅲ.7.5]. Once the dualizing sheaf exists, it is unique together with its trace map [7, Ⅲ.7.2].
It can be seen that the construction above involves a choice of ambient space PNk. However, for the nodal curve, there is a simpler description based on its normalization.
Let α:ˆX→X be the normalization of reduced curves [11, 7.5.1], namely we have ˆX=∐1≤i≤dˆXi, where each ˆXi is the normalization of the irreducible component Xi in X. Each ˆXi is a smooth curve, so ωˆX=ΩˆX is the canonical sheaf of X. Let ˆSX⊂ˆX be the preimage of the nodes in X, and then we define the sheaf of regular differentials on an open subset V⊆X by
ωregX(V):={η∈Γ(α−1(V),ΩˆX(ˆSX))∣∑ˆP∈α−1(P)ResˆP(η)=0,∀P∈V}. | (2.1) |
We can see that ωregX is an OX-module, and we call its global sections the regular differentials on X. We say a meromorphic differential φ0 in ˆX with at worst simple poles is regular at a point P∈X if ∑ˆP∈α−1(P)ResˆP(φ0)=0, namely the residue relation holds at P.
Remark 2.1. The definition and properties of residues for differentials on curves can be found in [7, Ⅲ.7.14]. Note that when X is a smooth curve, the regular differentials on X are exactly the holomorphic differentials. Therefore, regular differentials are the natural generalization of holomorphic differentials.
In fact, the following fact holds for dualizing sheaves on nodal curves:
Theorem 2.2. Let X be a nodal curve, and then we have the canonical isomorphism ωregX≃ωX.
This is a highly nontrivial result. In the exercises of [6, 3.A], one can verify the universal property of the dualizing sheaf by endowing ωregX with a trace map. For further verification on the compatibility, the readers are referred to the detailed proof in [3, 5.2].
In summary, the dualizing sheaf on a nodal curve has a precise characterization that depends only on the structure of meromorphic differentials with at worst simple poles in ΩˆX(ˆSX).
It is known that H0(X,ωX) is a k-vector space of dimension pa(X), the arithmetic genus of X. Since the finite group G acts on X, both the rational function field K(X) and H0(X,ωX) are naturally (right) k[G]-modules. Every 1-dimensional representation is its own character. Our goal is to compute the multiplicity of every character χ:G→k×, that is, the dimension of the eigenspace H0(X,ωX)χ:={φ∈H0(X,ωX)∣σφ=χ(σ)φ,∀σ∈G} over k.
For the sake of discussion, let X be smooth for the rest of this section. Let G⊆Aut(X) be a finite group of automorphisms of order n. Then the quotient map π:X→Y:=X/G is a Galois covering, i.e., K(X)/K(Y) is a Galois field extension, where K(X) and K(Y) are the rational function fields of the corresponding curves.
Proposition 2.3. Let χ:G→k× be a character, and then there exists a rational function fχ∈K(X)× such that σfχ=χ(σ)fχ,∀σ∈G.
This result is a special case of the following theorem (in [13, 5.23]):
Theorem 2.4. Let E/F be a Galois field extension, and let G=Gal(E/F). Then H1(G,E×)=0.
Remark 2.5. (NOTES below Corollary 5.25 in [13]). This theorem is a generalization of the famous Hilbert's Theorem 90, which was first discovered by Kummer in the case of Q[ξp]/Q, and later generalized by Emmy Noether. This theorem and its various generalizations are all referred to as Hilbert's Theorem 90.
Here we only prove Proposition 2.3. Before that, we first prove a lemma:
Lemma 2.6 (Dedekind's independence theorem). Let F be a field, and G be a group. Then any finite number of different group homomorphisms χ1,…,χm:G→F× are linearly independent over F, i.e.,
∑aiχi=0⟹a1=0,…,am=0. |
Proof. We use induction on m. The statement is obvious when m=1. Assume it holds for m−1. Suppose there exist ai∈F such that
a1χ1(x)+a2χ2(x)+⋯+amχm(x)=0,∀x∈G. |
Next, we prove ai=0. Without loss of generality, suppose for some g∈G, χ1(g)≠χ2(g), and then we have
a1χ1(g)χ1(x)+a2χ2(g)χ2(x)+⋯+amχm(g)χm(x)=0,∀x∈G. |
Subtracting the first equation multiplied by χ1(g) from this equation, we get
a′2χ2+⋯+a′mχm=0,a′i=ai(χi(g)−χ1(g)). |
By induction, a′i=0,i=2,⋯,m. Since χ2(g)−χ1(g)≠0, a2=0, thus we have
a1χ1+a3χ3+⋯+amχm=0. |
By the induction hypothesis, the rest of the ai=0.
Proof of Proposition 2.3. Consider the mapping
∑τ∈Gχ(τ)τ:K(X)→K(X). |
By the above lemma, this mapping is not zero, so there exists a rational function g∈K(X) such that
f:=∑τ∈Gχ(τ)τg≠0. |
Note that, for ∀σ∈G, we have
σf=∑τ∈Gχ(τ)⋅στ(g)=∑τ∈Gχ(σ)−1χ(στ)⋅στ(g)=χ(σ)−1f. |
Thus fχ:=f−1 is the desired function.
The quotient map π:X→Y is also a ramified cover. Let eP be the ramification index at P∈X, and then we have the ramification divisor
Rπ=∑P∈X(eP−1)P. |
For a divisor D=∑aiPi∈Div(X), define π∗D∈Div(Y) by
π∗D=∑aiπ(Pi). |
If D=∑aiQi∈Div(Y) is a divisor and r∈R, then define ⌊rD⌋∈Div(Y) by
⌊rD⌋=∑⌊rai⌋Qi, |
where ⌊rai⌋ denotes the greatest integer ≤rai. Define π∗D∈Div(Y) by
π∗D:=∑i∑P∈π−1(Qi)(aieP)P∈Div(Y). |
Proposition 2.7 (Kani [9]). Let G be a finite group (of order n) acting on a smooth curve X with Rπ the ramification divisor of π:X→Y=X/G. Suppose D∈Div(X) is a G-invariant divisor, and then for the trivial character χ=1G, we have
H0(X,OX(D))G=π∗H0(Y,OY⌊n−1π∗D⌋), | (2.2) |
H0(X,ΩX(D))G=π∗H0(Y,ΩY⌊n−1π∗(D+Rπ)⌋). | (2.3) |
For any character χ, let fχ∈K(X)∗ be such that σfχ=χ(σ)fχ for all σ∈G (whose existence is guaranteed by Hilbert's Theorem 90). Then
H0(X,OX(D))χ=fχ⋅π∗H0(Y,OY⌊n−1π∗(D+(fχ))⌋), | (2.4) |
H0(X,ΩX(D))χ=fχ⋅π∗H0(Y,ΩY⌊n−1π∗(D+(fχ)+Rπ)⌋). | (2.5) |
Proof. (More details are given here than in [9].) First, it is easily deduced from the definition that π∗π∗D=nD since D is G-invariant.
