1. Introduction
Let $\Delta(R)=\{z\in{\mathbb{C}}: |z| \lt R\}$ be a complex disc and r 0 be a fixed positive number so that $0 \lt r_0 \lt R$. Let ν be a divisor on $\Delta (R)$, which is regarded as a function on $\Delta (R)$ with values in $\mathbb{Z}$ such that $\mathrm{Supp}\,\nu:=\{z;\nu(z)\ne 0\}$ is a discrete subset of $\Delta (R)$. Let k be a positive integer or $+\infty$. The truncated counting function of ν is defined by:
We will omit the character $^{[k]}$ if $k=+\infty$.
Let $\varphi:\Delta (R)\rightarrow {\mathbb{C}}\cup\{\infty\}$ be a non-constant meromorphic function. We denote by $\nu^0_\varphi$ (resp. $\nu^\infty_\varphi$) the divisor of zeros (resp. divisor of poles) of φ and set $\nu_\varphi=\nu^0_\varphi-\nu^\infty_\varphi$. As usual, we will write $N^{[k]}_{\varphi}(r)$ and $N^{[k]}_{1/\varphi}(r)$ for $N^{[k]}(r,\nu^0_{\varphi})$ and $N^{[k]}(r,\nu^\infty_{\varphi})$ respectively.
Let $f:\Delta (R)\rightarrow{\mathbb{P}}^N({\mathbb{C}})$ be a holomorphic map and Ω be the Fubini-Study form on ${\mathbb{P}}^N({\mathbb{C}})$. The characteristic function of f is defined by
In [Reference Ru and Sibony8], Ru and Sibony defined the growth index of f by
For the convenient, we will set $c_f=+\infty$ if
A meromorphic function a on $\Delta (R)$ (which is regarded as a holomorphic map into ${\mathbb{P}}^1({\mathbb{C}})$) is said to be small with respect to f if $\| T_a(r)=o(T_f(r))$ as $r\rightarrow R$. Here (and throughout this paper), the notation ‘$\| P$’ means the assertion P holds for all $r\in (0,R)$ outside a subset S of $(0,R)$ with $\int_{S}{\rm exp}((c_f+\epsilon)T_f(r))dr \lt +\infty$ for some ϵ > 0.
Denote by $\mathcal H$ the ring of all holomorphic functions on $\Delta (R)$. Let Q be a homogeneous polynomial in $\mathcal H[x_0,\dots,x_n]$ of degree $d \geq 1$ given by
where $\mathcal T_d=\{(i_0,\ldots,i_N)\in\mathbf{N}_0^{N+1}\ ;\ i_0+\cdots +i_N=d\}$, $\omega^I=\omega_0^{i_0}\cdots\omega_n^{i_N}$ for $I=(i_0,\ldots,i_N)$ and all $a_I\in\mathcal H$ has no common zero. The homogeneous polynomial Q is called a moving hypersurface of ${\mathbb{P}}^N (\mathbf{C} )$. Throughout this paper, by changing the homogeneous coordinates of ${\mathbb{P}}^N({\mathbb{C}})$ if necessary, we may assume that $a_{I_0}\not\equiv 0$ each such given moving hypersurface Q, where $I_0=(d,0,\ldots,0)$. We put $\tilde Q=\sum_{I\in\mathcal T_d}\frac{a_{I}}{a_{I_0}}\omega^I$.
The moving hypersurface Q is said to be slow with respect to f if all $\frac{a_I}{a_{I_0}}\ (I\in\mathcal T_d)$ are small with respect to f. Let $\mathbf f=(f_0,\dots,f_N)$ be a reduced representation of f. We define
Then the truncated divisor $\nu^{[k]}_{Q({\bf{f}})}$ does not depend on the choice of the reduced representation ${\bf{f}}$ and hence is written by $\nu_{Q(f)}$. Its truncated counting function is denoted simply by $N^{[k]}_{Q(f)}(r)$. The proximity function of f with respect to Q is define by
If Q is slow with respect to f, then the first main theorem states that
The truncated defect of f with respect to Q is defined by
We omit the character $^{[k]}$ if $k=+\infty$. If all coefficients of Q are constant then we call Q a (fixed) hypersurface of ${\mathbb{P}}^N({\mathbb{C}})$ and set $Q^\ast=\{(\omega_0:\cdots:\omega_N)\;\vert\;\sum_{I\in{\mathcal T}_d}a_I\omega^I=0\}$.
