The aim of this paper is to establish the second main theorem for holomorphic curves from the annulus into a complex projective variety intersecting an arbitrary family of hypersurfaces. This is done by using the notion of "Distributive Constant" for a family of hypersurfaces with respect to a complex projective variety developed by Quang. We also give an explicit estimate for the level of truncation.
Citation: Liu Yang, Yuehuan Zhu. Second main theorem for holomorphic curves on annuli with arbitrary families of hypersurfaces[J]. Electronic Research Archive, 2024, 32(2): 1365-1379. doi: 10.3934/era.2024063
[1] | Won-Ki Seo . Fredholm inversion around a singularity: Application to autoregressive time series in Banach space. Electronic Research Archive, 2023, 31(8): 4925-4950. doi: 10.3934/era.2023252 |
[2] | Vladimir Rovenski . Willmore-type variational problem for foliated hypersurfaces. Electronic Research Archive, 2024, 32(6): 4025-4042. doi: 10.3934/era.2024181 |
[3] | Sang-Eon Han . Semi-Jordan curve theorem on the Marcus-Wyse topological plane. Electronic Research Archive, 2022, 30(12): 4341-4365. doi: 10.3934/era.2022220 |
[4] | Dong Li, Xiaxia Wu, Shuling Yan . Periodic traveling wave solutions of the Nicholson's blowflies model with delay and advection. Electronic Research Archive, 2023, 31(5): 2568-2579. doi: 10.3934/era.2023130 |
[5] | Ping Liu, Junping Shi . A degenerate bifurcation from simple eigenvalue theorem. Electronic Research Archive, 2022, 30(1): 116-125. doi: 10.3934/era.2022006 |
[6] | Li Du, Xiaoqin Yuan . The minimality of biharmonic hypersurfaces in pseudo-Euclidean spaces. Electronic Research Archive, 2023, 31(3): 1587-1595. doi: 10.3934/era.2023081 |
[7] | Chenghua Gao, Enming Yang, Huijuan Li . Solutions to a discrete resonance problem with eigenparameter-dependent boundary conditions. Electronic Research Archive, 2024, 32(3): 1692-1707. doi: 10.3934/era.2024077 |
[8] | Xiaoyan Xu, Xiaohua Xu, Jin Chen, Shixun Lin . On forbidden subgraphs of main supergraphs of groups. Electronic Research Archive, 2024, 32(8): 4845-4857. doi: 10.3934/era.2024222 |
[9] | Kwok-Pun Ho . Martingale transforms on Banach function spaces. Electronic Research Archive, 2022, 30(6): 2247-2262. doi: 10.3934/era.2022114 |
[10] | Fabrizio Catanese, Luca Cesarano . Canonical maps of general hypersurfaces in Abelian varieties. Electronic Research Archive, 2021, 29(6): 4315-4325. doi: 10.3934/era.2021087 |
The aim of this paper is to establish the second main theorem for holomorphic curves from the annulus into a complex projective variety intersecting an arbitrary family of hypersurfaces. This is done by using the notion of "Distributive Constant" for a family of hypersurfaces with respect to a complex projective variety developed by Quang. We also give an explicit estimate for the level of truncation.
The main result in Nevanlinna theory is called the second fundamental theorem. In 1925, Nevanlinna [1] established the second main theorem for meromorphic functions on the complex plane C. In 1933, H. Cartan [2] proved the second main theorem for holomorphic curves with targets in the form of hyperplanes in the general position in the complex projective spaces Pn(C). In 1983, E. I. Nochka [3] proved the second main theorem in the case of hyperplanes in the N-subgeneral position in Pn(C) with ramification. In 2004, M. Ru [4] established the second main theorem for holomorphic curves with targets in the form of hypersurfaces in the general position in Pn(C) without ramification. In 2009, Ru [5] made further extension to algebraically nondegenerate holomorphic curves into an arbitrary smooth complex projective variety. Since that time, the problem of investigation of the characteristics of holomorphic maps has attracted the attention of numerous authors.
In this paper, we mainly consider the case for holomorphic curves from the doubly connected domain into Pn(C). By the doubly connected mapping theorem [6], each doubly connected domain in C is conformally equivalent to the annulus A(r,R)={z:r<|z|<R},0≤r<R≤+∞. We need only consider two cases: r=0,R=+∞ simultaneously and 0<r<R<+∞. In the latter case the homothety z↦z√rR reduces the given domain to the annulus {z:1R0<|z|<R0} with R0=√Rr. Thus, in the two cases every annulus is invariant with respect to the inversion z↦1z. Observing the above facts, Khrystiyanyn and Kondratyuk [7,8] indicated the way to extend the value distribution of Nevanlinna theory to meromorphic functions in annuli.
