1 Introduction and results

Let \(\mathbf R\) and \(\mathbf{R}_+\) be the set of all real numbers and the set of all positive real numbers, respectively. We denote by \(\mathbf{R}^{n} (n\ge 2)\) the \(n\)-dimensional Euclidean space. A point in \(\mathbf{R}^{n}\) is denoted by \(P=(X,x_n),\) \(X=(x_1,x_2,\ldots ,x_{n-1}).\) The Euclidean distance of two points \(P\) and \(Q\) in \(\mathbf{R}^{n}\) is denoted by \(|P-Q|\). Also \(|P-O|\) with the origin \(O\) of \(\mathbf{R}^{n}\) is simply denoted by \(|P|\). The boundary and the closure of a set \(\mathbf S\) in \(\mathbf{R}^{n}\) are denoted by \(\partial \mathbf{S}\) and \(\overline{\mathbf{S}}\), respectively.

We introduce a system of spherical coordinates \((r,\Theta ),\) \(\Theta =(\theta _1,\theta _2,\ldots ,\theta _{n-1}),\) in \(\mathbf{R}^{n}\) which are related to cartesian coordinates \((X,x_n)=(x_1,x_2,\ldots ,x_{n-1},x_n)\) by \(x_n=r\cos \theta _1\).

For \(P\in \mathbf{R}^{n}\) and \(r>0\), let \(B(P,r)\) denote the open ball with center at \(P\) and radius \(r\) in \(\mathbf{R}^{n}\). \(S_{r}=\partial {B(O,r)}\). The unit sphere and the upper half unit sphere in \(\mathbf{R}^{n}\) are denoted by \(\mathbf{S}^{n-1}\) and \(\mathbf{S}_{+}^{n-1}\), respectively. For simplicity, a point \((1,\Theta )\) on \(\mathbf{S}^{n-1}\) and the set \(\{\Theta ; (1,\Theta )\in \Omega \}\) for a set \(\Omega \), \(\Omega \subset \mathbf{S}^{n-1},\) are often identified with \(\Theta \) and \(\Omega \), respectively. For two sets \(\Lambda \subset \mathbf{R}_+\) and \(\Omega \subset \mathbf{S}^{n-1},\) the set \(\{(r,\Theta )\in \mathbf{R}^{n}; r\in \Lambda ,(1,\Theta )\in \Omega \}\) in \(\mathbf{R}^{n}\) is simply denoted by \(\Lambda \times \Omega .\) In particular, the half space \(\mathbf{R}_{+}\times \mathbf{S}_{+}^{n-1}=\{(X,x_n)\in \mathbf{R}^{n}; x_n>0\}\) will be denoted by \(\mathbf{T}_n\).

By \(C_n(\Omega )\), we denote the set \(\mathbf{R}_+\times \Omega \) in \(\mathbf{R}^{n}\) with the domain \(\Omega \) on \(\mathbf{S}^{n-1}.\) We call it a cone. We denote the sets \(I\times \Omega \) and \(I\times \partial {\Omega }\) with an interval on \(\mathbf R\) by \(C_n(\Omega ;I)\) and \(S_n(\Omega ;I)\). By \(S_n(\Omega ; r)\) we denote \(C_n(\Omega )\cap S_{r}\). By \(S_n(\Omega )\) we denote \(S_n(\Omega ; (0,+\infty ))\) which is \(\partial {C_n(\Omega )}-\{O\}.\)

Furthermore, we denote by \(d\sigma _{Q}\) (resp. \(dS_{r}\)) the \((n-1)\)-dimensional volume elements induced by the Euclidean metric on \(\partial {C_n(\Omega )}\) (resp. \(S_{r}\)) and by \(dw\) the elements of the Euclidean volume in \(\mathbf{R}^{n}\).

Let \(\fancyscript{A}_a\) denote the class of nonnegative radial potentials \(a(P)\), i.e. \(0\le a(P)=a(r)\), \(P=(r,\Theta )\in C_n(\Omega )\), such that \(a\in L_{loc}^{b}(C_n(\Omega ))\) with some \(b> {n}/{2}\) if \(n\ge 4\) and with \(b=2\) if \(n=2\) or \(n=3\).

This article is devoted to the stationary Schrödinger equation

$$\begin{aligned} Sch_au(P)=-\Delta u(P)+a(P)u(P)=0\quad \mathrm{{for}}\quad P\in C_n(\Omega ), \end{aligned}$$

where \(\Delta \) is the Laplace operator and \(a\in \fancyscript{A}_a\). These solutions are called \(a\)-harmonic functions or generalized harmonic functions (g.h.f.s) associated with the operator \(Sch_a\). Note that they are classical harmonic functions in the case \(a=0\). Under these assumptions the operator \(Sch_a\) can be extended in the usual way from the space \(C_0^{\infty }(C_n(\Omega ))\) to an essentially self-adjoint operator on \(L^{2}(C_n(\Omega ))\) (see [11, Ch. 13]). We will denote it \(Sch_a\) as well. This last one has a Green function \(G(\Omega ,a)(P,Q)\) which is positive on \(C_n(\Omega )\) and its inner normal derivative \(\partial G(\Omega ,a)(P,Q)/{\partial n_Q}\ge 0\), where \({\partial }/{\partial n_{Q}}\) denotes the differentiation at \(Q\) along the inward normal into \(C_n(\Omega )\). We denote this derivative \(P(\Omega ,a)(P,Q)\), which is called the Poisson \(a\)-kernel with respect to \(C_n(\Omega )\).

Let \(\Delta ^{*}\) be a Laplace-Beltrami operator (spherical part of the Laplace) on \(\Omega \subset \mathbf{S}^{n-1}\) and \(\lambda _j~(j=1,2,3\ldots , 0<\lambda _1<\lambda _2\le \lambda _3\le \ldots )\) be the eigenvalues of the eigenvalue problem for \(\Delta ^{*}\) on \(\Omega \) (see, e.g., [12, p. 41])

$$\begin{aligned}&\Delta ^{*}\varphi (\Theta )+\lambda \varphi (\Theta )=0 \quad \mathrm{{in}} \quad \Omega ,\\&\varphi (\Theta )=0 \quad \mathrm{{on}} \quad \partial {\Omega }. \end{aligned}$$

Corresponding eigenfunctions are denoted by \(\varphi _{j}(\Theta )\). We set \(\lambda _0=0\), norm the eigenfunctions in \(L^2(\Omega )\) and \(\varphi _1(\Theta )>0\).

In order to ensure the existences of \(\lambda _j~(j=1,2,3,\ldots )\). We put a rather strong assumption on \(\Omega \): if \(n\ge 3,\) then \(\Omega \) is a \(C^{2,\alpha }\)-domain \((0<\alpha <1)\) on \(\mathbf{S}^{n-1}\) surrounded by a finite number of mutually disjoint closed hypersurfaces (e.g. see [5, p. 88–89] for the definition of \(C^{2,\alpha }\)-domain), \(\varphi _j\in C^2(\overline{\Omega })\) \((j=1,2,3,\ldots )\) and \({\partial \varphi _1}/{\partial n}>0\) on \(\partial {\Omega }\) (here and below, \({\partial }/{\partial n}\) denotes differentiation along the interior normal).

Here well-known estimates (see, e.g., [2], and also [4, p. 120 and p. 126–128] imply the following inequalities:

$$\begin{aligned} M_1j^{\frac{2}{n-1}}\le \lambda _j ~~(j=1,2,3,\ldots ) \end{aligned}$$
(1.1)

and

$$\begin{aligned} |\varphi _j(\Theta )|\le M_2j^{\frac{1}{2}}~~~~(\Theta \in \Omega , j=1,2,3,\ldots ), \end{aligned}$$
(1.2)

where \(M_1\) and \(M_2\) are two positive constants.

Let \(V_j(r)\) and \(W_j(r)\) stand, respectively, for the increasing and non-increasing, as \(r\rightarrow +\infty \), solutions of the equation

$$\begin{aligned} -Q^{\prime \prime }(r)-\frac{n-1}{r}Q^{\prime }(r)+\left( \frac{\lambda _j}{r^2}+a(r)\right) Q(r)=0, \quad 0<r<\infty , \end{aligned}$$
(1.3)

normalized under the condition \(V_j(1)=W_j(1)=1\).

We will also consider the class \(\fancyscript{B}_a\), consisting of the potentials \(a\in \fancyscript{A}_a\) such that there exists the finite limit \(\lim \nolimits _{r\rightarrow \infty }r^2 a(r)=k\in [0,\infty )\), moreover, \(r^{-1}|r^2 a(r)-k|\in L(1,\infty )\). If \(a\in \fancyscript{B}_a\), then the g.h.f.s are continuous (see [13]).

In the rest of paper, we assume that \(a\in \fancyscript{B}_a\) and we shall suppress this assumption for simplicity. Meanwhile, we use the standard notations \(u^{+}=\max \{u,0\}\), \(u^{-}=-\min \{u,0\}\) and \([d]\) is the integer part of \(d\), where \(d\) is a positive real number.

Denote

$$\begin{aligned} \iota _{j,k}^{\pm }=\frac{2-n\pm \sqrt{(n-2)^2+4(k+\lambda _j)}}{2}\quad (j=0,1,2,3\ldots ). \end{aligned}$$

The solutions to Eq. (1.3) have the asymptotic (see [6])

$$\begin{aligned} V_j(r)\sim M_3 r^{\iota _{j,k}^{+}}, W_j(r)\sim M_4 r^{\iota _{j,k}^{-}}, \mathrm{as}\,\, r\rightarrow \infty , \end{aligned}$$
(1.4)

where \(M_3\) and \(M_4\) are some positive constants.

Further, we have

$$\begin{aligned} \iota _{j,k}^{+}\ge \iota _{j,0}^{+}> M_5 j^{\frac{1}{n-1}}\quad (j=1,2,3\ldots ) \end{aligned}$$
(1.5)

from (1.1), where \(M_5\) is a positive constant independent of \(j\).

If \(a\in \fancyscript{A}_a\), it is known that the following expansion for the Green function \(G(\Omega ,a)(P,Q)\) (see [3, Ch. 11])

$$\begin{aligned} G(\Omega ,a)(P,Q)=\sum _{j=0}^{\infty }\frac{1}{\chi ^{\prime }(1)}V_j(\min \{r,t\}) W_j(\max \{r,t\})\varphi _{j}(\Theta )\varphi _{j}(\Phi ), \end{aligned}$$

where \(P=(r,\Theta )\), \(Q=(t,\Phi )\), \(r\ne t\) and \(\chi ^{\prime }(t)=w\left( W_1(r),V_1(r)\right) |_{r=1}\) is their Wronskian. This series converges uniformly if either \(r\le s t\) or \(t\le s r\) \((0<s<1)\). In the case \(a=0\), this expansion coincides with the well-known result by Lelong-Ferrand (see [9]).

For a nonnegative integer \(m\) and two points \(P=(r,\Theta ), Q=(t,\Phi )\in C_n(\Omega )\), we put

$$\begin{aligned} K(\Omega ,a,m)(P,Q)=\left\{ \begin{array}{ll} 0 &{} \quad \text{ if }\ \ 0<t<1, \\ \widetilde{K}(\Omega ,a,m)(P,Q) &{}\quad {\text{ if }} \ \ 1\le t<\infty , \end{array}\right. \end{aligned}$$

where

$$\begin{aligned} \widetilde{K}(\Omega ,a,m)(P,Q)=\sum _{j=0}^{m}\frac{1}{\chi ^{\prime }(1)}V_j(r)W_j(t)\varphi _{j} (\Theta )\varphi _{j}(\Phi ). \end{aligned}$$

To obtain the modified Poisson integral representation for the Schrödinger operator in a cone, we use the following modified kernel function defined by

$$\begin{aligned} G(\Omega ,a,m)(P,Q)=G(\Omega ,a)(P,Q)-K(\Omega ,a,m)(P,Q) \end{aligned}$$

for two points \(P=(r,\Theta ), Q=(t,\Phi )\in C_n(\Omega ).\)

Write

$$\begin{aligned} U(\Omega ,a,m;u)(P)=\int _{S_n(\Omega )}P(\Omega ,a,m)(P,Q)u(Q)d\sigma _Q, \end{aligned}$$

where

$$\begin{aligned} P(\Omega ,a,m)(P,Q)=\frac{\partial G(\Omega ,a,m)(P,Q)}{\partial n_Q},\quad P(\Omega ,a,0)(P,Q)=P(\Omega ,a)(P,Q) \end{aligned}$$

and \(u(Q)\) is a continuous function on \(\partial C_n(\Omega )\).

Now we define the function \(\rho (R)\) under consideration. Hereafter, the function \(\rho (R)~(\ge 1)\) is always supposed to be nondecreasing and continuously differentiable on the interval \([0,+\infty )\). We assume further that

$$\begin{aligned} \epsilon _{0}=\limsup _{R\rightarrow \infty } \frac{(\iota _{{[\rho (R)]+1},k}^{+})^{\prime }R\ln R}{\iota _{{[\rho (R)]+1},k}^{+}}<1. \end{aligned}$$
(1.6)

Remark

\(\iota _{{[\rho (R)]+1},k}^{+}\) in (1.6) is not the function \(V_j(R)\). For any \(\epsilon \) \((0<\epsilon <1-\epsilon _0)\), there exists a sufficiently large positive number \(R_{\epsilon }\) such that \(R>R_{\epsilon }\), by (1.5) and (1.6) we have

$$\begin{aligned} M(\rho (R))^{\frac{1}{n-1}}<\iota _{{[\rho (R)]+1},k}^{+}< \iota _{{[\rho (e)]+1,k}}^{+}(\ln R)^{\epsilon _0+\epsilon }, \end{aligned}$$

where \(M\) is a positive constant.

For positive real numbers \(\beta \), we denote \(\mathcal C _{\Omega ,\beta ,a}\) the class of all measurable functions \(f(t,\Phi )\) \((Q=(t,\Phi )\in C_n(\Omega ))\) satisfying the following inequality

$$\begin{aligned} \int _{C_n(\Omega )}\frac{|f(t,\Phi )|\varphi _1}{1+V_{[\rho (t)]+1} (t)t^{n+\beta -1}}dw<\infty \end{aligned}$$
(1.7)

and the class \(\mathcal D _{\Omega ,\beta ,a}\), consists of all measurable functions \(g(t,\Phi )\,(Q=(t,\Phi )\in S_n(\Omega ))\) satisfying

$$\begin{aligned} \int _{S_n(\Omega )}\frac{|g(t,\Phi )|V_1(t)W_1(t)}{1+\chi ^{\prime }(t)V_{[\rho (t)]+1}(t)t^{n+\beta -2}} \frac{\partial \varphi _1}{\partial n}d\sigma _Q<\infty , \end{aligned}$$
(1.8)

where \(\chi ^{\prime }(t)=w\left( W_1(r),V_1(r)\right) |_{r=t}\) is their Wronskian.

We will also consider the class of all continuous functions \(u(t,\Phi )\) \(((t,\Phi )\in \overline{C_n(\Omega )})\) generalized harmonic in \(C_n(\Omega )\) with \(u^{+}(t,\Phi )\in \mathcal C _{\Omega ,\beta ,a}\,((t,\Phi )\in C_n(\Omega ))\) and \(u^{+}(t,\Phi )\in \mathcal D _{\Omega ,\beta ,a}\,((t,\Phi )\in S_n(\Omega ))\) is denoted by \(\mathcal E _{\Omega ,\beta ,a}\).

Next we define the order of g.h.f, which is similar to the F. Riesz’ definition for the order of classical harmonic function (see [7, Definition 4.1]). We shall say that a g.h.f.-\(u(P)\,(P=(r,\Theta )\in C_n(\Omega ))\) is of order \(\lambda \) if

$$\begin{aligned} \lambda =\limsup _{r\rightarrow \infty } \frac{\log \left( \sup _{C_n(\Omega )\cap S_{r}}|u|\right) }{\log r}. \end{aligned}$$

If \(\lambda <\infty \), then \(u\) is said to be of finite order.

In case \(\lambda <\infty \), about the solutions of the Dirichlet problem for the Schrödinger operator with continuous data in \(\mathbf{T}_n\), we refer the readers to the paper by Kheyfits (see [8]).

Motivated by Kheyfits’s conclusions, we prove the following results for the g.h.f.s of infinite order. In the case \(a=0\), we refer readers to the paper by Qiao (see [10]).

Theorem 1

If \(u\in \mathcal E _{\Omega ,\beta ,a}\), then \(u\in \mathcal D _{\Omega ,\beta ,a}\).

Theorem 2

If \(u\in \mathcal E _{\Omega ,\beta ,a}\), then the following properties hold:

  1. (I)

    \(U(\Omega , a,[\rho (t)];u)(P)\) is a g.h.f. on \(C_n(\Omega )\) and can be continuously extended to \(\overline{C_n(\Omega )}\) such that \(U(\Omega , a,[\rho (t)];u)(P)=u(P)\) for \(P=(r,\Theta )\in S_n(\Omega )\).

  2. (II)

    There exists an infinite series \(h(P)=\sum \nolimits _{j=1}^{\infty }A_jV_j(r)\varphi _j(\Theta )\) vanishing continuously on \(\partial {C_n(\Omega )}\) such that

    $$\begin{aligned} u(P)=U(\Omega , a,[\rho (t)];u)(P)+h(P) \end{aligned}$$

    for \(P=(r,\Theta )\in C_n(\Omega )\), where \(A_j\) \((j=1,2,3,\ldots )\) is a constant.

2 Lemmas

The following Lemma generalizes the Carleman’s formula (referring to the holomorphic functions in the half space) (see [1]) to the g.h.f.s in a cone, which is due to Levin and Kheyfits (see [3, Ch. 11]).

Lemma 1

If \(u(t,\Phi )\) is a g.h.f. on a domain containing \(C_n(\Omega ;(1,R))\), then

$$\begin{aligned}&m_{+}(R)+\int _{S_n(\Omega ;(1,R))}u^{+}\Psi (t) \frac{\partial \varphi _1}{\partial n} d\sigma _Q+M_6+\frac{W_1(R)}{V_1(R)}M_7=m_{-}(R)\\&\quad + \int _{S_n(\Omega ;(1,R))}u^{-}\Psi (t) \frac{\partial \varphi _1}{\partial n} d\sigma _Q, \end{aligned}$$

where

$$\begin{aligned} \Psi (t)=W_1(t)-\frac{W_1(R)}{V_1(R)}V_1(t),~~~~m_{\pm }(R)= \int _{S_n(\Omega ;R)}\frac{\chi ^{\prime }(R)}{V_1(R)}u^{\pm }\varphi _1 d S_{R}, \end{aligned}$$
$$\begin{aligned} M_6=\int _{S_n(\Omega ;1)}u\varphi _1W^{\prime }_1(1)-W_1(1)\varphi _1 \frac{\partial u}{\partial n}dS_1\quad \mathrm{{and}} \quad M_7=\int _{S_n(\Omega ;1)}V_1(1)\varphi _1\frac{\partial u}{\partial n}-u\varphi _1V^{\prime }_1(1) dS_1. \end{aligned}$$

Lemma 2

(see [3, Ch. 11]) For a non-negative integer \(m\), we have

$$\begin{aligned} |P(\Omega ,a,m)(P,Q)|\le M_8V_{m+1}(2r)\frac{W_{m+1}(t)}{t}\varphi _1(\Theta )\frac{\partial \varphi _1(\Phi )}{\partial n_\Phi } \end{aligned}$$
(2.1)

for any \(P=(r,\Theta )\in C_n(\Omega )\) and \(Q=(t,\Phi )\in S_n(\Omega )\) satisfying \(2r\le t\), where \(M_8\) is a constant depending only \(n\).

Lemma 3

If \(h(r,\Theta )\) is a g.h.f. in \(C_n(\Omega )\) vanishing continuously on \(\partial {C_n(\Omega )}\), then

$$\begin{aligned} h(r,\Theta )=\sum _{j=1}^{\infty }B_jV_j(r)\varphi _j(\Theta ), \end{aligned}$$
(2.2)

where the series converges uniformly and absolutely in any compact set of \(\overline{C_n(\Omega )}\), and \(B_j\) \((j=1,2,3,\ldots )\) is a constant satisfying

$$\begin{aligned} B_jV_j(r)=\int _\Omega h(r,\Theta )\varphi _j(\Theta )d S_1 \end{aligned}$$
(2.3)

for every \(r~(0<r<\infty )\).

Proof

Set

$$\begin{aligned} y_{j}(r)=\int _\Omega h(r,\Theta )\varphi _j(\Theta )d S_1~~~~~(j=1,2,3,\ldots ). \end{aligned}$$

Making use of the assumptions on \(h\) and self-adjoint of the Laplace-Beltrami operator \(\Delta ^{*}\), one can check directly (by differentiating under the integral sign) that the functions \(y_j~(j=1,2,3,\ldots )\) satisfy the Eq. (1.3). This equation has a general solution \(y_j(r)=B_jV_j(r)+D_jW_j(r)\), where \(B_j\) and \(D_j\) are constants independent of \(r~(j=1,2,3,\ldots )\). We note that \(h(r,\Theta )\) converges uniformly to zero as \(r\rightarrow 0\) and hence \(\lim \nolimits _{r\rightarrow 0}y_j(r)=0~(j=1,2,3,\ldots ).\) Thus we see that \(D_j=0~(j=1,2,3,\ldots )\). Since \(y_j(r)\) takes the value \(y_j(r_1)\) at \(r=r_1\), we have

$$\begin{aligned} y_j(r)=\frac{V_j(r)}{V_j(r_1)}y_j(r_1) \end{aligned}$$

for any \(r\) and \(r_1\) \((0<r,r_1<R_1)\), where \(0<R_1\le +\infty \). In particular, if \(R_1=\infty \), then

$$\begin{aligned} \lim _{r\rightarrow \infty }\frac{y_j(r)}{V_j(r)}=B_{j}~~~(j=1,2,3,\ldots ) \end{aligned}$$
(2.4)

exists.

Since \(y_j(r_1)\rightarrow y_j(R_1)\) as \(r_1\rightarrow R_1\), we see that

$$\begin{aligned} y_j(r)=\frac{V_j(r)}{V_j(R_1)}y_j(R_1)~~~(j=1,2,3,\ldots ), \end{aligned}$$
(2.5)

which gives

$$\begin{aligned} |y_j(r)|\le M_2w_n\left( \frac{r}{R_1}\right) ^{\iota _{j,k}^{+}}j^{\frac{1}{2}}\sup _{\Theta \in \Omega }|h(R_1,\Theta )| \end{aligned}$$

from (1.2) and (1.4), where \(w_n\) is the surface area \(2\pi ^{n/2}\{\Gamma (n/2)\}^{-1}\) of \(\mathbf{S}^{n-1}\).

Now we set

$$\begin{aligned} J=\max \left\{ j;~\iota _{j,k}^{+}-\iota _{{l+1},k}^{+} <\frac{1}{2}M_5j^{\frac{1}{n-1}}\right\} ~~(l=1,2,3,\ldots ). \end{aligned}$$

The existence of this \(J\) is known from (1.5). Hence if we put

$$\begin{aligned} M_{9}=w_nM_2^2\left\{ \sum _{j=l+1}^{J}j\left( \frac{1}{2}\right) ^{\iota _{j,k}^{+} -\iota _{{m+1},k}^{+}}+\sum _{j=J+1}^{\infty } j\left( \frac{1}{2}\right) ^{2^{-1}M_5j^{\frac{1}{n-1}}}\right\} \end{aligned}$$

and use (1.2), then from the completeness of \(\{\varphi _j(\Theta )\}\) we can expand \(h(r,\Theta )\) into the Fourier series

$$\begin{aligned} h(r,\Theta )=\sum _{j=1}^{\infty }y_j(r)\varphi _j(\Theta ) \end{aligned}$$

satisfying

$$\begin{aligned} \sum _{j=l+1}^{\infty }|y_j(r)||\varphi _j(\Theta )|\le M_{9}\left( \frac{r}{R_1}\right) ^{\iota _{{l+1},k}^{+}}\sup _{\Theta \in \Omega }|h(R_1,\Theta )|~~(l=1,2,3,\ldots ) \qquad \end{aligned}$$
(2.6)

on \(C_n(\Omega ;(0,\frac{R_1}{2}))\), where \(M_{9}\) is a positive constant independent of \(r\) and \(R_1\).

Take any compact \(H\), \(H\subset \overline{C_n(\Omega )}\) and a number \(R_1\) satisfying \(R_1>2\max \{r;(r,\Theta )\in H\}\). So we can represent \(h(r,\Theta )\) as

$$\begin{aligned} h(r,\Theta )=\sum _{j=1}^{\infty }y_j(r)\varphi _j(\Theta ), \end{aligned}$$
(2.7)

where \((r,\Theta )\) is a point in \(H\). Hence we observe in (2.5) that \(y_j(r)\) is a number independent of \(R_1\). Hence as \(R_1\rightarrow \infty \), we see from (2.4) that \(y_j(r)=B_jV_j(r)\), which is (2.3). This and (2.7) give (2.2).

To prove the absolute and uniform convergence of (2.7) on \(H\), see from (2.6) that

$$\begin{aligned} \sum _{j=l+1}^{\infty }|y_j(r)||\varphi _j(\Theta )|\le M_{9}2^{-\iota _{{l+1},k}^{+}}\sup _{\Theta \in \Omega }|h(R_1,\Theta )|, \end{aligned}$$

which converges to 0 as \(l\rightarrow \infty \). Then Lemma 3 is proved. \(\square \)

3 Proof of Theorem 1

Since \(u\in \mathcal E _{\Omega ,\beta ,a}\), we obtain by (1.7)

$$\begin{aligned} \int _{1}^{\infty }\frac{m_{+}(R)V_1(R)}{\chi ^{\prime }(R)V_{[\rho (R)]+1}(R)R^{n+\beta -1}}dR \!\le \! 2\int _{C_n(\Omega )}\frac{u^{+}\varphi _1}{1+V_{[\rho (t)]+1}(t)t^{n+\beta -1}}dw\!<\!\infty . \qquad \quad \end{aligned}$$
(3.1)

From (1.4) and (1.8), we conclude that

$$\begin{aligned}&\int _{1}^{\infty }\frac{V_1(R)}{\chi ^{\prime }(R)V_{[\rho (R)]+1} (R)R^{n+\beta -1}}\int _{S_n(\Omega ;(1,R))}u^{+}\Psi (t) \frac{\partial \varphi _1}{\partial n} d\sigma _Q d R\nonumber \\&\quad \le \frac{2\chi _{1,k}}{(\chi _{1,k}+\beta )\beta } \int _{S_n(\Omega )}\frac{u^{+}V_1(t)W_1(t)}{1+\chi ^{\prime }(t)V_{[\rho (t)]+1} (t)t^{n+\beta -2}}\frac{\partial \varphi _1}{\partial n}d\sigma _Q \nonumber \\&\quad <\infty , \end{aligned}$$
(3.2)

where \(\chi _{1,k}=\iota _{1,k}^{+}-\iota _{1,k}^{-}.\)

It follows from (1.4), Remark and the L’hospital’s rule

$$\begin{aligned} \lim _{t\rightarrow \infty }\frac{\chi ^{\prime }(t)V_{[\rho (t)]+1} (t)t^{n+\beta -2}}{W_1(t)}\int _{t}^{\infty }\frac{V_1(R)}{\chi ^{\prime }(R)V_{[\rho (R)]+1}(R)R^{n+\frac{\beta }{2}-1}} \left( \frac{W_1(t)}{V_1(t)}-\frac{W_1(R)}{V_1(R)}\right) d R=+\infty , \end{aligned}$$

which yields that there exists a positive constant \(M_{10}\) such that for any \(t\ge 1\),

$$\begin{aligned} \int _{t}^{\infty }\frac{V_1(R)}{\chi ^{\prime }(R)V_{[\rho (t)]+1} (R)R^{n+\frac{\beta }{2}-1}}\Psi (t)dR\ge \frac{M_{10}V_1(t)W_1(t)}{\chi ^{\prime }(t)V_{[\rho (t)]+1}(t)t^{n+\beta -2}}. \end{aligned}$$

From (3.1), (3.2) and Lemma 1 we see that

$$\begin{aligned}&M_{10}\int _{S_n(\Omega ;(1,\infty ))} \frac{u^{-}V_1(t)W_1(t)}{\chi ^{\prime }(t)V_{[\rho (t)]+1}(t)t^{n+\beta -2}} \frac{\partial \varphi _1}{\partial n}d\sigma _Q\\&\quad \le \int _{S_n(\Omega ;(1,\infty ))}u^{-}\int _{t}^{\infty } \frac{V_1(R)}{\chi ^{\prime }(R)V_{[\rho (t)]+1}(R)R^{n+\frac{\beta }{2}-1}} \Psi (t) dR \frac{\partial \varphi _1}{\partial n} d\sigma _Q\\&\quad <\infty . \end{aligned}$$

Then Theorem 1 is proved from \(|u|=u^{+}+u^{-}\).

4 Proof of Theorem 2

Let \(l_1\) be any positive number such that \(l_1\ge 2\beta \). For any fixed \(P=(r,\Theta )\in C_n(\Omega )\), take a number \(\sigma \) satisfying \(\sigma >\sigma _r=\max \{[2r]+1,\vartheta _r \}\), where \(\vartheta _r=\exp (\frac{l_1}{\beta }\iota _{{[\rho (e)]+1,k}}^{+} 2^{1+\epsilon _0+\epsilon }\ln 2r)^{\frac{1}{1-\epsilon _0-\epsilon }}\).

From the Remark we see that there exists a constant \(M(r)\) dependent only on \(r\) such that \(M(r)\ge (2r)^{\iota _{{[\rho (i+1)]+1,k}}^{+}}i^{-\frac{\beta }{l_1}}\) from \(\sigma \ge \vartheta _r\).

By (1.4), (1.8), (2.1) and Theorem 1, we have

$$\begin{aligned}&\int _{S_n(\Omega ;(\sigma ,\infty ))}|P(\Omega ,a,[\rho (t)]) (P,Q)||u(Q)|d\sigma _{Q}\\&\quad \le M_8\varphi _1(\Theta )\sum _{i=\sigma _r}^{\infty } \int _{S_n(\Omega ;[i,i+1))}\frac{(2r)^{\iota _{{[\rho (t)]+1},k}^{+}}}{t^{\frac{\beta }{l_1}}}\frac{|u(t,\Phi )|}{V_{[\rho (t)]+1} (t)t^{n-2+\frac{\beta }{l_1}}}d\sigma _{Q} \\&\quad \le M_8\sum _{i=\sigma _r}^{\infty } \frac{(2r)^{\iota _{{[\rho (i+1)]+1,k}}^{+}}}{i^{\frac{\beta }{l_1}}} \int _{S_n(\Omega ;[i,i+1))}\frac{|u(t,\Phi )|}{V_{[\rho (t)]+1} (t)t^{n-2+\frac{\beta }{l_1}}}d\sigma _{Q}\\&\quad \le M_8M(r)\varphi _1(\Theta ) \int _{S_n(\Omega ;[\sigma _r,\infty ))}\frac{|u(t,\Phi )|}{V_{[\rho (t)]+1}(t)t^{n-2+\frac{\beta }{l_1}}}d\sigma _{Q}\\&\quad <\infty . \end{aligned}$$

Hence \(U(\Omega ,a,[\rho (t)];u)(P)\) is absolutely convergent and finite for any \(P\in C_n(\Omega )\). Thus \(U(\Omega ,a,[\rho (t)];u)(P)\) is generalized harmonic on \(C_n(\Omega )\).

Now we study the boundary behavior of \(U(\Omega ,a,[\rho (t)];u)(P)\). Let \(Q^{\prime }=(t^{\prime },\Phi ^{\prime })\in \partial {C_n(\Omega )}\) be any fixed point and \(l_2\) be any positive number such that \(l_2>t^{\prime }+1\).

Set \(\chi _{S(l_2)}\) is the characteristic function of \(S(l_2)=\{Q=(t,\Phi )\in \partial {C_n(\Omega )},t\le l_2\}\) and write

$$\begin{aligned} U(\Omega ,a,[\rho (t)];u)(P)=U^{\prime }(P)-U^{\prime \prime }(P)+U^{\prime \prime \prime }(P), \end{aligned}$$

where

$$\begin{aligned} U^{\prime }(P)&= \int _{S_n(\Omega ;(0,2l_2])}P(\Omega ,a)(P,Q)u(Q)d\sigma _Q,\\ U^{\prime \prime }(P)&= \int _{S_n(\Omega ;(1,2l_2])}\frac{\partial { K(\Omega ,a,[\rho (t)])(P,Q)}}{\partial n_Q}u(Q)d\sigma _Q \end{aligned}$$

and

$$\begin{aligned} U^{\prime \prime \prime }(P)=\int _{S_n(\Omega ;(2l_2,\infty ))}P (\Omega ,a,[\rho (t)])(P,Q)u(Q)d\sigma _Q. \end{aligned}$$

Notice that \(U^{\prime }(P)\) is the Poisson integral of \(u(Q)\chi _{S(2l_2)}\), we have \(\lim \nolimits _{P\in C_n(\Omega ),P\rightarrow Q^{\prime }} U^{\prime }(P)=u(Q^{\prime })\). Since \(\lim \nolimits _{\Theta \rightarrow \Phi ^{\prime }}\varphi _{j}(\Theta )=0~(j=1,2,3\ldots )\) as \(P=(r,\Theta )\rightarrow Q^{\prime }=(t^{\prime },\Phi ^{\prime })\in S_n(\Omega )\), we have \(\lim \nolimits _{P\in C_n(\Omega ),P\rightarrow Q^{\prime }}U^{\prime \prime }(P)=0\) from the definition of the kernel function \(K(\Omega ,a,[\rho (t)])(P,Q)\). \(U^{\prime \prime \prime }(P)=O(M(r)\varphi _1(\Theta ))\) and therefore tends to zero.

So the function \(U(\Omega ,a,[\rho (t)];u)(P)\) can be continuously extended to \(\overline{C_n(\Omega )}\) such that

$$\begin{aligned} \lim \limits _{P\in C_n(\Omega ),P\rightarrow Q^{\prime }}U(\Omega ,a,[\rho (t)];u)(P)=u(Q^{\prime }) \end{aligned}$$

for any \(Q^{\prime }=(t^{\prime },\Phi ^{\prime })\in \partial {C_n(\Omega )}\) from the arbitrariness of \(l_2\).

So (I) is proved. Finally (I) and Lemma 3 give the conclusion of (II). Then we complete the proof of Theorem 2.