Loading [MathJax]/jax/output/SVG/jax.js
Research article

Second main theorem for holomorphic curves on annuli with arbitrary families of hypersurfaces

  • Received: 02 October 2023 Revised: 23 December 2023 Accepted: 09 January 2024 Published: 01 February 2024
  • 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

    Related Papers:

    [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},0r<R+. We need only consider two cases: r=0,R=+ simultaneously and 0<r<R<+. In the latter case the homothety zzrR 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 z1z. 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={zC:1R0<|z|<R0}

    be an annuli in C. Moreover, for any real number r such that 1<r<R0, we denote

    Ar={zC:1r<|z|<r},
    A1,r={zC:1r<|z|1},

    and

    A2,r={zC: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)=12π2π0logf(reiθ)dθ+12π2π0logf(r1eiθ)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):=12π2π0logf(reiθ)d|Qf(reiθ)|dθ+12π2π0logf(r1eiθ)d|Qf(r1eiθ)|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 Qf in Aj,r, counting multiplicity, and let nMj,f(r,D) be the number of zeros of Qf 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):=1r1n1,f(t,D)tdt,N2,f(r,D):=r1n2,f(t,D)tdt,N[M]1,f(r,D):=1r1nM1,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(log1R0r+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:APn(C) be a linearly nondegenerate holomorphic curve. Let H1,,Hq be hyperplanes in Pn(C), located in general position, then we have

    ||(qn1)Tf(r)qj=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 Ar with Arrλ1dr<+. At the same time, if R0<+, then the assertion P holds for all r(1,R0) outside a set Ar with Ar1(R0r)λ+1dr<+, where λ0.

    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:=dimVdim(VS).

    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(jQ). We define the distributive constant of {Dj}qj=1 with respect to V by

    ΔV:=maxΓQΓcodimV(jΓDj).

    Definition 2. Let {Dj}qj=1 be the hypersurfaces in Pn(C). Denote by Q the index set {1,,q}. Let Nn and qN+1. The family {Dj}qj=1 is said to be in N-subgeneral position with respect to V if for every subset RQ with the cardinality R=N+1, then

    jRDjV=.

    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 ΔVNdimV+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:AV be an algebraically nondegenerate holomorphic curve with 0<R0+. Let {Dj}qj=1 be q hypersurfaces in Pn(C) with degDj=dj(1jq). Let d be the least common multiple of djs, 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,

    ||(qΔV(k+1)ε)Tf(r)qj=11djN[Mε]f(r,Dj)+Of(r),

    where

    Mε=deg(V)k+1ekdk2+kΔkV(2k+4)klkεk

    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:AV 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(1jq). Let d be the least common multiple of djs, i.e., d=lcm(d1,,dq). Then for any ε>0,

    ||(q(Nk+1)(k+1)ε)Tf(r)qj=11djN[˜Mε]f(r,Dj)+Of(r).

    where

    ˜Mε=deg(V)k+1ekdk2+k(Nk+1)k(2k+4)klkεk

    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:APn(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

    2π0maxKjKlogf(reiθ)f(reiθ);Hjdθ2π+2π0maxKjKlogf(r1eiθ)f(r1eiθ);Hjdθ2π(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,jK, 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 d1, such that pi=0DiV= and

    dim(si=0Di)V=kl,tl1s<tl,1lk

    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=tlj=0cljDj,cljC,l=0,,k

    such that (kl=0Pl)V=.

    Lemma 3. ([16]) Let {Qi}iR 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 iRQiV=, then there exist positive constants α and β such that

    αfdmaxiR|Qi(f)|βfd.

    Lemma 4. ([11]) Let t0,t1,,tn be n+1 integers such that 1=t0<t1<<tn, and let Δ=max1sntst0s, then for every n real numbers a0,a1,,an1 with a0a1an11, we have

    at1t00at2t11atntn1n1(a0a1an1)Δ.

    We recall the notion of Chow weights and Hilbert weights from [5] (see also [17]). Let XPn(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, XHu0Huk, 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,,GrC[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)(1ik) 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)=tjJcj[J](u00,,u0n,,uk0,,ukn)

    For a=(a0,,an)Zn+1 we write xa for the monomial xa00xann. 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]uI.

    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)=δunn!+O(un1).