For (2.2), we have D≥π∗⌊n−1π∗D⌋ and hence H0(X,OX(D))G⊇π∗H0(Y,OY⌊n−1π∗D⌋). Conversely, if f∈H0(X,OX(D))G, then f=π∗e with some e∈K(Y). Hence
π∗((f)+D)=n(e)+π∗D≥0, |
which implies (e)+⌊n−1π∗D⌋≥0.
To prove (2.3), fix a meromorphic differential 0≠φ∈Ω(Y), which exists by Riemann-Roch provided that the pole multiplicities are sufficiently high. By
H0(X,ΩX(D))=H0(X,OX(D+(π∗φ)))⋅π∗φ=H0(X,OX(D+π∗(φ)+Rπ))⋅π∗φ, |
we have
H0(X,ΩX(D))G=H0(X,OX(D+π∗(φ)+Rπ))G⋅π∗φ=π∗H0(Y,OY⌊n−1π∗((D+π∗(φ)+Rπ))⌋)⋅π∗φ=π∗(H0(Y,OY(⌊n−1π∗(D+Rπ)⌋)+(φ))⋅φ)=π∗H0(Y,ΩY⌊n−1π∗(D+Rπ)⌋). |
Finally, (2.4) and (2.5) for general χ follow from
H0(X,OX(D))χ=fχ⋅H0(X,OX(D+(fχ)))G, |
H0(X,ΩX(D))χ=fχ⋅H0(X,ΩX(D+(fχ)))G. |
Note that the divisor (fχ) is G-invariant, so it suffices to repeat the discussion above with D replaced by D+(fχ).
The constructions in this section are based on Kani's work[9]. Fix a point P∈X, and let GP:={σ∈G∣σ(P)=P} be the stabilizer subgroup of G at P, which is a cyclic group of order eP. Then there is a unique character θP:GP→k× such that for any f∈K(X)×,
σff≡θP(σ)vP(f)(modmP),∀σ∈GP, |
where vP denotes the valuation at P and mP the maximal ideal of the local ring OP.
Set
RG,P:=IndGGP(eP−1⨁d=0d⋅θdP).∗ |
*Here d⋅θdP means ⊕dθdP.
Definition 2.8. Let Bl(Y) be the branch locus of π:X→Y, namely the subset of all branch points in Y. For a point Q∈Y, we define the ramification module of Q by
RG,Q:=⨁P∈π−1(Q)RG,P, |
and the ramification module of π by
RG:=⨁Q∈YRG,Qi. |
Note that this is a finite sum because RG,Q=0 for Q∉Bl(Y).
Let χ:G→k× be a character and fχ∈K(X)χ as in Proposition 2.3. Since χn=1G, we have fnχ∈k(X)G=π∗k(Y). Write (fnχ)=π∗(nA+B) where A,B∈Div(Y) and ⌊n−1B⌋=0. Note that Supp(B)⊆Bl(Y), so we write B=∑Q∈Bl(Y)bQQ. By definition, we have
bQ=n⟨vQ(fnχ)n⟩,† |
†In the expression vQ(fnχ), the function fnχ is regarded as a rational function on Y. Consequently, we have vP(fnχ)=eQvQ(fnχ).
where ⟨r⟩=r−⌊r⌋ denotes the fractional part of r.
The following lemma shows that this B is independent of the choices of fχ:
Lemma 2.9. Let χ:G→k× be a character. Then for any Q∈Bl(Y), we have
n⟨vQ(fnχ)n⟩=⟨χ,RG,Q⟩G. | (2.6) |
Proof. Let P∈π−1(Q). Then by Frobenius reciprocity, we have
⟨χ,RG,P⟩G=⟨χ|GP,eP−1⨁d=0d⋅θdP⟩GP. | (2.7) |
Note that θdP runs through are all the irreducible representations of GP, and hence we have
⟨χ,RG,P⟩G=a⇔χ|GP=θaP, | (2.8) |
for 0≤a<eP. Choose a generator σ of GP, and then by the definition of fχ, we have
σfχ=χ(σ)fχ=θP(σ)afχ. |
Furthermore, by the definition of θP, we have
θP(σ)a=σfχfχ≡θP(σ)vP(fχ)(modmP), |
which implies a≡vP(fχ)(modeP) since θP(σ) has order eP in k×. Finally,
⟨χ,RG,P⟩GeP=⟨vP(fχ)eP⟩=⟨vQ(fnχ)n⟩=bQn, |
and hence we have
⟨χ,RG,Q⟩G=⟨χ,⨁P∈π−1(Q)RG,P⟩G=neP⟨χ,RG,P⟩G=bQ. |
Here is a basic result on the nodal curve quotient [11, 10.3.48]:
Proposition 3.1. Let X be a reduced nodal curve, and G⊆Aut(X) be a finite automorphism group on X. Then the quotient curve Y=X/G is a nodal curve. More precisely, let P∈X be a closed point, and Q be its image in Y. Then we have the following results:
(a) If X is smooth at P, then Y is smooth at Q;
(b) If P is an ordinary double point on X, then Q is either a smooth point or an ordinary double point.
Let X be an irreducible nodal curve in this section.
Let G be a finite group acting on an irreducible nodal curve X and Y=X/G the quotient curve, which is also an irreducible nodal curve. We want to study the G-invariant space of regular differentials H0(X,ωX)G. It is a classical fact that:
Proposition 3.2. If X is smooth, then
dimkH0(X,ΩX)G=gY. |
Proof. With the notations of Section 2.3, let eQ:=eP for any P∈π−1(Q). Note that
⌊n−1π∗Rπ⌋=∑Q∈Y⌊eQ−1eQ⌋Q=0. |
By Proposition 2.7 (2.3), we have
H0(X,ΩX)G=π∗H0(Y,ΩY⌊n−1π∗Rπ⌋)=π∗H0(Y,ΩY), |
which is of dimension gY, the geometric genus of Y.
Here comes a natural question: For the covering π:X→Y of irreducible nodal curves, do we still have the equality
dimkH0(X,ωX)G=pa(Y)? | (3.1) |
The answer is yes.
Consider the normalizations ˆX→X and ˆY→Y, respectively. We have a canonical isomorphism ˆX/G=ˆY, so there is a commutative diagram:
![]() |
This induces the corresponding morphisms of differentials
![]() |
(3.2) |
The lower row is obtained by ˆπ−1(ˆSY)⊆ˆSX, since X→Y maps smooth points to smooth points.
Lemma 3.3. There is a canonical inclusion
π∗H0(Y,ωY)⊆H0(X,ωX), |
that makes (3.2) commute.
Proof. We say {P1,P2}⊆ˆSX is a pair if the two points are precisely the preimages of the same node in X. In this context, a differential φ∈H0(ˆX,ΩˆX(ˆSX)) is regular if and only if
ResP1φ=−ResP2φ, |
for every pair {P1,P2}.
In this situation, we say that {P1,P2} is the preimage of the corresponding node, denoted by P. Note that for any pair {P1,P2}, the ramification indices are equal, i.e.,
eP1=eP2, |
since the orbits of both points have the same cardinality, namely, #{σ(P)∈X∣σ∈G}.
Now, consider a regular differential φY∈H0(Y,ωY). For any pair {P1,P2}=α−1(P)⊆ˆSX, we have
Resˆπ(P1)φY=−Resˆπ(P2)φY. |
Moreover, for any point P0∈ˆX, the pullback satisfies
ResP0(ˆπ∗φY)=eP0⋅Resˆπ(P0)φY. |
Thus, applying these equalities for the pair {P1,P2}, we obtain
ResP1(ˆπ∗φY)=eP1⋅Resˆπ(P1)φY=−eP2⋅Resˆπ(P2)φY=−ResP2(ˆπ∗φY). | (3.3) |
This shows that ˆπ∗φY is regular at P, and hence is a regular differential in X.
Furthermore, we have
Lemma 3.4. For the left column of (3.2), we have
H0(X,ωX)G↪H0(ˆX,ΩˆX(π−1(ˆSY)))G. |
Proof. All we need is to show a G-invariant regular differential has no poles at ˆSX−ˆπ−1(ˆSY). Suppose ˆSX−ˆπ−1(ˆSY)≠∅, otherwise there is nothing to prove.
Given a pair {P1,P2}⊆ˆSX−ˆπ−1(ˆSY), we know that the points of ˆSX−ˆπ−1(ˆSY) are mapped to the smooth locus of Y, and hence there exists a σ∈G such that σ(P1)=P2. So for any regular differential φ∈H0(X,ωX)G, we have ResP1φ=ResP1σφ=Resσ(P1)φ=ResP2φ. As ResP1φ=−ResP2φ by definition, it forces that ResP1φ=ResP2φ=0, which implies φ has no poles at {P1,P2}.
Now we can give a positive answer to (3.1).
Theorem 3.5. With the notations above, the lower row of the canonical commutative diagram
![]() |
is an isomorphism. Moreover, both H0(X,ωX)G and H0(Y,ωY) are the subspaces of log differentials satisfying the residue relations, so we have the isomorphism for the upper row. In particular, we have dimkH0(X,ωX)G=pa(Y).
Back to the smooth case, we have the following:
Proposition 3.6. Suppose π:X→Y is the quotient morphism of smooth curves and S⊂Y is a finite set. Then
H0(Y,ΩY(S))→H0(X,ΩX(π−1(S))G | (3.4) |
is an isomorphism.
Proof. By Proposition 2.7 (2.3), we have
H0(X,ΩX(π−1(S))G=π∗H0(Y,ΩY(⌊n−1π∗(π−1(S)+Rπ)⌋)).∗ |
*Here and there, we adopt the convention that every finite subset is interpreted as an effective divisor when needed.
First, note that for the divisor n−1π∗π−1(S), the coefficient corresponding to a point Q∈S is exactly 1/eQ. Now, consider
⌊n−1π∗π−1(S)+n−1π∗Rπ⌋=∑aQQ, |
and analyze the coefficient aQ for each prime divisor Q by considering three cases:
● If Q∈Bl(Y)−S (i.e., Q belongs to the branch locus but is not in S), then
aQ=⌊eQ−1eQ⌋=0. |
● If Q∈S−Bl(Y) (i.e., Q is in S but not in the branch locus), then
aQ=1. |
● If Q∈S∩Bl(Y) (i.e., Q is both in S and in the branch locus), then
aQ=⌊1eQ+eQ−1eQ⌋=⌊1⌋=1. |
Since these are the only possibilities, it follows that ⌊n−1π∗(π−1(S)+Rπ)⌋=S.
Proof of Theorem 3.5. Apply Proposition 3.6 to ˆπ:ˆX→ˆY and ˆSY⊂ˆY.
Let χ be a character of G, and α:ˆX→X the normalization. With the notations in (3.2), consider the embedding
H0(X,ωX)χ↪H0(ˆX,ΩˆX(ˆSX))χ. |
We want to determine the image of H0(X,ωX)χ in H0(ˆX,ΩˆX(ˆSX))χ.
Proposition 3.7. Suppose that Y is smooth. Define
ˆSχX:={ˆP∈ˆSX∣∃τ∈Gα(ˆP)s.t.τ(ˆP)≠ˆPandχ(τ)=−1}. | (3.5) |
Then the image of H0(X,ωX)χ in H0(ˆX,ΩˆX(ˆSX))χ is equal to H0(ˆX,ΩˆX(ˆSχX))χ. So we have an isomorphism
H0(X,ωX)χ∼→H0(ˆX,ΩˆX(ˆSχX))χ. | (3.6) |
We call ˆSχX the singular χ-set of π.
Proof. Let φ0∈H0(X,ωX)χ a regular differential. If it has poles on a pair {P1,P2}⊆ˆSX, then there is some T∈GP such that T(P1)=P2†, which is guaranteed by the smoothness of Y. So we have
χ(T)ResP1(φ0)=ResP1(Tφ0)=ResP2(φ0)=−ResP1(φ0), |
†Hence T(P2)=P1 by symmetry.
which implies χ(T)=−1 and P1,P2∈ˆSχX. In particular, when ˆSχX=∅, we see that any regular differential in H0(X,ωX)χ has no pole on ˆX.
Conversely, assume φ∈H0(ˆX,ΩˆX(ˆSχX))χ and ˆSχX≠∅. Let {P1,P2}⊆ˆSχX be a pair. By definition, there is some automorphism T∈GP such that T(P1)=P2 and χ(T)=−1. Hence
−ResP1φ=ResP1Tφ=ResT−1(P1)φ=ResP2φ. |
This means φ is regular.
Remark 3.8. If χ=1G is the trivial representation, then χ(σ)≠−1 for any σ∈G, and hence ˆS1GX=∅. By Propositions 3.7 and 3.2, we obtain
H0(X,ωX)G∼→H0(ˆX,ΩˆX)G=π∗H0(Y,ΩY). |
This conclusion is consistent with Proposition 3.5 in the case that Y is smooth, in which case ˆSY=∅.
Remark 3.9. Here we explain why the condition of Y being smooth is needed. Suppose the meromorphic differential φ0∈H0(ˆX,ΩˆX(S)) has a simple pole at some point P1∈S, i.e., ResP1φ0≠0. If φ0∈H0(X,ωX)χ, then φ0 is regular at α(P1) (α is the normalization), so that there exists a pair {P1,P2} on ˆX satisfying the following two conditions:
(a) vP2(φ0)=vP1(φ0)=−1;
(b) ResP2φ0=−ResP1φ0.
For condition (a), generally, we cannot determine the value vP2(φ0) from vP1(φ0). However, when Y is smooth, we have the following commutative diagram:
![]() |
Here ˆπ is a ramified cover of smooth curves. Since P1,P2 are mapped to the same point in Y (through X), there exists an automorphism τ that permutes the pair {P1,P2} and satisfies τφ0=χ(τ)φ0. Consequently, we must have
vP2(φ0)=vP1(φ0) |
and
ResP2φ0=χ(τ)ResP1φ0. |
For condition (b), under the hypothesis Y being smooth, it is equivalent to say χ(τ)=−1 for some (and hence any) automorphism τ permuting {P1,P2}. Furthermore, this criterion works for the intersection points in a nodal curve with several irreducible components (see Section 4.2).
Assume Y=X/G is smooth for the rest of this section, and then ˆπ:ˆX→Y is the induced ramified cover of π:X→Y. Now we compute the dimension of H0(ˆX,ΩˆX(ˆSχX))χ through Proposition 2.7.
Let fχ be a rational function on ˆX such that σfχ=χ(σ)fχ,∀σ∈G. Set
Dχ:=⌊n−1ˆπ∗(ˆSχX+(fχ)+Rˆπ)⌋. |
By Proposition 2.7 (2.5), we have
H0(ˆX,ΩˆX(ˆSχX))χ=fχ⋅ˆπ∗H0(Y,ΩY(Dχ)). |
At this point, we reduce the problem to calculating the dimension of H0(Y,ΩY(Dχ)).
Using the Riemann-Roch theorem on Y, we have
dimkH0(Y,ΩY(Dχ))=dimkH0(Y,OY(−Dχ))+degDχ+gY−1. | (3.7) |
For dimkH0(Y,OY(−Dχ)), we have:
Lemma 3.10. The space H0(Y,OY(−Dχ)) vanishes except when χ=1G, and in this case, we have dimkH0(Y,−D1G)=1.
Proof. We will prove the following:
1) If ˆSχX≠∅, then degDχ>0.
2) The divisor Dχ is principal if and only if χ=1G; in this case, ˆSχX=∅.
Recall that we have
ˆπ∗(fχ)=∑Q∈YneQaQ⋅Q, |
where aQ=vP(fχ), ∀P∈ˆπ−1(Q). Note that
⌊n−1ˆπ∗((fχ)+Rˆπ)⌋=∑Q⌊aQ+eQ−1eQ⌋Q≥∑QaQeQQ=n−1ˆπ∗(fχ). |
Therefore, we have the inequalities
deg⌊n−1ˆπ∗(ˆSχX+(fχ)+Rˆπ)⌋≥deg⌊n−1ˆπ∗((fχ)+Rπ)⌋≥degn−1ˆπ∗(fχ)=0. | (3.8) |
Now, write
ˆπ∗(ˆSχX)=∑Q∈YneQcQ⋅Q. |
If ˆSχX≠∅, then there exists some point Q′ such that cQ′≥1. Hence, we can write
degDχ=∑Q≠Q′⌊cQ+aQ+eQ−1eQ⌋+⌊cQ′+aQ′+eQ′−1eQ′⌋≥∑Q≠Q′⌊cQ+aQ+eQ−1eQ⌋+⌊aQ′+eQ′eQ′⌋>∑QaQeQ=0. | (3.9) |
Thus we have degDχ>0 provided ˆSχX≠∅, and consequently
dimkH0(Y,OY(−Dχ))=0. |
Next, assume that χ=1G. It follows that ˆSχX=∅ and fχ=ˆπ∗h for some rational function h∈K(Y). Then we have
Dχ=⌊n−1ˆπ∗(ˆπ∗h)+n−1ˆπ∗Rˆπ⌋=n−1ˆπ∗(ˆπ∗h)+⌊n−1ˆπ∗Rˆπ⌋=n−1ˆπ∗(ˆπ∗h)+0=(h), | (3.10) |
which shows that Dχ is principal. Therefore, in this case, dimkH0(Y,OY(−Dχ))=1.
Conversely, suppose that Dχ is a principal divisor; that is, there exists some rational function h∈K(Y) such that Dχ=(h). Then degDχ=0, which forces ˆSχX=∅. Then, by the previous calculations,
0=deg⌊n−1ˆπ∗((fχ)+Rˆπ)⌋=∑Q⌊aQ+eQ−1eQ⌋≥∑QaQeQ=degn−1ˆπ∗(fχ)=0. |
This chain of equalities implies that
⌊aQ+eQ−1eQ⌋=aQeQ, |
which means that eQ∣aQ. Therefore,
Dχ=n−1ˆπ∗(fχ)=(h). |
Applying ˆπ∗ to both sides, we obtain
n−1ˆπ∗ˆπ∗(fχ)=ˆπ∗(h), |
which implies that (fχ)=(ˆπ∗h). In other words, fχ∈K(X)G, i.e., χ=1G.
Now we want to calculate degDχ.
Definition 3.11. Let S⊆ˆX be a finite subset stable by G, and define
mχ(S):=#ˆπ(S)+∑Q∉ˆπ(S)⌊eQ−1eQ+1n⟨χ,RG,Q⟩G⌋−1n⟨χ,RG⟩G, | (3.11) |
where RG,Q and RG are ramification modules of ˆπ:ˆX→Y.
Lemma 3.12. We have degDχ=mχ(ˆSχX), which is independent of the choice of fχ.
Proof. Let (fnχ)=ˆπ∗(nA+B) as in the discussion before Lemma 2.9 satisfying ⌊n−1B⌋=0. Then we have
Dχ=⌊n−1ˆπ∗ˆSχX+n−1ˆπ∗(fχ)+n−1ˆπ∗Rˆπ⌋=⌊n−1ˆπ∗ˆSχX+n−2ˆπ∗ˆπ∗(nA+B)+n−1ˆπ∗Rˆπ⌋. |
Note that
n−1ˆπ∗ˆSχX=∑Q∈ˆπ(ˆSχX)1eQQ , n−1ˆπ∗Rˆπ=∑Q∈Bl(Y)eQ−1eQQ. |
So we decompose n−1ˆπ∗(ˆSχX)=U+V into two parts according to whether the point is branched, so that V is a Q-divisor which is supported on the branch locus of ˆπ, i.e., Supp(V)=Bl(Y)∩ˆπ(ˆSχX). So U is an integer divisor with all coefficients equal to 1. Now we have
Dχ=U+A+⌊V+n−1B+n−1ˆπ∗Rˆπ⌋. | (3.12) |
Let B=∑QbQQ (supported on Bl(Y)) and according to degB=−degnA, we get
degDχ=degU+∑Q∈Supp(V)⌊1+bQn⌋+∑Q∉ˆπ(ˆSχX)⌊eQ−1eQ+bQn⌋−∑QbQn=#ˆπ(ˆSχX)+∑Q∉ˆπ(ˆSχX)⌊eQ−1eQ+bQn⌋−∑QbQn. | (3.13) |
By (2.6) in Lemma 2.9, we have bQ=⟨χ,RG,Q⟩G, so
degDχ=#ˆπ(ˆSχX)+∑Q∉ˆπ(ˆSχX)⌊eQ−1eQ+1n⟨χ,RG,Q⟩G⌋−1n⟨χ,RG⟩G=mχ(ˆSχX). | (3.14) |
We summarize the discussion in this section.
Let X be an irreducible nodal curve over k, G be a finite automorphism group of order n on X, and Y=X/G be a smooth curve. Let ˆπ:ˆX→Y be the ramified cover induced by the normalization π:X→Y, and ˆSX⊆ˆX be the preimage of the singular points of X on ˆX. Denote RG the ramification module of ˆπ, and RG,Q the ramification module of Q∈Y.
Theorem 3.13 (The Chevalley-Weil formula on irreducible nodal curves). Let χ:G→k× be a character. If the quotient curve Y=X/G is smooth, then
dimkH0(X,ωX)χ=gY−1+mχ(ˆSχX)+⟨χ,1G⟩. | (3.15) |
Here ˆSχX is the singular χ-set of π as in (3.5), and in Definition 3.11,
mχ(ˆSχX)=#ˆπ(ˆSχX)+∑Q∉ˆπ(ˆSχX)⌊eQ−1eQ+1n⟨χ,RG,Q⟩G⌋−1n⟨χ,RG⟩G. |
In particular, when χ=1G, we have ˆSχX=∅ and dimkH0(X,ωX)G=gY. This is the special case that ˆSY=∅ in Proposition 3.5.
Example 3.14 (Hyperelliptic stable curves). We call a stable curve C a hyperelliptic stable curve if there exists a 2-order automorphism J:C→C such that C/⟨J⟩=P1. Such an automorphism of C is called an involution.
Let C be an irreducible hyperelliptic stable curve with N (≥1) nodes, and ˆC be the normalization of C with genus g. Then ˆπ:ˆC→P1 has 2g+2 fixed points, which are also all the ramification points of ˆπ, with ramification index all equal to 2. So the number of branch points on P1 is #Bl(P1)=2g+2 and no branch point lies in ˆπ(ˆSC).
Note that the Galois group G=⟨J⟩≅Z2 has two characters 1G and χ−, where χ−(J)=−1.
For χ=1G, we have dimkH0(C,ωC)G=g(P1)=0.
For χ−, we have ˆSχ−C=ˆSC, so sχ−=N. The induced character at any branch point P is
θP:GP=G→k∗,J↦−1. |
Therefore RG,Q=θP, so for all Q∈Bl(P1), we have ⟨χ−,RG,Q⟩G=1, and then ⟨χ−,RG⟩G=2g+2. We calculate
mχ−(ˆSχ−C)=sχ−+∑Q∈Bl(P1)⌊eQ−1eQ+12⟨χ−1,RG,Q⟩G⌋−12⟨χ−,RG⟩G=N+2g+2−(g+1)=pa(C)+1. | (3.16) |
Finally, we get dimkH0(C,ωC)χ−=g(P1)−1+mχ−(ˆSχ−C)=pa(C).
In this section, let X be a connected nodal curve with several irreducible components. Suppose that X admits a finite automorphism group G, and let π:X→Y=X/G be the quotient morphism. Consequently, Y=X/G is also a connected nodal curve.
We will prove that H0(X,ωX)G=π∗H0(Y,ωY) and then generalize Theorem 3.13 to the Chevalley-Weil formula for connected nodal curves, assuming the smoothness of Y for the same reasons stated in Remark 3.9.
Due to the abundance of symbols, in the following, we will replace some round brackets with square brackets, such as using H0(Y,ωY[IY]) instead of H0(Y,ωY(IY)).
Let Y=∪jYj be the decomposition into irreducible components. Then
∐Yj→Y |
is the normalization at the intersection points. Let IYj be the set of intersection points of Yj with other components, and we call it the intersection locus of Yj. Note that these points are nodes in the whole Y, but smooth points in each Yj.
Then we can decompose the regular differential of Y by the irreducible components:
H0(Y,ωY)↪⊕jH0(Yj,ωYj[IYj]),φ↦(φ|Yj)j. | (4.1) |
Furthermore let ˆYj→Yj be the normalization. Then we have
H0(Yj,ωYj[IYj])↪H0(ˆYj,ΩˆYj[ˆSYj⊔IYj]). |
Now let X′j:=π−1(Yj) be the preimage of an irreducible component. This is a nodal curve but could be disconnected and we have*
H0(X,ωX)↪⊕jH0(X′j,ωX′j[IX′j]), | (4.2) |
*If a curve X decomposes as X=X1⊔X2, then the space of differentials decomposes as H0(X,ωX)=H0(X1,ωX1)⊕H0(X2,ωX2).
where IX′j=π−1(IYj). In fact, the information of the eigenspace H0(X′j,ωX′j)χ of the regular differentials is contained in any single irreducible component.
Let X be a (possibly disconnected) nodal curve decomposed into its irreducible components
X=d⋃i=1Xi, |
and suppose that Y=X/G is irreducible. Then the action of G on the set of irreducible components {X1,…,Xd} is transitive, and in particular, all these irreducible components are isomorphic. Let S⊂X be a G-invariant divisor, and set Si:=S|Xi. Consequently, G acts on
⊕di=1H0(Xi,ωXi[Si]). |
More precisely, for any σ∈G we define the permutation of the indices by setting σ(i) via the relation σ(Xi)=Xσ(i). Then for any φ∈H0(Xi,ωXi[Si]) we have:
σφ∈H0(Xσ−1(i),ωXσ−1(i)[Sσ−1(i)]).† |
† σ acts by pullback on the differentials (that is, if σ:Xj→Xi and φ∈H0(Xi,ωXj), then σφ:=σ∗φ∈H0(Xj,ωXj).
Let Gi:={σ∈G∣σ(Xi)=Xi} denote the stabilizer subgroup of Xi. Then we have:
Proposition 4.1. Given a character χ:G→k× with restriction χi:=χ|Gi:Gi→k× to Gi, we have a canonical projection
pi:(⊕di=1H0(Xi,ωXi[Si]))χ→H0(Xi,ωXi[Si])χi, | (4.3) |
which is an isomorphism.
Proof. We show this holds for p1.
Suppose
(φ1,⋯,φd)∈(⊕di=1H0(Xi,ωXi[Si]))χ. |
Then for any σ∈G we have
σ(φ1,⋯,φd)=(σφi)σ−1(i)=(χ(σ)φi)i. |
Since G acts transitively on the set {X1,…,Xd}, for each i there exists some σi:Xi→X1. Hence, we can express the tuple as
(φ1,…,φd)=(φ1,χ(σ2)−1σ2φ1,…,χ(σd)−1σdφ1), |
which shows that the tuple (φi)i is completely determined by its first component φ1. This proves that the projection
p1:(⊕di=1H0(Xi,ωXi[Si]))χ⟶H0(X1,ωX1[S1]) |
is injective. It remains to show that φ1 lies in the χ1-eigenspace. For any element τ∈G1 (the stabilizer of X1), we have τ(1)=1 and χ(τ)=χ1(τ). Therefore,
τφ1=χ1(τ)φ1, |
which confirms that φ1 indeed belongs to H0(X1,ωX1[S1])χ1.
Conversely, let
φ1∈H0(X1,ωX1[S1])χ1, |
and for each i choose an isomorphism σi:Xi→X1 as before. We need to show that
(χ(σi)−1σiφ1)i |
defines an element of
(⊕di=1H0(Xi,ωXi[Si]))χ. |
First, note that if we choose two different isomorphisms σi,τi:Xi→X1, then τ−1iσi is an element of the stabilizer G1 of X1. Since φ1 lies in the χ1-eigenspace, we have
τ−1iσiφ1=χ1(τ−1iσi)φ1. |
Because the character χ restricts on G1 to χ1, this can be rewritten as
τ−1iσiφ1=χ(τi)−1χ(σi)φ1. |
Multiplying both sides on the left by χ(σi)−1χ(τi)τi yields
χ(σi)−1σiφ1=χ(τi)−1τiφ1. |
Thus, the section
(φ1,χ(σ2)−1σ2φ1,…,χ(σd)−1σdφ1) |
is independent of the particular choice of the isomorphisms σi.
Next, let τ∈G be arbitrary. We may assume that τ sends the index j to i, that is, τ(j)=i. Then, by the definition of the action, we have
τ(χ(σi)−1σiφ1)=χ(σi)−1(τσi)φ1. |
Since τσi and σj are both isomorphisms from Xj to X1, we have
χ(τσi)−1(τσi)φ1=χ(σj)−1σjφ1. |
Thus,
τ(χ(σi)−1σiφ1)=χ(σi)−1χ(τσi)χ(σj)−1σjφ1=χ(τ)χ(σj)−1σjφ1. |
It follows that
τ(φ1,χ(σ2)−1σ2φ1,…,χ(σd)−1σdφ1)=χ(τ)(φ1,χ(σ2)−1σ2φ1,…,χ(σd)−1σdφ1). |
In conclusion, we obtain that p1 is an isomorphism.
Let Xj be a chosen irreducible component of X′j. By Proposition 4.1, there are immersions
H0(X′j,ωX′j)χ↪H0(Xj,ωXj[IXj])χj, |
and
H0(X′j,ωX′j[IX′j])χ↪H0(Xj,ωXj[I′j⊔IXj])χj, |
where χj=χ|Gj. Consequently, we obtain the chain of inclusions
H0(X,ωX)χ↪⊕jH0(Xj,ωXj[I′j⊔IXj])χj↪⊕jH0(ˆXj,ΩˆXj[ˆSXj⊔IXj⊔I′j])χj. | (4.4) |
Here:
● Gj={σ∈G∣σ(Xj)=Xj} and |Gj|=|G|/dj if X′j has dj irreducible components.
● I′j=IX′j∩Xj.
● IXj is the set of intersection points of Xj with other components in X′j.
● ˆSXj is the preimage of nodes in the normalization ˆαj:ˆXj→Xj.
Now we consider the G-invariant regular differentials. Suppose χ=1G and consider the following diagram as (3.2):
![]() |
(4.5) |
Lemma 4.2. There is a canonical inclusion
π∗H0(Y,ωY)⊆H0(X,ωX) |
that makes the diagram above commute.
Proof. Let
φY∈H0(Y,ωY) |
be a regular differential on Y and consider its image in the lower left-hand corner:
ˆπ∗φY∈⊕jH0(ˆXj,ΩˆXj[ˆSXj⊔IXj⊔I′j])Gj. |
We need to verify that the residues of ˆπ∗φY satisfy the relation
ResP1(ˆπ∗φY)=−ResP2(ˆπ∗φY), |
where {P1,P2} is a pair in ˆX, i.e., P1 and P2 are points lying over the same node P in X.
Without loss of generality, assume that P1∈ˆX1, the normalization of X1, which is a chosen irreducible component of X′1. Then
ResP1(ˆπ∗φY)=eP1⋅Resˆπ(P1)φY, |
where eP1 is the ramification index of P1 in the cover of smooth curves
ˆπ1:ˆX1→ˆY1. |
According to Lemma 3.3, the remaining task is to establish the equality
eP1=eP2. |
To see this, let lP denote the total number of elements in the orbit of the point P (which is common to both P1 and P2) under the action of the group G. Then the orbit of P1 has exactly lP/d1 elements in ˆX1, where d1 is the number of irreducible components in X′1. Hence, by the very definition of the ramification index we obtain
eP1=|G1|lP/d1=|G|lP. |
By the same argument to P2, we also have
eP2=|G|lP. |
This completes the proof.
The same argument as Lemma 3.4 tells us
H0(X,ωX)G↪⊕jH0(ˆXj,ΩˆXj[ˆπ−1j(ˆSYj⊔IYj)])Gj. |
With the notations above, we have:
Theorem 4.3. The upper and lower row of the canonical commutative diagram
![]() |
are both isomorphisms. In particular, we have dimkH0(X,ωX)G=pa(Y).
Proof. By Proposition 3.6, we have
ˆπ∗jH0(ˆYj,ΩˆYj[ˆSYj⊔IYj])∼→H0(ˆXj,ΩˆXj[ˆπ−1j(ˆSYj⊔IYj)])Gj. |
So we have the isomorphism for the lower row. Moreover, both H0(X,ωX)G and H0(Y,ωY) are the subspaces of log differentials satisfying the residue relations, so we have the isomorphism for the upper row.
Let Y be smooth for the remaining part of this section. We are going to calculate dimkH0(X,ωX)χ for general χ.
Let X=∪di=1Xi be the decomposition of irreducible components, and then we can decompose the regular differential space on X into each irreducible component:
H0(X,ωX)↪⊕di=1H0(Xi,ωXi[Ii]),φ↦(φ|Xi), | (4.6) |
where Ii is the intersection locus of Xi.
Since Y is smooth, then by the criterion in Remark 3.9, set
Iχi:={P∈Ii∣∃τ∈GP−Gi s.t. χ(τ)=−1} | (4.7) |
as those intersection points that could be the poles of φ|Xi for φ∈H0(X,ωX)χ, and then we have the isomorphism
H0(X,ωX)χ∼→(⊕di=1H0(Xi,ωXi[Iχi]))χ. | (4.8) |
Applying Proposition 4.1 shows that
(⊕di=1H0(Xi,ωXi[Iχi]))χ∼→H0(X1,ωX1[Iχ1])χ1. | (4.9) |
So far, our research object has been reduced to the action of the stabilizer subgroup G1 on the irreducible nodal curve X1.
Remark 4.4. With the notations above, note that Iχi=∅ when χ=1G. For the quotient morphism π1:X1→X1/G1∼=Y, we have
H0(X,ωX)G∼→H0(X1,ωX1)G1∼→H0(Y,ωY). | (4.10) |
This first isomorphism is by (4.9) and the second is by Theorem 3.5. This is a special case of Theorem 4.3, and we have
dimkH0(X,ωX)G=g(Y). | (4.11) |
Now it remains to calculate dimkH0(X1,ωX1[Iχ1])χ1.
Let ˆπ1:^X1→Y be the normalization of π1, Bl(Y) the branch locus, RG1 the ramification module of ˆπ1, and RG1,Q the ramification module of Q∈Y. If we let #G=n, then n1:=#G1=n/d.
Let ˆSχ1X1 be the singular χ1-set of X1 as (3.5), and then we have the isomorphism
H0(X1,ωX1[Iχ1])χ1∼→H0(ˆX1,ΩˆX1[ˆSχ1X1∪Iχ1])χ1 | (4.12) |
by the same argument as in Proposition 3.7.
By Proposition 2.7 (2.5) again, we have
H0(^X1,ΩˆX1[ˆSχ1X1∪Iχ1])χ1=fχ1⋅ˆπ∗1H0(Y,ΩY⌊ n−11π∗(ˆSχ1X1∪Iχ1+(fχ1)+Rˆπ1)⌋), |
where fχ1∈K(X1)× satisfies σfχ1=χ1(σ)fχ1,∀σ∈G1.
Set Dχ1:=⌊ n−11π∗(ˆSχ1X1∪Iχ1+(fχ1)+Rˆπ1)⌋. By the Riemann-Roch theorem, we have
dimkH0(X,ωX)χ=dimkH0(Y,ΩY(Dχ1))=dimkH0(Y,OY(−Dχ1))+degDχ1+gY−1. | (4.13) |
Applying the calculation of Lemma 3.12, we have
degDχ1=mχ1(ˆSχ1X1∪Iχ1), |
and see that
mχ1(ˆSχ1X1∪Iχ1)=#ˆπ1(ˆSχ1X1∪Iχ1)+∑Q∉ˆπ1(ˆSχ1X1∪Iχ1)⌊eQ−1eQ+dn⟨χ1,RG1,Q⟩G1⌋−dn⟨χ1RG1⟩G1 | (4.14) |
in Definition 3.11.
Finally, through the discussion of Lemma 3.10, we get
dimkH0(Y,OY(−Dχ1))=δχ, | (4.15) |
where δχ=0 or 1. δχ=1 if and only if Dχ1 is principal; in this case, Iχ1=∅ and χ1=1G1. ‡
‡In this context, we require that ˆSχ1X1∪Iχ1 is empty. In particular, when χ1=1G1, note that ˆSχ1X1=∅; however, Iχ1 is determined by χ rather than χ1. See Example 4.6 where even though χ−1=id, Iχ−1 is nonempty.
In summary, we have obtained the following:
Let X be a connected nodal curve with d irreducible components, and G an automorphism group on X of order n. Assume that the quotient curve Y=X/G is smooth. Then there exists a canonical ramified cover ˆπ1:ˆX1→X1/G1≃Y for an irreducible component X1 with G1={σ∈G∣σ(X1)=X1}.
Theorem 4.5 (The Chevalley-Weil formula on connected nodal curves). Let χ:G→k× be a character. If the quotient curve Y=X/G is smooth, then
dimkH0(X,ωX)χ=gY−1+mχ1(ˆSχ1X1∪Iχ1)+δχ, | (4.16) |
where χ1 is the restriction of χ on G1, and see mχ1(ˆSχ1X1∪Iχ1) in (4.14) and δχ in (4.15).
In particular, when χ=1G, we have dimkH0(X,ωX)G=gY.
Example 4.6. Let the curves C1≃C2≃P1 intersect transversely at m>2 points, and consider the hyperelliptic stable curve C=C1∪C2, with the involution J which permutes C1 and C2. Then we have the quotient mapping π:C→C/⟨J⟩=P1. The arithmetic genus of C is pa(C)=m−1 (see [11, Proposition 7.5.4 and Lemma 10.3.18]) and the Galois group ⟨J⟩ has two characters 1G and χ−.
For χ−, note that the cover π1:C1→P1 induced by π is an isomorphism, so we have χ−1=id. Therefore ˆSidC1=∅ and mχ−1(ˆSidC1∪Iχ−1)=#π(Iχ−1)=m. By Theorem 4.5, we have
dimkH0(C,ωC)χ−=gP1−1+m=m−1=pa(C). | (4.17) |
For χ=1G, by (4.11), we have dimkH0(C,ωC)G=pa(P1)=0.
In this work, we have systematically investigated the eigenspaces of regular differentials on a nodal curve X under the action of a finite automorphism group G. Our primary achievement is the demonstration that the dimension of the space of G-invariant regular differentials is the arithmetic genus of the quotient curve X/G, providing a non-trivial generalization of the smooth case. Furthermore, under the condition that X/G is smooth, we have successfully extended the Chevalley-Weil formula to the nodal setting, obtaining an exact expression for the dimension of the eigenspace for any one-dimensional character χ. This was accomplished by first addressing irreducible nodal curves and incorporating a singular χ-set, then reducing the analysis on a conncected nodal curve to that on one of its irreducible components. This reduction involves adding correction terms that account for certain intersection points on that component. These explicit formulas illuminate how nodal singularities (including intersection points) contribute to the dimensions of these eigenspaces, advancing our understanding of group actions on differentials of nodal curves.
The author declares that he has not used Artificial Intelligence (AI) tools in the creation of this article.
This article is part of the author's doctoral thesis presented to Xiamen University. The author would like to thank his supervisor Wenfei Liu for his advice and support, and he is grateful to Professor Qing Liu for helpful discussions during his visit to the University of Bordeaux. This work has been supported by the NSFC (No.11971399) and by the Presidential Research Fund of Xiamen University (No.20720210006).
The author declares no competing interests.
[1] |
B. Y. Chen, S. Ishikawa, Biharmonic pseudo-Riemannian submanifolds in pseudo-Euclidean spaces, Kyushu J. Math., 52 (1998), 167–185. https://doi.org/10.2206/kyushujm.52.167 doi: 10.2206/kyushujm.52.167
![]() |
[2] | G. Y. Jiang, Some non-existence theorems of 2-harmonic isometric immersions into Euclidean spaces, Chin. Ann. Math. Ser. A, 8 (1987), 376–383. |
[3] |
F. D. O. Leuven, Hypersurfaces of E4 with harmonic mean curvature vector, Math. Nachr., 196 (1998), 61–69. https://doi.org/10.1002/mana.19981960104 doi: 10.1002/mana.19981960104
![]() |
[4] |
T. Hasanis, T. Vlachos, Hypersurface in E4 with harmonic mean curvature vector field, Math. Nachr., 172 (1995), 145–169. https://doi.org/10.1002/mana.19951720112 doi: 10.1002/mana.19951720112
![]() |
[5] |
Y. Fu, M. C. Hong, X. Zhan, On Chen's biharmonic conjecture for hypersurfaces in R5, Adv. Math., 383 (2021), 107697. https://doi.org/10.1016/j.aim.2021.107697 doi: 10.1016/j.aim.2021.107697
![]() |
[6] |
B. Y. Chen, M. I. Munteanu, Biharmonic ideal hypersurfaces in Euclidean spaces, Differ. Geom. Appl., 31 (2013), 1–16. https://doi.org/10.1016/j.difgeo.2012.10.008 doi: 10.1016/j.difgeo.2012.10.008
![]() |
[7] |
S. Montaldo, C. Oniciuc, A. Ratto, On cohomogeneity one biharmonic hypersurfaces into the Euclidean space, J. Geom. Phys., 106 (2016), 305–313. https://10.1016/j.geomphys.2016.04.012 doi: 10.1016/j.geomphys.2016.04.012
![]() |
[8] | N. Mosadegh, E. Abedi, Biharmonic hypersurfaces with recurrent operators in the Euclidean space, arXiv preprint, (2021), arXiv: math/2108.06193v1. https://doi.org/10.48550/arXiv.2108.06193 |
[9] | Y. L. Ou, B. Y. Chen, Biharmonic Submanifolds and Biharmonic Maps in Riemannian Geometry, World Scientific, Hackensack, USA, 2020. https://doi.org/10.1142/11610 |
[10] |
B. Y. Chen, S. Ishikawa, Biharmonic surfaces in pseudo-Euclidean spaces, Mem. Fac. Sci. Kyushu Univ. Ser. A, Math., 45 (1991), 323–347. https://doi.org/10.2206/kyushumfs.45.323 doi: 10.2206/kyushumfs.45.323
![]() |
[11] |
F. Defever, G. Kaimakamis, V. Papantoniou, Biharmonic hypersurfaces of the 4-dimensional semi-Euclidean space E4s, J. Math. Anal. Appl., 315 (2006), 276–286. https://doi.org/10.1016/j.jmaa.2005.05.049 doi: 10.1016/j.jmaa.2005.05.049
![]() |
[12] | Y. Fu, Z. H. Hou, D. Yang, X. Zhan, Biharmonic hypersurface in E5s, Acta Math. Sinica (Chinese Ser.), 65 (2022), 335–352. |
[13] |
J. C. Liu, C. Yang, Hypersurfaces in En+1s satisfying Δ→H=λ→H with at most three distinct principal curvatures, J. Math. Anal. Appl., 419 (2014), 562–573. http://dx.doi.org/10.1016/j.jmaa.2014.04.066 doi: 10.1016/j.jmaa.2014.04.066
![]() |
[14] |
L. Du, On η-biharmonic hypersurfaces with constant scalar curvature in higher dimensional pseudo-Riemannian space forms, J. Math. Anal. Appl., 518 (2023), 126670. https://doi.org/10.1016/j.jmaa.2022.126670 doi: 10.1016/j.jmaa.2022.126670
![]() |
[15] | B. Y. Chen, Total Mean Curvature and Submanifolds of Finite Type, World Scientific, Hackensack, USA, 2014. |