Let R0<+∞ be a fixed positive real number or +∞ and let
A={z∈C:1R0<|z|<R0} |
be an annuli in C. Moreover, for any real number r such that 1<r<R0, we denote
Ar={z∈C:1r<|z|<r}, |
A1,r={z∈C:1r<|z|≤1}, |
and
A2,r={z∈C:1<|z|<r}. |
Let f=(f0:…:fn+1) be a holomorphic curve from the annuli A into the complex projective space Pn(C). For 1<r<R0, the characteristic function of f is defined by
Tf(r)=12π∫2π0log‖f(reiθ)‖dθ+12π∫2π0log‖f(r−1eiθ)‖dθ. |
where
‖f(z)‖=max{|f0(z)|,…,|fn(z)|}. |
Remark 1. The above definition is independent, up to an additive constant, of the choice of the reduced representation of f.
Let D be a hypersurface in Pn(C) of degree d. Let Q be the homogeneous polynomial of degree d, defining D. The proximity function of f is defined by
mf(r,D):=12π∫2π0log‖f(reiθ)‖d|Q∘f(reiθ)|dθ+12π∫2π0log‖f(r−1eiθ)‖d|Q∘f(r−1eiθ)|dθ. |
To be within an additive constant, this definition is independent of the choice of the reduced representation of f and the choice of the defining polynomial Q.
Further, for j=1,2, by nj,f(r,D) we denote the number of zeros of Q∘f in Aj,r, counting multiplicity, and let nMj,f(r,D) be the number of zeros of Q∘f in the disk Aj,r, where any zero of multiplicity greater than M is “truncated” and counted as if it only had multiplicity M. We set
N1,f(r,D):=∫1r−1n1,f(t,D)tdt,N2,f(r,D):=∫r1n2,f(t,D)tdt,N[M]1,f(r,D):=∫1r−1nM1,f(t,D)tdt,N[M]2,f(r,D):=∫r1nM2,f(t,D)tdt. |
The integrated counting and truncated counting functions are defined by
Nf(r,D):=N1,f(r,D)+N2,f(r,D),N[M]f(r,D):=N[M]1,f(r,D)+N[M]2,f(r,D). |
When we want to emphasize Q, we sometimes also write Nf(r,D) as Nf(r,Q) and N[M]f(r,D) as N[M]f(r,Q).
In the present paper, we set the small error term by
Of(r)={O(logr+logTf(r)),ifR0=+∞,O(log1R0−r+logTf(r)),ifR0<+∞. |
In 2015, H. T. Phuong and N. V. Thin [9] considered the extension of the second main theorem for holomorphic curves from A into Pn(C) crossing a finite set of fixed hyperplanes in general position.
Theorem A. ([9]) Let f:A→Pn(C) be a linearly nondegenerate holomorphic curve. Let H1,…,Hq be hyperplanes in Pn(C), located in general position, then we have
||(q−n−1)Tf(r)≤q∑j=1N[n]f(r,Hj)+Of(r). |
Here and in the following, the notation “||P” means that if R0=+∞, then the assertion P holds for all r∈(1,+∞) outside a set A′r with ∫A′rrλ−1dr<+∞. At the same time, if R0<+∞, then the assertion P holds for all r∈(1,R0) outside a set A′r with ∫A′r1(R0−r)λ+1dr<+∞, where λ≥0.
Recently, J.L. Chen and T.B. Cao [10] obtained the second main theorem for holomorphic curves on annuli crossing a finite set of moving hyperplanes in sub-general position in Pn(C). It is well known that all known second main theorems and uniqueness results hold under the conditions that the hyperplanes or hypersurfaces are located in general position (or in sub-general position). More recently, S.D. Quang [11] introduced the notion of “distributive constant” ΔV of a family of moving hypersurfaces with respect to a subvariety V of Pn(C) and generalized some results to the case of meromorphic mappings into a projective subvariety V and an arbitrary family of moving hypersurfaces. Motivated by this new notion, we show the second main theorem for holomorphic curves from the annulus into the complex projective space which is ramified over an arbitrary family of hypersurfaces. We also give an explicit estimate for the level of truncation.
For the purpose of this article, we recall some definitions. For a subvariety V and an analytic subset S of Pn(C), the codimension of S in V is defined by
codimVS:=dimV−dim(V∩S). |
According to Quang [11], we give the following definition.
Definition 1. Let {Dj}qj=1 be the hypersurfaces in Pn(C). Denote by Q the index set {1,…,q}. Let V be a subvariety of Pn(C) of dimension k. Assume that V is not contained in any Dj(j∈Q). We define the distributive constant of {Dj}qj=1 with respect to V by
ΔV:=maxΓ⊂Q♯ΓcodimV(⋂j∈ΓDj). |
Definition 2. Let {Dj}qj=1 be the hypersurfaces in Pn(C). Denote by Q the index set {1,…,q}. Let N≥n and q≥N+1. The family {Dj}qj=1 is said to be in N-subgeneral position with respect to V if for every subset R⊂Q with the cardinality ♯R=N+1, then
⋂j∈RDj∩V=∅. |
If N=dimV then we say that {Dj}qj=1 is in general position with respect to V.
Remark 2. If the family {Dj}qj=1 is in N-subgeneral position with respect to V, then ΔV≤N−dimV+1 (see [11]).
In this paper, we establish the following second main theorem for holomorphic curves from the annulus into a complex projective variety intersecting an arbitrary family of hypersurfaces. The proof of our result follows from the paper [12,13].
Theorem 1. Let V be a projective subvariety of Pn(C) of dimension k. Let f:A→V be an algebraically nondegenerate holomorphic curve with 0<R0≤+∞. Let {Dj}qj=1 be q hypersurfaces in Pn(C) with degDj=dj(1≤j≤q). Let d be the least common multiple of d′js, i.e., d=lcm(d1,…,dq). Let ΔV be the distributive constant of {Dj}qj=1 with respect to V, then for any ε>0,
||(q−ΔV(k+1)−ε)Tf(r)≤q∑j=11djN[Mε]f(r,Dj)+Of(r), |
where
Mε=⌊deg(V)k+1ekdk2+kΔkV(2k+4)klkε−k⌋ |
with l=(k+1)q!.
Here and in the following, ⌊x⌋ denotes the greatest integer less than or equal to the real number x.
By Remark 2, we have an immediate corollary.
Corollary 1. Let V be a projective subvariety of Pn(C) of dimension k. Let f:A→V be an algebraically nondegenerate holomorphic curve with 0<R0≤+∞. Let D1,…,Dq be the hypersurfaces in Pn(C), located in N-subgeneral position with respect to V with dj:=degDj(1≤j≤q). Let d be the least common multiple of d′js, i.e., d=lcm(d1,…,dq). Then for any ε>0,
||(q−(N−k+1)(k+1)−ε)Tf(r)≤q∑j=11djN[˜Mε]f(r,Dj)+Of(r). |
where
˜Mε=⌊deg(V)k+1ekdk2+k(N−k+1)k(2k+4)klkε−k⌋ |
with l=(k+1)q!.
To prove our result, we need the following second main theorem for holomorphic curves on the annulus (see [10,14]).
Lemma 1. (A general form of the second main theorem) Let f:A→Pn(C) be a linearly nondegenerate holomorphic curve (i.e. its image is not contained in any proper subspace of Pn(C)). Let H1,…,Hq (or linear forms a1,…,aq) be arbitrary hyperplanes in Pn(C), then
‖∫2π0maxK∑j∈Klog‖f(reiθ)‖‖⟨f(reiθ);Hj⟩‖dθ2π+∫2π0maxK∑j∈Klog‖f(r−1eiθ)‖‖⟨f(r−1eiθ);Hj⟩‖dθ2π≤(n+1)Tf(r)−NW(r,0)+Of(r). |
Here, the maximum is taken over all subsets K of {1,…,q} such that the linear forms aj,j∈K, are linearly independent.
We also need the following lemmas.
Lemma 2. ([11,15]) Let V be a projective subvariety of Pn(C) of dimension k. Let D0,…,Dp be p+1 hypersurfaces in Pn(C) of the same degree d≥1, such that ⋂pi=0Di∩V=∅ and
dim(s⋂i=0Di)∩V=k−l,∀tl−1≤s<tl,1≤l≤k |
where t0,t1,…,tk integers with 0=t0<t1<⋯<tk=p, then there exist k+1 hypersurfaces P0,…,Pk in Pn(C) of the forms
Pl=tl∑j=0cljDj,clj∈C,l=0,…,k |
such that (⋂kl=0Pl)∩V=∅.
Lemma 3. ([16]) Let {Qi}i∈R be a set of hypersurfaces in Pn(C) of the common degree d, let V be a projective subvariety of Pn(C) and let f be a meromorphic mapping of Cm into V. Assume that ⋂i∈RQi∩V=∅, then there exist positive constants α and β such that
α‖f‖d≤maxi∈R|Qi(f)|≤β‖f‖d. |
Lemma 4. ([11]) Let t0,t1,…,tn be n+1 integers such that 1=t0<t1<⋯<tn, and let Δ=max1≤s≤nts−t0s, then for every n real numbers a0,a1,…,an−1 with a0≥a1≥⋯≥an−1≥1, we have
at1−t00at2−t11⋯atn−tn−1n−1≤(a0a1⋯an−1)Δ. |
We recall the notion of Chow weights and Hilbert weights from [5] (see also [17]). Let X⊂Pn(C) be a projective variety of dimension k and degree δ. The Chow form of X is the unique polynomial, up to a constant scalar,
FX(u0,…,uk)=FX(u00,…,u0n;…;uk0,…,ukn) |
in n+1 blocks of variables ui=(ui0,…,uin),i=0,…,k with the following properties:
(ⅰ) FX is irreducible in k[u00,…,ukn];
(ⅱ) FX is homogeneous of degree δ in each block ui,i=0,…,k;
(ⅲ) FX(u0,…,uk)=0, if and only if, X∩Hu0∩⋯∩Huk≠∅, where Hui,i= 0,…,k, are the hyperplanes given by
ui0x0+⋯+uinxn=0. |
Let c=(c0,…,cn) be a tuple of real numbers and t be an auxiliary variable. We consider the decomposition
FX(tc0u00,…,tcNu0n;…;tc0uk0,…,tcnukn) |
=te0G0(u0,…,un)+⋯+terGr(u0,…,un) |
with G0,…,Gr∈C[u00,…,u0n;…;uk0,…,ukn] and e0>e1>⋯>er. The Chow weight of X with respect to c is defined by
eX(c):=e0 |
For each subset J={j0,…,jk} of {0,…,n} with j0<j1<⋯<jk, we define the bracket
[J]=[J](u0,…,uk):=det(uijt),i,t=0,…,k, |
where ui=(ui0,…,uin)(1≤i≤k) denote the blocks of n+1 variables. Let J1,…,Jβ with β=(n+1k+1) be all subsets of {0,…,n} of cardinality k+1.
Therefore, FX can be written as a homogeneous polynomial of degree δ in [J1],…,[Jβ]. We may see that for c=(c0,…,cn)∈Rn+1 and for any J among J1,…,Jβ,
(tc0u00,…,tcnu0n,…,tc0uk0,…,tcnukn)=t∑j∈Jcj[J](u00,…,u0n,…,uk0,…,ukn) |
For a=(a0,…,an)∈Zn+1 we write xa for the monomial xa00⋯xann. Denote by C[x0,…,xn]u the vector space of homogeneous polynomials in C[x0,…,xn] of degree u (including 0). For an ideal I in C[x0,…,xn], we put
Iu:=C[x0,…,xn]u∩I. |
Let I(X) be the prime ideal in C[x0,…,xn] defining X. The Hilbert function HX of X is defined by, for u=1,2,…,
HX(u):=dim(C[x0,…,xn]u/I(X)u) |
By the usual theory of Hilbert polynomials,
HX(u)=δ⋅unn!+O(un−1). |
The u-th Hilbert weight SX(u,c) of X with respect to the tuple c=(c0,…,cn)∈ Rn+1 is defined by
SX(u,c):=max(HX(u)∑i=1ai⋅c) |
where the maximum is taken over all sets of monomials xa1,…,xaHX(u) whose residue classes modulo I form a basis of C[x0,…,xn]u/Iu. The following theorems are due to J. Evertse and R. Ferretti [18].
Lemma 5. Let X⊂Pn(C) be an algebraic variety of dimension k and degree δ. Let u>δ be an integer and let c=(c0,…,cn)∈Rn+1≥0, then
1uHX(u)SX(u,c)≥1(k+1)δeX(c)−(2k+1)δu⋅(maxi=0,…,nci). |
Lemma 6. Let Y⊂Pn(C) be an algebraic variety of dimension k and degree δ. Let c=(c1,…,cq) be a tuple of positive reals. Let {i0,…,in} be a subset of {1,…,q} such that
Y∩{yi0=⋯=yik=0}=∅, |
then
eY(c)≥(ci0+⋯+cik)δ. |
Proof. Assume that V is a projective subvariety of Pn(C) of dimension k. If there exists i0∈Q={1,2,…,q} such that ∩j∈Q∖{i0}Dj⋂V≠∅, then it follows from the definition that
ΔV≥q−1k−dim(∩j∈Q∖{i0}Dj⋂V)≥q−1k>qk+1. |
Hence, q<ΔV(k+1), which implies the conclusion of Theorem 1 is trivial. Therefore, we only need to consider the case that for each i∈Q, the set ∩j∈Q∖{i}Dj⋂V=∅.