    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=1aic)

    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 XPn(C) be an algebraic variety of dimension k and degree δ. Let u>δ be an integer and let c=(c0,,cn)Rn+10, then

    1uHX(u)SX(u,c)1(k+1)δeX(c)(2k+1)δu(maxi=0,,nci).

    Lemma 6. Let YPn(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 i0Q={1,2,,q} such that jQ{i0}DjV, then it follows from the definition that

    ΔVq1kdim(jQ{i0}DjV)q1k>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 iQ, the set jQ{i}DjV=.

    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}iI the set of all permutations of the set {1,,q}, where I={1,2,,n0} and n0=q!. For each iI, since q1j=1Dσi(j)V=, there exist k+1 integers ti,0,ti,1,,ti,k with 1=ti,0<<ti,k=pi, where piq1 such that pij=1Dσi(j)V= and

    dim(sj=1Dσi(j)V)=klti,l1s<ti,l,1lk.

    For each iI, 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 xV 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:=degYdkdegV.

    For every

    a=(a1,0,,a1,k,a2,0,,a2,k,,an0,0,,an0,k)Zl+10

    and

    y=(y1,0,,y1,k,y2,0,,y2,k,,yn0,0,,yn0,k)

    we denote ya=ya1,01,0ya1,k1,kyan0,0n0,0yan0,kn0,k. Let u be a positive integer. We set

    nu:=HY(u)1,mu:=(m+u1u)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)):ACnu+1.

    Since f is algebraically nondegenerate, the holomorphic curve F:APnu(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 iI, such that

    ˜f(z)dAmax0jpi|Dσi(j)(˜f(z))|.

    According to the definition of Pi,j, we may choose a positive constant B1, commonly for all iI, such that

    |Pi,j(x)|Bmax1sti,j|Dσi(s)(x)|,

    for all x=(x0,,xn)Cn+1 and for all 0jk. It is easily seen that, there exists a positive constant C, such that

    |Pi,j(x)|Cxd

    for all x=(x0,,xn)Cn+1,1in0, and 0jk.

    Fix an element iI. Denote by S(i) the set of all points

    z(R){qi=1Di(˜f(z))1({0})0jkiIPi,j(˜f(z))1({0})}

    such that

    |Dσi(1)(˜f(z))||Dσi(2)(˜f(z))||Dσi(q)(˜f(z))|.

    Therefore, for each zS(i), by Lemma 4 we have

    qi=1˜f(z)d|Di(˜f(z))|Aqpipij=1˜f(z)d|Dσi(j)(˜f(z))|Aqpikj=1(˜f(z)d|Dσi(tj1)(˜f(z))|)ti,jti,j1Aqpikj=1(˜f(z)d|Dσi(tj1)(˜f(z))|)ΔVAqpiBkΔVk1j=0(˜f(z)d|Pi,j(˜f(z))|)ΔVAqpiBkΔVCΔVkj=0(˜f(z)d|Pi,j(˜f(z))|)ΔV,

    which implies that

    logqi=1˜f(z)d|Di(˜f(z))|log(AqpiBkΔVCΔV)+ΔVlogkj=0(˜f(z)d|Pi,j(˜f(z))|). (3.1)

    Now, we fix an index iI and a point zS(i) and define

    cz=(c1,0,z,,c1,k,z,c2,0,z,,c2,k,z,,cn0,0,z,,cn0,k,z)Rl+10,

    where ci,j,z:=log˜f(z)dPi,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+10 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,zcz.

    Since yai,zYu (modulo (IY)u), we may write

    yai,z=Li,z(v1,,vHY(u)),

    where Li(1iHY(u)) are independent linear forms. We see at once that

    logHY(u)i=1|Li,z(˜F(z))|=logHY(u)i=11tn00jk|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)maxJLlogLJ˜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 JL with J=HY(u) and where {L;LJ} is linearly independent. From Lemma 5, we have

    SY(u,cz)uHY(u)(k+1)δeY(cz)(2k+1)δHY(u)(max1jk+11in0ci,j,z). (3.3)

    We choose an index i0 such that zS(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=(log0jk˜f(z)dPi0,j|Pi0,j(˜f)(z)|)δ (3.4)