Let V be a smooth projective subvariety of ${\mathbb{P}}^N({\mathbb{C}})$ of dimension n. Let $\mathcal Q=\{Q_1,\ldots,Q_q\}$ be a family of moving hypersurfaces in ${\mathbb{P}}^N(\mathbf{C})$, where $Q_i=\sum_{I\in\mathcal T_{d_i}}a_{iI}x^I$. Denote by $\mathcal K_{\mathcal Q}$ the smallest field which contains ${\mathbb{C}}$ and all functions $\frac{a_{iI}}{a_{iI_0}}\ (I\in\mathcal T_{d_i})$. As usual, the family $\{Q_1,\ldots,Q_q\}$ is said to be in weakly $\ell$-subgeneral position if $\bigcap_{s=1}^{\ell+1}Q_{j_s}(z)^*\cap V=\emptyset$ for every $1\le j_1 \lt \cdots \lt j_{\ell+1}\le q$ and for generic points $z\in\Delta(R)$ (i.e. for all $z\in\Delta(R)$ outside a discrete subset). Here, we note that $\dim\emptyset=-\infty$. In [Reference Quang6], we define the distributive constant of the family $\mathcal Q$ with respect to V by
for generic points $z\in\Delta (R)$. From [Reference Quang6, Remark 3.7], we know that if $\mathcal Q$ is in weakly $\ell$-subgeneral position with respect to V then $\Delta_{V}\le\ell-n+1$.
For the case of holomorphic curves from ${\mathbb{C}}$ into V and families of fixed hypersurfaces in general position (i.e. in n-subgeneral position), Ru [Reference Ru7] proved the following.
Theorem A. (see [Reference Ru7])
Let f be an algebraically non-degenerate holomorphic map of ${\mathbb{C}}$ into a smooth subvariety $V\subset{\mathbb{P}}^N(\mathbf{C})$ of dimension n. Let $\{Q_i\}_{i=1}^q$ be a family of q hypersurfaces in general position with respective to V. Then for any ϵ > 0, we have
From the above theorem, the number n + 1 is an above bound of the sum of defects (without truncated multiplicity) for hypersufaces in this case. Later on, many mathematicians generalized Theorem A to the case of slowly moving hypersurfaces in general position with respect to V. Recently, in [Reference Quang4] Quang considered the case of holomorphic maps from $\Delta(R)$ into ${\mathbb{P}}^N({\mathbb{C}})$ and proved the following result.
Theorem B. (reformulation of [Reference Quang4, Theorem 1.3]) Let $V\subset{\mathbb{P}}^N({\mathbb{C}})$ be a smooth complex projective variety of dimension $n\ge 1$. Let $\{Q_1,\ldots,Q_q\}$ be a family of hypersurfaces in ${\mathbb{P}}^N({\mathbb{C}})$ with the distributive constant Δ with respect to V, $\deg Q_i=d_i\ (1\le i\le q)$, and let d be the least common multiple of $d_1,\ldots,d_q$. Let $f:\Delta(R)\to V$ be an algebraically non-degenerate holomorphic curve with $c_f \lt +\infty$. Then, for every ϵ > 0,
where $L=\bigl[d^{n^2+n}\deg (V)^{n+1}e^n\Delta^n(2n+4)^n(n+1)^n(q!)^n\epsilon^{-n}\bigl].$
Here, by $[x]$ we denote the largest integer not exceeding the real number x. However, the coefficient of $T_f(r)$ in the right hand side of the above inequality has a factor $c_f(q!\epsilon^{-1})^{n-1}$. Then if cf or q is large enough, this coefficient always exceeds q and the theorem is meaningless. Hence, it may not imply the defect relation. Our purpose in this paper is to improve Theorem B by reducing the truncation level L and that coefficients so that they do not depend on q. In order to do so, we will apply the new below bound of Chow weight in [Reference Quang5] (see Lemma 2.2) and also give some new technique to control the error term occuring when the theorem on the estimate of Hilbert weights (Theorem 2.1) is applied. We also consider the case of moving hypersurfaces. Our main result is stated as follows.
Theorem 1.1. Let f be a non-constant holomorphic map of $\Delta (R)$ into an n-dimensional smooth projective subvariety $V\subset{\mathbb{P}}^N(\mathbf{C})$ with finite growth index cf. Let $\{Q_i\}_{i=1}^q$ be a family of slow (with respect to f) moving hypersurfaces with the distributive constant $\Delta_{V}$ with respect to V. Assume that f is algebraically non-degenerate over $\mathcal K_{\mathcal Q}$.
a) For any $\epsilon' \gt 0$ and $(n+1)\Delta_V \gt \epsilon \gt 0$, we have
\begin{align*} \bigl\|\ &(q-\Delta_{V}(n+1)-\epsilon)T_f(r)\\ &\le\sum_{j=1}^{q}\frac{1}{d}N^{[L-1]}_{Q_j(f)}(r)+\frac{(\Delta_{V}(n+1)+\epsilon)(c_f+\epsilon')(L-1)}{2du}T_f (r), \end{align*}where
\begin{equation*}L=d^n\deg V(u+1)^n\left[\biggl(1+\frac{\epsilon}{2(n+1)\Delta_{V}}\biggl)^{\left[\frac{d^n\deg V(u+1)^{n+q}}{\log^2(1+\frac{\epsilon}{2(n+1)\Delta_{V}})}\right]+1}\right]\end{equation*}with $u=\lceil 2\Delta_{V}(2n+1)(n+1)d^n\deg V(\Delta_{V}(n+1)+\epsilon)\epsilon^{-1}\rceil$.
b) Assume further that all $Q_i\ (1\le i\le q)$ are assumed to be fixed hypersurfaces. For any $\epsilon' \gt 0$ and ϵ > 0 we have
\begin{align*} \bigl\|\ &(q-\Delta_{V}(n+1)-\epsilon)T_f(r)\\ &\le\sum_{j=1}^{q}\frac{1}{d}N^{[L'-1]}_{Q_j(f)}(r)+\frac{(\Delta_{V}(n+1)+\epsilon)(c_f+\epsilon')(L'-1)}{2du'}T_f (r), \end{align*}where $L'=\bigl[d^{n^2+n}(\deg V)^{n+1}e^n(2n+5)^n(\Delta^2_V(n+1)\epsilon^{-1}+\Delta_V)^n\bigl]$ with $u'=\lceil \Delta_{V}(2n+1)(n+1)d^n\deg V(\Delta_{V}(n+1)+\epsilon)\epsilon^{-1}\rceil$.
Here, $\lceil x\rceil$ stands for the smallest integer bigger than or equal to the real number x. By this theorem, we get the following truncated defect relation for fixed hypersurfaces.
Corollary 1.2. With the assumption and notation as in Theorem 1.1 and suppose that all Qi are fixed hypersurfaces. Then for any ϵ > 0, we have
where $L=[d^{n^2+n}(\deg V)^{n+1}e^n(2n+5)^n(\Delta^2_V(n+1)\epsilon^{-1}+\Delta_V)^n]$ with $u=\lceil 2\Delta_{V}(2n+1)(n+1)d^n\deg V(\Delta_{V}(n+1)+\epsilon)\epsilon^{-1}\rceil$.
2. Some auxiliary results
Let $X\subset{\mathbb{P}}^n({\mathbb{C}})$ be a projective variety of dimension k and degree δ. For $\textbf{a} = (a_0,\ldots,a_n)\in\mathbb Z^{n+1}_{\ge 0}$ we write $\textbf{x}^\textbf{a}$ for the monomial $x^{a_0}_0\cdots x^{a_n}_n$. Let IX be the prime ideal in ${\mathbb{C}}[x_0,\ldots,x_n]$ defining X. Let ${\mathbb{C}}[x_0,\ldots,x_n]_u$ be the vector space of homogeneous polynomials in ${\mathbb{C}}[x_0,\ldots,x_n]$ of degree u (including 0). For $u= 1, 2,\ldots,$ put $(I_X)_u:={\mathbb{C}}[x_0,\ldots,x_n]_u\cap I_X$ and define the Hilbert function HX of X by
Let $\textbf{c}=(c_0,\ldots,c_n)$ be a tuple in $\mathbb R^{n+1}_{\ge 0}$ and let $e_X(\textbf{c})$ be the Chow weight of X with respect to c. The u-th Hilbert weight $S_X(u,\textbf{c})$ of X with respect to c is defined by
where the maximum is taken over all sets of monomials $\textbf{x}^{\textbf{a}_1},\ldots,\textbf{x}^{\textbf{a}_{H_X(u)}}$ whose residue classes modulo IX form a basis of ${\mathbb{C}}[x_0,\ldots,x_n]_u/(I_X)_u.$
The following theorem is due to Evertse and Ferretti [Reference Evertse and Ferretti2].
Theorem 2.1. (see [Reference Evertse and Ferretti2, Theorem 4.1]) Let $X\subset{\mathbb{P}}^n({\mathbb{C}})$ be an algebraic variety of dimension k and degree δ. Let $u \gt \delta$ be an integer and let $\textbf{c}=(c_0,\ldots,c_n)\in\mathbb R^{n+1}_{\geqslant 0}$. Then
The following lemma is due to the Quang [Reference Quang5].
Lemma 2.2. (see [Reference Quang5, Lemma 3.2]) Let Y be a projective subvariety of ${\mathbb{P}}^R({\mathbb{C}})$ of dimension $k\ge 1$ and degree δY. Let $\ell\ (\ell\ge k+1)$ be an integer and let $\textbf{c}=(c_0,\ldots,c_R)$ be a tuple of non-negative reals. Let $\mathcal H=\{H_0,\ldots,H_R\}$ be a set of hyperplanes in ${\mathbb{P}}^R({\mathbb{C}})$ defined by $H_{i}=\{y_{i}=0\}\ (0\le i\le R)$. Let $\{i_1,\ldots, i_\ell\}$ be a subset of $\{0,\ldots,R\}$ such that:
(1) $c_{i_\ell}=\min\{c_{i_1},\ldots,c_{i_\ell}\}$,
(2) $Y\cap\,\bigcap_{j=1}^{\ell-1}H_{i_j}\ne \emptyset$,
(3) and $Y\not\subset H_{i_j}$ for all $j=1,\ldots,\ell$.
Let $\Delta_{\mathcal H,Y}$ be the distributive constant of the family $\mathcal H=\{H_{i_j}\}_{j=1}^\ell$ with respect to Y. Then
The following theorem is due to Ru and Sibony [Reference Ru and Sibony8].
Theorem 2.3. (reformulation of [Reference Ru and Sibony8, Theorem 4.8]) Let f be a linearly non-degenerate holomorphic map from $\Delta (R)\ (0 \lt R\le +\infty)$ into $\mathbb{P}^n(\mathbb{C})$. Let $H_1,\ldots,H_q$ be q arbitrary hyperplanes in ${\mathbb{P}}^n({\mathbb{C}})$. Then, for every ɛ > 0, we have
where $W=\det (f_i^{(k)}; 0\le i,k\le n)$ for a reduced representation $\mathbf f=(f_0,\dots,f_n)$ of f.
Note that, in the original theorem [Reference Ru and Sibony8, Theorem 4.8], the last term of the right hand side of the above inequality is more complex and the inequality holds for all r outside an exceptional set $S\subset(0,R)$ such that $\int_S{\rm exp}((c_f+\epsilon)T_f(r))dr \lt +\infty$. In this reformulation, we just simplify that term but the exceptional set is a subset $S'\subset(0,R)$ such that $\int_{S'}{\rm exp}((c_f+\epsilon')T_f(r))dr \lt +\infty$ for some $\epsilon' \gt 0$.
Let $\mathcal Q=\{Q_1,\ldots,Q_q\}$ be a family moving hypersurfaces in ${\mathbb{P}}^N({\mathbb{C}})$ given by
where $\textbf{x}=(x_0,\ldots,x_N)$, $\textbf{x}^I=x_0^{i_0}\cdots x_N^{i_N}$ for $I=(i_0,\ldots,i_N)$. Denote by $\mathcal C_{\mathcal Q}$ the set of all non-negative functions $h : \mathbf{C}^m\setminus A\longrightarrow [0,+\infty]$, which are of the form:
where $k,l\in\mathbf{N},\ g_1,...., g_{l+k}\in\mathcal K_{\mathcal Q}\setminus\{0\}$ and A is a discrete subset of $\Delta(R)$, which may depend on $g_1,....,g_{l+k}$. Then, for $h\in\mathcal C_{\mathcal Q}$ we have
Also, for every moving hypersurface Q in $\mathcal K_{\mathcal Q}[x_0,\ldots,x_N]$ of degree d, we have
for some $c\in\mathcal C_{\mathcal Q}$.
Lemma 2.4. (see [Reference Quang6, Lemma 3.2]) Let V be a projective variety of ${\mathbb{P}}^N({\mathbb{C}})$. With the above notation, let $1\le j_1\le\cdots\le j_k\le q$. Suppose that there exists $z_0\in\Delta(R)$ such that $V\cap\,\bigcap_{s=1}^kQ_{j_s}(z_0)^*=\emptyset.$ Then we have $V\cap\bigcap_{s=1}^kQ_{j_s}(z)^*=\emptyset$ for every $z\in\Delta(R)$ outside a discrete subset, and there exists a function $c\in\mathcal C_{\mathcal K}$ such that
3. Proof of theorem 1.1
Replacing Qj by $Q^{\frac{d}{d_{j}}}$ if necessary, we may assume that $Q_{1},\ldots,Q_{q}$ have the same degree d and $Q_i=\sum_{I\in\mathcal T_d}a_{iI}x^I\ (i=1,\ldots,q)$. Take a point z 0 such that $a_{iI_0}(z_0)\ne 0$ for all i, and
It is suffice for us to consider the case where $q \gt \Delta_V(n+1)$. Denote by $\sigma_1,\ldots,\sigma_{n_0}$ all bijections from $\{0,\ldots,q-1\}$ into $\{1,\ldots,q\}$, where $n_0=q!$. For each σi, it is easy to see that $\bigcap_{j=0}^{q-2}\tilde Q_{\sigma_i(j)}(z_0)^*\cap V=\emptyset$. Then there exists a smallest index $\ell_i\le q-2$ such that $\bigcap_{j=0}^{\ell_i}\tilde Q_{\sigma_i(j)}(z_0)^*\cap V=\emptyset$. Hence $\bigcap_{j=0}^{\ell_i}\tilde Q_{\sigma_i(j)}(z)^*\cap V=\emptyset$ for generic points z and for all $i=1,\ldots,n_0$. Denote by $\mathcal S$ the set of all points $z\in\Delta(R)$ such that $\bigcap_{j=0}^{\ell_i}\tilde Q_{\sigma_i(j)}(z)^*\cap V\ne \emptyset$ for some i. Then $\mathcal S$ is a discrete subset of $\Delta(R)$.
By Lemma 2.4, there is a function $A\in\mathcal C_{\mathcal Q}$, chosen common for all σi, such that
Denote by S(i) the set of all z not in $\mathcal S$ such that $\tilde Q_j({\bf{f}})(z)\ne 0$ for all $j=1,\ldots,q$ and
Therefore, for every generic point $z\in S(i)$, we have
where $C(z)=\sum_{i=1}^{n_0}A(z)^{q-\ell_i-1}\in\mathcal C_{\mathcal Q}$.
For $z\not\in\mathcal S$, consider the mapping $\Phi_z$ from V into ${\mathbb{P}}^{q-1}({\mathbb{C}})$ defined by
for every $\textbf{x}=(x_0:\cdots:x_N)\in V$, where $x=(x_0,\ldots,x_N)$. We set
Let $Y_z=\Phi_z(V)$. Since $V\cap\bigcap_{j=1}^{q}\tilde Q_j(z)^*=\emptyset$, $\Phi_z$ is a finite morphism on V and Yz is a projective subvariety of ${\mathbb{P}}^{q-1}({\mathbb{C}})$ with $\dim Y_z=n$ and of degree
For every $\textbf{a} = (a_1,\ldots,a_q)\in\mathbb Z^q_{\ge 0}$ and $\textbf{y} = (y_1,\ldots,y_q)$ we denote $\textbf{y}^\textbf{a} = y_{1}^{a_{1}}\ldots y_{q}^{a_{q}}$. Let u be a positive integer. We set $\xi_u:=\binom{q+u}{u}$ and define the ${\mathbb{C}}$-vector space
Denote by $(I_Y)_u$ the subspace of the $\mathcal K_{\mathcal Q}$-vector space $\mathcal K_{\mathcal Q}[y_1,\ldots,y_q]_u$ consisting of all homogeneous polynomials $P\in \mathcal K_{\mathcal Q}[y_1,\ldots,y_q]_u$ (including the zero polynomial) such that
Let $(\tilde R_1,\ldots,\tilde R_p)$ be an $\mathcal K_{\mathcal Q}$-basis of $(I_Y)_u$. By enlarging $\mathcal S$ if necessary, we may assume that all zeros and poles of all non-zero coefficients of $\tilde R_i\ (1\le i\le p)$ are contained in $\mathcal S$, also all above assertions for generic points z still hold for all $z\not\in\mathcal S$. Choose $\xi_u-p$ non-zero monic monomial $v_1,\ldots,v_{\xi_u-p}$ of degree of u in variables $y_1,\ldots,y_q$ such that $\{\tilde R_1,\ldots,\tilde R_p,v_1,\ldots,v_{\xi_u-p}\}$ is a $\mathcal K_{\mathcal Q}$-basis of $\mathcal K_{\mathcal Q}[y_1,\ldots,y_q]_u$.
Denote by $\mathcal T=\{T_1,\ldots,T_{\xi_u}\}$ the set of all non-zero monic monomials of degree of u in variables $y_1,\ldots,y_q$. Then $\{T_1,\ldots,T_{\xi_u}\}$ is a $\mathcal K_{\mathcal Q}$-basis of $\mathcal K_{\mathcal Q}[y_1,\ldots,y_q]_u$, and also is an ${\mathbb{C}}$-basis of ${\mathbb{C}}[y_1,\ldots,y_q]_u$.
From [Reference Quang6, Claim 4.3], we have the following claim.
Claim 3.1.
There is a discrete subset $\mathcal S'$ of $\Delta (R)$ such that for all $z\not\in\mathcal S'$, we have:
(i) the family of equivalent classes of $v_1,\ldots,v_{\xi_u-p}$ is a basis of $Y_{z,u}$ and the family $\{\tilde R_1(z),\ldots,\tilde R_p(z)\}$ is a ${\mathbb{C}}$-basis of $(I_{Y_z})_u$;
(ii) for a subset $\{v_1',\ldots,v'_{\xi_u-p}\}$ of $\mathcal T$, if $\{\tilde R_1,\ldots,\tilde R_p,v_1',\ldots,v'_{\xi_u-p}\}$ is a $\mathcal K_{\mathcal Q}$-basis of $\mathcal K_{\mathcal Q}[y_1,\ldots,y_q]_u$ then the set of equivalent classes of $v'_1,\ldots,v'_{\xi_u-p}$ modulo $(I_{Y_z})_u$ is a ${\mathbb{C}}$-basis of $Y_{z,u}$ for everey $z\not\in\mathcal S$;
(iii) otherwise if $\{\tilde R_1,\ldots,\tilde R_p,v_1',\ldots,v'_{\xi_u-p}\}$ is linearly dependent over $\mathcal K_{\mathcal Q}$ then the set of equivalent classes of $v_1',\ldots,v'_{\xi_u-p}$ modulo $(I_{Y_z})_u$ is not a ${\mathbb{C}}$-basis of $Y_{z,u}$.
Then, we have $\xi_u-p=H_{Y_z}(u)$ for all z outside $\mathcal S\cup\mathcal S'.$ Now, consider the holomorphic map F from $\Delta(R)$ into ${\mathbb{P}}^{\xi_u-p-1}({\mathbb{C}})$ with the representation
Since f is algebraically non-degenerate over $\mathcal K_{\mathcal Q}$, F is linearly non-degenerate over $\mathcal K_{\mathcal Q}$.
Now, for $z\not\in\mathcal S\cup\mathcal S'$, we set $\textbf{c}_z = (c_{1,z},\ldots,c_{q,z})\in\mathbb Z^{q},$ where
By the definition of the Hilbert weight, there are $\textbf{a}_{1,z},\ldots,\textbf{a}_{\xi_u-p,z}\in\mathbb N^{q}$ with
where $a_{i,j,z}\in\{1,\ldots,\xi_u\},$ such that the residue classes modulo $(I_Y)_u$ of $\textbf{y}^{\textbf{a}_{1,z}},\ldots,$$\textbf{y}^{\textbf{a}_{\xi_u-p,z}}$ form a basic of ${\mathbb{C}}[y_1,\ldots,y_q]_u/(I_{Y_z})_u$ and
Note that $\textbf{y}^{\textbf{a}_{i,z}}\in\mathcal T$ and the set $\{\tilde R_1,\ldots,\tilde R_p,\textbf{y}^{\textbf{a}_{1,z}},\ldots,\textbf{y}^{\textbf{a}_{\xi_u-p,z}}\}$ is a basis of $\mathcal K_{\mathcal Q}[y_1,\ldots,y_q]$ (by Claim 3.1(iii)). Therefore, the set of equivalent classes of $\left\{\textbf{y}^{\textbf{a}_{1,z}},\ldots,\right.$$\left.\textbf{y}^{\textbf{a}_{\xi_u-p,z}}\right\}$ is a basis of $\dfrac{\mathcal K_{\mathcal Q}[y_1,\ldots,y_q]_u}{I(Y)_u}$. Then $ \textbf{y}^{\textbf{a}_{i,z}}=L_{i,z}\left(v_1,\ldots,v_{H_Y(u)}\right)\ \text{modulo}$ $I(Y)_u, $ where $L_{i,z}\ (1\le i\le \xi_u-p)$ are $\mathcal K_{\mathcal Q}$-independent linear forms with coefficients in $\mathcal K_{\mathcal Q}$. We have
where $C_1\in\mathcal C_{\mathcal Q}$. Note that the number of these linear forms $L_{i,z}$ is finite, at most ξu. Denote by $\mathcal L$ the set of all $L_{i,z}$ occurring in the above inequalities. The above inequality follows that
where $C_2(z)\in\mathcal C_{\mathcal Q}$. Then, we have
where the maximum is taken over all subsets $\mathcal J\subset\mathcal L$ with $\sharp\mathcal J=\xi_u-p$ and $\{L|L\in\mathcal J\}$ is linearly independent over $\mathcal K$. From Theorem 2.1 we have
Combining (3.2) and (3.3), we get
where $C_3(z)\in\mathcal C_{\mathcal Q}$, for every $z\in \Delta (R)$ outside a discrete subset.
Fix a point $z\in\Delta(R)\setminus(\mathcal S\cup \mathcal S')$. Choose $i\in\{1,\ldots,n_0\}$ such that
Since $\bigcap_{j=0}^{\ell_i-1}\tilde Q_{\sigma_i(j)}(z)^*\cap V\ne\emptyset$, by Lemma 2.2, we have
Then, from (3.2), (3.4) and (3.5) we have
for every $z\in \Delta (R)$ outside a discrete subset.
Denote by Ψ the set of all the coefficients of all linear forms $L_{i,z}$ and suppose that $\Psi=\{a_1,\ldots,a_{q_0}\}$. Then, we see that $\sharp\mathcal L\le\xi_u,\sharp\Psi=q_0\le \xi_u(\xi_u-p)$. For each positive integer m, denote by $\mathcal L(\Psi(m))$ the $C-$vector space generated by the set $\{a_1^{i_1}\ldots a_{q_0}^{i_{q_0}}|i_j\ge 0\ \text{and }\sum_{j=1}^{q_0}i_j\le m\}$. By Remark 3.4 in [Reference Thai and Quang10], there exists the smallest integer p ʹ such that
Put $s=\dim\mathcal L(\Psi (p')), t=\dim\mathcal L(\Psi (p'+1))$. Again, by [Reference Thai and Quang10, Remark 3.4], we have
Here, the last inequality comes from the fact that $\xi_u\le (u+1)^q$ and
Choose $\{b_1, \ldots ,b_s\}$ an ${\mathbb{C}}$-basis of $\mathcal L(\Psi(p'))$ and $\{b_1, \ldots ,b_t\}$ an ${\mathbb{C}}$-basis of $\mathcal L(\Psi(p'+1))$. Consider the holomorphic map $\tilde F:\Delta(R)\rightarrow\mathbb{P}^{t(\xi_u-p)-1}(\mathbb{C})$ with a presentation
Note that $\|\ T_{\tilde F}(r)=duT_f(r)+o(T_f(r))$ and $c_{\tilde F}=\frac{1}{du}c_f$. By Theorem 2.3, we have
where $\max_{\mathcal J\subset\mathcal L}$ is taken over all subsets $\mathcal J$ of the system $\mathcal L$ of linear forms such that $\mathcal J$ is linearly independent over $\mathcal {K}_{\mathcal {Q}}$, $\epsilon'$ is an arbitrary positive number. Then, by integrating (3.6) and using the above inequality, we obtain
We now estimate the quantity $N_{W(\tilde F)}(r)$. Let $z\in\Delta(R)$ which is neither zero nor pole of any coefficients of $\tilde Q_i\ (1\le i\le q)$ and $b_i\ (1\le i\le t)$. We set
Then there are
such that $\textbf{y}^{\textbf{a}_1},...,\textbf{y}^{\textbf{a}_{H_Y(u)}}$ is a basic of ${\mathbb{C}}[y_1,\ldots,y_q]_u/(I_{Y_z})_u$ and
Similarly as above, we write $\textbf{y}^{\textbf{a}_i}=L_i(v_1,...,v_{\xi_u-p})$, where $L_1,...,L_{\xi_u-p}$ are linearly independent linear forms. We see that
where $n_u=(\xi_u-p)t-1$. It is easy to see that
and hence
Thus, we have
Choose an index $\sigma_{i_0}$ such that $\nu^0_{\tilde Q_{\sigma_{i_0}(0)}({\bf{f}})}(z)\ge \nu^0_{\tilde Q_{\sigma_{i_0}(1)}({\bf{f}})}(z)\ge\cdots\ge \nu^0_{\tilde Q_{\sigma_{i_0}(q-1)}({\bf{f}})}(z)$. By Lemma 2.2 we have
where R i is the resultant of the family $\{Q_{\sigma_i(j)}\}_{j=0}^{\ell_i}$ for $i=1,\ldots,q$.
On the other hand, by Theorem 2.1 we have that
Combining this inequality and (3.8), we have
Integrating both sides of this inequality, we obtain
Seting $m_0=\frac{1}{\Delta_{V}}\,-\,\frac{(2n+1)(n+1)\delta}{u}$ and combining inequalities (3.7) with the above inequality, we get
This inequality implies that
a) We choose $u=\lceil 2\Delta_{V}(2n+1)(n+1)d^n\deg V(\Delta_{V}(n+1)+\epsilon)\epsilon^{-1}\rceil$.
Then $u\ge \biggl\lceil \dfrac{\Delta_{V}(2n+1)(n+1)\delta(\Delta_{V}(n+1)+\epsilon)}{\Delta_{V}(n+1)+\epsilon-\frac{\Delta_{V}(n+1)t}{s}}\biggl\rceil,$ and we have:
• $q-\dfrac{(n+1)t}{sm_0}\ge q-\dfrac{\Delta_V(n+1)t}{(1-\Delta_V(2n+1)(n+1)\delta/u)s}$
$\ge q-\Delta_{V}(n+1)-\epsilon;$
• $n_u+1=(\xi_u-p)t$
$ \le d^n\deg V(u+1)^n\left[\bigl(1+\frac{\epsilon}{2(n+1)\Delta_{V}}\bigl)^{\bigl[\frac{d^n\deg V(u+1)^{n+q}}{\log^2(1+\frac{\epsilon}{2(n+1)\Delta_{V}})}\bigl]+1}\right]=L;$
• $\dfrac{(t(\xi_u-p)-1)t(n+1)}{2sdm_0u} \lt \dfrac{t(n+1)}{sm_0}\cdot\dfrac{(L-1)}{2du}$
$ \le \dfrac{(\Delta_{V}(n+1)+\epsilon)(L-1)}{2du}.$
Then, from (3.9) we have
\begin{align*} \bigl\|\ &(q-\Delta_{V}(n+1)-\epsilon)T_f(r)\\ &\le\sum_{j=1}^{q}\frac{1}{d}N^{[L-1]}_{Q_j(f)}(r)+\frac{(\Delta_{V}(n+1)+\epsilon)(c_f+\epsilon')(L-1)}{2du}T_f (r). \end{align*}The assertion a) is proved.
b) If all Qi are fixed hypersurfaces then $t=s=1$. Choosing $u'=\lceil \Delta_{V}(2n+1)$$(n+1)d^n\deg V(\Delta_{V}(n+1)+\epsilon)\epsilon^{-1}\rceil$ and replacing u in the above by u ʹ, we have
• $q-\dfrac{(n+1)}{m_0}\ge q-\Delta_{V}(n+1)-\epsilon,$
• $n_{u'}+1=\xi_{u'}-p \lt \delta\left(\begin{matrix}{n+u'}\\{n}\end{matrix}\right)\le d^n\deg V\left(\begin{matrix}{n+u'}\\{n}\end{matrix}\right)$
$\le \left[d^n\deg Ve^n\left(1+\dfrac{u'}{n}\right)^{n}\right]$
$\le \left[d^{n^2+n}(\deg V)^{n+1}e^n(2n+5)^n(\Delta^2_V(n+1)\epsilon^{-1}+\Delta_V)^n\right]=L'.$
Similarly as above, from (3.9) we have the desired inequality of the assertion b).
Hence, the proof of the theorem is completed.$\square$
Remark. For the case $\Delta(R)={\mathbb{C}}$, we have $c_f=0$ if f is non-constant. In this case, (3.7) can be re-written as follows:
Here, the notation “$\|$” means that the inequality hold for all $r\in [1,+\infty)$ outside a set of finite measure.
Using (3.10) instead of (3.7), from the above proof, we obtain the following second main theorem for holomorphic maps from ${\mathbb{C}}$.
Theorem 3.11. Let V be a smooth subvriety of dimension n of $\mathbb{P}^{N}(\mathbb{C})$. Let $f:{\mathbb{C}}\rightarrow V$ be a holomorphic mapping. Let $\mathcal {Q}=\{Q_{1},\ldots,Q_{q}\}$ be a set of slowly (with respect to f) moving hypersurfaces with the distributive constant $\Delta_{V}$ with respect to V. Assume that f is algebraically non-degenerate over $\mathcal{K}_{\mathcal {Q}}$. Let $d=lcm(\deg Q_1,\ldots,\deg Q_q)$. Then for every $(n+1)\Delta_V \gt \epsilon \gt 0,$
where
with $u=\lceil 2\Delta_{V}(2n+1)(n+1)d^n\deg V(\Delta_{V}(n+1)+\epsilon)\epsilon^{-1}\rceil$.
Moreover, if all $Q_i\ (1\le i\le q)$ are assumed to be fixed hypersurfaces, then for every ϵ > 0 we have the inequality (3.12) with $L=\bigl[d^{n^2+n}(\deg V)^{n+1}e^n(2n+5)^n(\Delta^2_V(n+1)\epsilon^{-1}+\Delta_V)^n\bigl]$.
We also note that our proof is valid for the case of holomorphic maps from higher dimension complex spaces ${\mathbb{C}}^m$ into V. This theorem is an improvement of many previous second main theorem for hypersurface targets, such as [Reference Dethloff and Tan1, Reference Giang3, Reference Quang4, Reference Quang6, Reference Ru7, Reference Shi9].
Competing interests
No potential conflict of interest was reported by the author.