Replacing Dj by Dd/djj if necessary, without loss of generality, we may assume that all hypersurfaces D1,…,Dq are of the same degree d. We denote by {σi}i∈I the set of all permutations of the set {1,…,q}, where I={1,2,⋯,n0} and n0=q!. For each i∈I, since ⋂q−1j=1Dσi(j)∩V=∅, there exist k+1 integers ti,0,ti,1,…,ti,k with 1=ti,0<⋯<ti,k=pi, where pi≤q−1 such that ⋂pij=1Dσi(j)∩V=∅ and
dim(s⋂j=1Dσi(j)∩V)=k−l∀ti,l−1≤s<ti,l,1≤l≤k. |
For each i∈I, we denote by Pi,0,…,Pi,k the hypersurfaces obtained in Lemma 2 with respect to the hypersurfaces Dσi(1),…,Dσi(pi).
We consider the mapping Φ from V into Pl(C)(l=n0(k+1)−1), which maps a point x∈V into the point Φ(x)∈Pl(C) given by
Φ(x)=(P1,0(x):⋯:P1,k(x):P2,0(x):⋯:P2,k(x):⋯:Pn0,0(x):⋯:Pn0,k(x)) |
Let Y:=Φ(V). Since V∩(⋂kj=0P1,j)=∅,Φ is a finite morphism on V and Y is a complex projective subvariety of Pl(C) with dimY=k and
δ:=degY≤dk⋅degV. |
For every
a=(a1,0,…,a1,k,a2,0,…,a2,k,…,an0,0,…,an0,k)∈Zl+1≥0 |
and
y=(y1,0,…,y1,k,y2,0,…,y2,k,…,yn0,0,…,yn0,k) |
we denote ya=ya1,01,0⋯ya1,k1,k⋯yan0,0n0,0⋯yan0,kn0,k. Let u be a positive integer. We set
nu:=HY(u)−1,mu:=(m+u−1u)−1 |
and define the space
Yu=C[y0,…,yl]u/(IY)u, |
which is a vector space of dimension HY(u). We fix a basis {v1,…,vHY(u)} of Yu and consider the meromorphic mapping F with a reduced representation
˜F=(v1(Φ∘˜f),…,vHY(u)(Φ∘˜f)):A→Cnu+1. |
Since f is algebraically nondegenerate, the holomorphic curve F:A→Pnu(C) is linearly nondegenerate (i.e., its image is not contained in any hyperplanes in Pnu(C)).
By Lemma 3, there exists a constant A>0, which is chosen common for all i∈I, such that
‖˜f(z)‖d≤Amax0≤j≤pi|Dσi(j)(˜f(z))|. |
According to the definition of Pi,j, we may choose a positive constant B≥1, commonly for all i∈I, such that
|Pi,j(x)|≤Bmax1≤s≤ti,j|Dσi(s)(x)|, |
for all x=(x0,…,xn)∈Cn+1 and for all 0≤j≤k. It is easily seen that, there exists a positive constant C, such that
|Pi,j(x)|≤C‖x‖d |
for all x=(x0,…,xn)∈Cn+1,1≤i≤n0, and 0≤j≤k.
Fix an element i∈I. Denote by S(i) the set of all points
z∈△(R)∖{q⋃i=1Di(˜f(z))−1({0})∪⋃0≤j≤ki∈IPi,j(˜f(z))−1({0})} |
such that
|Dσi(1)(˜f(z))|≤|Dσi(2)(˜f(z))|≤⋯≤|Dσi(q)(˜f(z))|. |
Therefore, for each z∈S(i), by Lemma 4 we have
q∏i=1‖˜f(z)‖d|Di(˜f(z))|≤Aq−pipi∏j=1‖˜f(z)‖d|Dσi(j)(˜f(z))|≤Aq−pik∏j=1(‖˜f(z)‖d|Dσi(tj−1)(˜f(z))|)ti,j−ti,j−1≤Aq−pik∏j=1(‖˜f(z)‖d|Dσi(tj−1)(˜f(z))|)ΔV≤Aq−piBkΔVk−1∏j=0(‖˜f(z)‖d|Pi,j(˜f(z))|)ΔV≤Aq−piBkΔVCΔVk∏j=0(‖˜f(z)‖d|Pi,j(˜f(z))|)ΔV, |
which implies that
logq∏i=1‖˜f(z)‖d|Di(˜f(z))|≤log(Aq−piBkΔVCΔV)+ΔVlogk∏j=0(‖˜f(z)‖d|Pi,j(˜f(z))|). | (3.1) |
Now, we fix an index i∈I and a point z∈S(i) and define
cz=(c1,0,z,…,c1,k,z,c2,0,z,…,c2,k,z,…,cn0,0,z,…,cn0,k,z)∈Rl+1≥0, |
where ci,j,z:=log‖˜f(z)‖d‖Pi,j‖|Pi,j(˜f(z))| for i=1,…,n0 and j=0,…,k. By the definition of the Hilbert weight, there are a1,…,aHY(u)∈Zl+1≥0 with
ai,z=(ai,1,0,z,…,ai,1,k,z,…,ai,n0,,z,…,ai,n0,k,z),ai,j,s,z∈{1,…,mu+1}, |
such that the residue classes modulo (IY)u of ya1,z,…,yaHY(u),z forms a basis of C[y0,…,yl]u/(IY)u and
SY(u,cz)=HY(u)∑i=1ai,z⋅cz. |
Since yai,z∈Yu (modulo (IY)u), we may write
yai,z=Li,z(v1,…,vHY(u)), |
where Li(1≤i≤HY(u)) are independent linear forms. We see at once that
logHY(u)∏i=1|Li,z(˜F(z))|=logHY(u)∏i=1∏1≤t≤n00≤j≤k|Ptj(˜f(z))|ai,t,j,z=−SY(u,cz)+duHY(u)log‖˜f(z)‖+O(uHY(u)). |
This implies that
logHY(u)∏i=1‖˜F(z)‖⋅‖Li,z‖|Li,z(˜F(z))|=SY(u,cz)−duHY(u)log‖˜f(z)‖+HY(u)log‖˜F(z)‖+O(uHY(u)). |
Here, we note that Li,z depends on i and z, but the number of these linear forms is finite. We denote by L the set of all Li,z occurring in the above equalities, then we have
SY(u,cz)≤maxJ⊂Llog∏L∈J‖˜F(z)‖⋅‖L‖|L(˜F(z))|+duHY(u)log‖˜f(z)‖−HY(u)log‖˜F(z)‖+O(uHY(u)), | (3.2) |
where the maximum is taken over all subsets J⊂L with ♯J=HY(u) and where {L;L∈J} is linearly independent. From Lemma 5, we have
SY(u,cz)≥uHY(u)(k+1)δeY(cz)−(2k+1)δHY(u)(max1≤j≤k+11≤i≤n0ci,j,z). | (3.3) |
We choose an index i0 such that z∈S(i0). Since Pi0,1,…,Pi0,k+1 are in general with respect to V, by Lemma 6, we have
eY(cz)≥(ci0,0,z+⋯+ci0,k,z)⋅δ=(log∏0≤j≤k‖˜f(z)‖d‖Pi0,j‖|Pi0,j(˜f)(z)|)⋅δ | (3.4) |
By combining (3.2), (3.3), and (3.4), we get
log∏0≤j≤k‖˜f(z)‖d‖Pi0,j‖|Pi0,j(˜f)(z)|≤k+1uHY(u)(maxJ⊂Llog∏L∈J‖˜F(z)‖⋅‖L‖|L(˜F(z))|−HY(u)log‖˜F(z)‖)+d(k+1)log‖˜f(z)‖+(2k+1)(k+1)δu(max1≤j≤k+11≤i≤n0ci,j,z). | (3.5) |
From (3.1) and (3.5), we have
1ΔVlogq∏i=1‖˜f(z)‖d|Di(˜f)(z)|≤k+1uHY(u)(maxJ⊂Llog∏L∈J‖˜F(z)‖⋅‖L‖|L(˜F(z))|−HY(u)log‖˜F(z)‖)+d(k+1)log‖˜f(z)‖+(2k+1)(k+1)δu∑0≤j≤k1≤i≤n0log‖˜f(z)‖d‖Pi,j‖|Pi,j(˜f)(z)|+O(1), |
where the term O(1) does not depend on z. Integrating both sides of the above inequality, we then obtain
1ΔVq∑i=1mf(r,Di)≤k+1uHY(u)∫2π0maxJ⊂Llog∏L∈J||˜F(reiθ)‖⋅‖L‖|L(˜F(reiθ))|dθ2π+k+1uHY(u)∫2π0maxJ⊂Llog∏L∈J||˜F(r−1eiθ)‖⋅‖L‖|L(˜F(r−1eiθ))|dθ2π−k+1uTF(r)+d(k+1)Tf(r)+(2k+1)(k+1)δu∑0≤j≤k1≤i≤n0mf(r,Pi,j). | (3.6) |
Applying Lemma 1 with ϵ′>0 (which will be chosen later) to the holomorphic curve F and linear forms Li(1≤i≤HY(u)), we obtain that
‖∫2π0maxJ⊂Llog∏L∈J||˜F(reiθ)‖⋅‖L‖|L(˜F(reiθ))|dθ2π+∫2π0maxJ⊂Llog∏L∈J||˜F(r−1eiθ)‖⋅‖L‖|L(˜F(r−1eiθ))|dθ2π≤HY(u)TF(r)−NW(˜F)(r)+OF(r). |
Combining this inequality with (3.6), we have
‖(q−ΔV(k+1))Tf(r)≤q∑i=11dNf(r,Di)+ΔV(2k+1)(k+1)δdu∑0≤j≤k1≤i≤n0mf(r,Pi,j)−ΔV(k+1)duHY(u)NW(˜F)(r)+Of(r). | (3.7) |
We now estimate the quantity NW(˜F)(r). In this case, we define
c=(c1,0,…,c1,k,c2,0,…,c2,k,…,cn0,0,…,cn0,k)∈Zl+1≥0, |
where ci,j:=max{νPi,j(f)(z)−nu,0} for i=1,…,n0 and j=0,…,k. By the definition of the Hilbert weight, there are a1,…,aHY(u)∈Zl+1≥0 with
ai=(ai,1,0,…,ai,1,k,…,ai,n0,0,…,ai,n0,k),ai,j,s∈{1,…,mu+1}, |
such that the residue classes modulo (IY)u of ya1,…,yaHY(u) forms a basis of C[y0,…,yl]u/(IY)u and
SY(u,c)=HY(u)∑i=1ai⋅c. |
Again, there exist independent linear forms ^Li(1≤i≤HY(u)) such that
yai=^Li(v1,…,vHY(u))(1≤i≤HY(u)). |
We also easily see that
max{ν^Li(F)(z)−nu,0}≥∑0≤s≤k∑1≤j≤n0ai,j,smax{νPj,s(f)(z)−nu,0}=ai⋅c |
and hence
νW(˜F)(z)≥HY(u)∑i=1max{ν^Li(F)(z)−nu,0}≥HY(u)∑i=1ai⋅c=SY(u,c). | (3.8) |
For the fixed point z∈△(R), without lose of generality, we may assume that
νD1(f)(z)≥⋯≥νDq(f)(z) |
and σ1=(1,2,…,q).Since P1,0,…,P1,k are in general position with respect to V, by Lemma 6 we have
eY(c)≥δ⋅k∑j=0c1,j=δ⋅k∑j=0max{νP1,j(f)(z)−nu,0}. |
This, together with Lemma 2, gives that
SY(u,c)≥uHY(u)k+1k∑j=0max{νP1,j(f)(z)−nu,0}−(2k+1)δHY(u)max1≤i≤n00≤j≤kνPi,j(f)(z). | (3.9) |
Note the definition of P1,j(0≤j≤k). We then have
ΔVk∑j=0max{νP1,j(f)(z)−nu,0}≥ΔVk∑j=0max{νDtj(f)(z)−nu,0}≥k∑j=0(t1,j−t1,j−1)max{νDtj(f)(z)−nu,0}≥p1∑i=1max{νDi(f)(z)−nu,0}=q∑i=1max{νDi(f)(z)−nu,0}. |
Again, we set t1,−1=0. Thus, we derive from (3.9) that
SY(u,c)≥uHY(u)ΔV(k+1)q∑i=1max{νDi(f)(z)−nu,0}−(2k+1)δHY(u)max1≤i≤n00≤j≤kνPi,j(f)(z). | (3.10) |
Therefore, we derive from (3.8) and (3.10) that
νW(˜F)(z)≥uHY(u)ΔV(k+1)q∑i=1max{νDi(f)(z)−nu,0}−(2k+1)δHY(u)max1≤i≤n00≤j≤kνPi,j(f)(z)≥uHY(u)ΔV(k+1)q∑i=1(νDi(f)(z)−min{νDi(f)(z),nu})−(2k+1)δHY(u)max1≤i≤n00≤j≤kνPi,j(f)(z). | (3.11) |
Integrating both sides of this inequality, we obtain
NW(˜F)(r)≥uHY(u)ΔV(k+1)q∑i=1(Nf(r,Di)−N[nu]f(r,Di))−(2k+1)δHY(u)max1≤i≤n00≤j≤kNf(r,Pi,j). | (3.12) |
Combining inequalities (3.7) and (3.11), we get
‖(q−ΔV(k+1))Tf(r)≤q∑i=11dN[nu]f(r,Di)+ΔV(2k+1)(k+1)δdu∑0≤j≤k1≤i≤n0(mf(r,Pi,j)+Nf(r,Pi,j))+Of(r). | (3.13) |
For each ε>0, we now choose u as the biggest integer such that
u>ΔV(2k+1)(k+1)2n0δε. | (3.14) |
From (3.13) we have
‖(q−ΔV(k+1)−ε)Tf(r)≤q∑i=11dN[nu]f(r,Di)+Of(r). |
Note that degY=δ≤dkdeg(V),
nu=HY(u)−1≤δ(k+uk)≤dkdeg(V)ek(1+uk)k<dkdeg(V)ek(ΔV(2k+4)δlε−1)k≤⌊deg(V)k+1ekdk2+kΔkV(2k+4)klkε−k⌋=Mε. |
Thus, it follows from (3.15) that
‖(q−ΔV(k+1)−ε)Tf(r)≤q∑i=11dN[Mε]f(r,Di)+Of(r). |
The proof of the theorem is finally completed.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This research was supported by Natural Science Research Project for Colleges and Universities of Anhui Province(Nos. 2022AH050329, 2022AH050290).
The authors declare there is no conflicts of interest.
[1] |
R. Nevanlinna, Zur theorie der meromorphen funktionen, Acta. Math., 46 (1925), 1–99. https://doi.org/10.1007/BF02543858 doi: 10.1007/BF02543858
![]() |
[2] | H. Cartan, Sur les zeros des combinaisions linearires de p fonctions holomorpes donnees, Mathematica, 7 (1933), 80–103. |
[3] | E. I. Nochka, On the theory of meromorphic curves, Dokl. Akad. Nauk SSSR, 269 (1983), 547–552. |
[4] |
M. Ru, A defect relation for holomorphic curves intersecting hypersurfaces, Amer. J. Math., 126 (2004), 215–226. https://doi.org/10.1353/ajm.2004.0006 doi: 10.1353/ajm.2004.0006
![]() |
[5] |
M. Ru, Holomorphic curves into algebraic varieties, Ann. Math., 169 (2009), 255–267. https://doi.org/10.4007/annals.2009.169.255 doi: 10.4007/annals.2009.169.255
![]() |
[6] |
S. Axler, Harmomic functions from a complex analysis viewpoit, Amer. Math. Monthly, 93 (1986), 246–258. https://doi.org/10.1080/00029890.1986.11971799 doi: 10.1080/00029890.1986.11971799
![]() |
[7] | A. Y. Khrystiyanyn, A. A. Kondratyuk, On the Nevanlinna Theory for meromorphic functions on annuli Ⅰ, Mathematychni Studii, 23 (2005), 19–30. |
[8] | A. Y. Khrystiyanyn, A. A. Kondratyuk, On the Nevanlinna Theory for meromorphic functions on annuli Ⅱ, Mathematychni Studii, 24 (2005), 57–68. |
[9] |
H. T. Phuong, N. V. Thin, On fundamental theorems for holomorphic curves on the annuli, Ukra. Math. J., 67 (2015), 1111–1125. https://doi.org/10.1007/s11253-015-1138-5 doi: 10.1007/s11253-015-1138-5
![]() |
[10] | J. L. Chen, T. B. Cao, Second main theorem for holomorphic curves on annuli, Math. Rep., 22 (2020), 261–275. |
[11] |
S. D. Quang, Meromorphic mappings into projective varieties with arbitrary families of moving hypersurfaces, J. Geom. Anal., 32 (2022), Article number: 52. https://doi.org/10.1007/s12220-021-00765-3 doi: 10.1007/s12220-021-00765-3
![]() |
[12] |
S. D. Quang, Generalizations of degeneracy second main theorem and Schmidts subspace theorem, Pacific J. Math., 318 (2022), 153–188. https://doi.org/10.2140/pjm.2022.318.153 doi: 10.2140/pjm.2022.318.153
![]() |
[13] | S. D. Quang, Second main theorem for holomorphic maps from complex disks with moving hyperplanes and application, Complex Var Elliptic Equ., 65 (2023), 150–156. |
[14] | M. Ru, Nevanlinna Theory and its Relation to Diophantine Approximation, 2nd edition, World Scientific, Singapore, 2021. https://doi.org/10.1142/12188 |
[15] |
S. D. Quang, Degeneracy second main theorems for meromorphic mappings into projective varieties with hypersurfaces, Trans. Amer. Math. Soc., 371(4) (2019), 2431–2453. https://doi.org/10.1090/tran/7433 doi: 10.1090/tran/7433
![]() |
[16] |
S. D. Quang, D. P. An, Second main theorem and unicity of meromorphic mappings for hypersurfaces in projective varieties, Acta Math Vietnam, 42 (2017), 455–470. https://doi.org/10.1007/s40306-016-0196-6 doi: 10.1007/s40306-016-0196-6
![]() |
[17] | P. Griffiths, J. Harris, Principles of Algebraic Geometry, Wiley, New York, 1994. https://doi.org/10.1002/9781118032527 |
[18] |
J. H. Evertse, R. G. Ferretti, Diophantine inequalities on projective varieties, Internat. Math. Res. Notices, 25 (2002), 1295–1330. https://doi.org/10.1155/S107379280210804X doi: 10.1155/S107379280210804X
![]() |