    By combining (3.2), (3.3), and (3.4), we get

    log0jk˜f(z)dPi0,j|Pi0,j(˜f)(z)|k+1uHY(u)(maxJLlogLJ˜F(z)L|L(˜F(z))|HY(u)log˜F(z))+d(k+1)log˜f(z)+(2k+1)(k+1)δu(max1jk+11in0ci,j,z). (3.5)

    From (3.1) and (3.5), we have

    1ΔVlogqi=1˜f(z)d|Di(˜f)(z)|k+1uHY(u)(maxJLlogLJ˜F(z)L|L(˜F(z))|HY(u)log˜F(z))+d(k+1)log˜f(z)+(2k+1)(k+1)δu0jk1in0log˜f(z)dPi,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

    1ΔVqi=1mf(r,Di)k+1uHY(u)2π0maxJLlogLJ||˜F(reiθ)L|L(˜F(reiθ))|dθ2π+k+1uHY(u)2π0maxJLlogLJ||˜F(r1eiθ)L|L(˜F(r1eiθ))|dθ2πk+1uTF(r)+d(k+1)Tf(r)+(2k+1)(k+1)δu0jk1in0mf(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(1iHY(u)), we obtain that

    2π0maxJLlogLJ||˜F(reiθ)L|L(˜F(reiθ))|dθ2π+2π0maxJLlogLJ||˜F(r1eiθ)L|L(˜F(r1eiθ))|dθ2πHY(u)TF(r)NW(˜F)(r)+OF(r).

    Combining this inequality with (3.6), we have

    (qΔV(k+1))Tf(r)qi=11dNf(r,Di)+ΔV(2k+1)(k+1)δdu0jk1in0mf(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+10,

    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+10 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=1aic.

    Again, there exist independent linear forms ^Li(1iHY(u)) such that

    yai=^Li(v1,,vHY(u))(1iHY(u)).

    We also easily see that

    max{ν^Li(F)(z)nu,0}0sk1jn0ai,j,smax{νPj,s(f)(z)nu,0}=aic

    and hence

    νW(˜F)(z)HY(u)i=1max{ν^Li(F)(z)nu,0}HY(u)i=1aic=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)δkj=0c1,j=δkj=0max{νP1,j(f)(z)nu,0}.

    This, together with Lemma 2, gives that

    SY(u,c)uHY(u)k+1kj=0max{νP1,j(f)(z)nu,0}(2k+1)δHY(u)max1in00jkνPi,j(f)(z). (3.9)

    Note the definition of P1,j(0jk). We then have

    ΔVkj=0max{νP1,j(f)(z)nu,0}ΔVkj=0max{νDtj(f)(z)nu,0}kj=0(t1,jt1,j1)max{νDtj(f)(z)nu,0}p1i=1max{νDi(f)(z)nu,0}=qi=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)qi=1max{νDi(f)(z)nu,0}(2k+1)δHY(u)max1in00jkνPi,j(f)(z). (3.10)

    Therefore, we derive from (3.8) and (3.10) that

    νW(˜F)(z)uHY(u)ΔV(k+1)qi=1max{νDi(f)(z)nu,0}(2k+1)δHY(u)max1in00jkνPi,j(f)(z)uHY(u)ΔV(k+1)qi=1(νDi(f)(z)min{νDi(f)(z),nu})(2k+1)δHY(u)max1in00jkνPi,j(f)(z). (3.11)

    Integrating both sides of this inequality, we obtain

    NW(˜F)(r)uHY(u)ΔV(k+1)qi=1(Nf(r,Di)N[nu]f(r,Di))(2k+1)δHY(u)max1in00jkNf(r,Pi,j). (3.12)

    Combining inequalities (3.7) and (3.11), we get

    (qΔV(k+1))Tf(r)qi=11dN[nu]f(r,Di)+ΔV(2k+1)(k+1)δdu0jk1in0(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

    (qΔV(k+1)ε)Tf(r)qi=11dN[nu]f(r,Di)+Of(r).

    Note that degY=δdkdeg(V),

    nu=HY(u)1δ(k+uk)dkdeg(V)ek(1+uk)k<dkdeg(V)ek(ΔV(2k+4)δlε1)kdeg(V)k+1ekdk2+kΔkV(2k+4)klkεk=Mε.

    Thus, it follows from (3.15) that

    (qΔV(k+1)ε)Tf(r)qi=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
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(906) PDF downloads(48) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog