\documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2003(2003), No. 03, pp. 1--31.\newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu (login: ftp)} \thanks{\copyright 2002 Southwest Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE--2003/03\hfil Singular solutions to Protter's problem] {Singular solutions to Protter's problem for the 3-D wave equation involving lower order terms} \author[M. K. Grammatikopoulos, T. D. Hristov, \& N. I. Popivanov \hfil EJDE--2003/03\hfilneg] {Myron. K. Grammatikopoulos, Tzvetan D. Hristov, \& Nedyu I. Popivanov} \address{Myron K. Grammatikopoulos \hfill\break Department of Mathematics, University of Ioannina\\ 451 10 Ioannina, Greece} \email{mgrammat@cc.uoi.gr} \address{Tzvetan D. Hristov \hfill\break Department of Mathematics and Informatics, University of Sofia, 1164 Sofia, Bulgaria} \email{tsvetan@fmi.uni-sofia.bg} \address{Nedyul I. Popivanov \hfill\break Department of Mathematics and Informatics, University of Sofia, 1164 Sofia, Bulgaria} \email{nedyu@fmi.uni-sofia.bg} \date{} \thanks{Submitted May 27, 2002. Published January 2, 2003.} \subjclass[2000]{35L05,35L20, 35D05, 35A20} \keywords{Wave equation, boundary value problems, generalized solutions, \hfill\break\indent singular solutions, propagation of singularities} \begin{abstract} In 1952, at a conference in New York, Protter formulated some boundary value problems for the wave equation, which are three-dimensional analogues of the Darboux problems (or Cauchy-Goursat problems) on the plane. Protter studied these problems in a 3-D domain $\Omega_0$, bounded by two characteristic cones $\Sigma_1$ and $\Sigma_{2,0}$, and by a plane region $\Sigma_0$. It is well known that, for an infinite number of smooth functions in the right-hand side, these problems do not have classical solutions. Popivanov and Schneider (1995) discovered the reason of this fact for the case of Dirichlet's and Neumann's conditions on $\Sigma_0$: the strong power-type singularity appears in the generalized solution on the characteristic cone $\Sigma_{2,0}$. In the present paper we consider the case of third boundary-value problem on $\Sigma_0$ and obtain the existence of many singular solutions for the wave equation involving lower order terms. Especifica ally, for Protter's problems in $\mathbb{R}^{3}$ it is shown here that for any $n\in N$ there exists a $C^{n}(\bar{\Omega}_0)$-function, for which the corresponding unique generalized solution belongs to $C^{n}(\bar{\Omega}_0\backslash O)$ and has a strong power type singularity at the point $O$. This singularity is isolated at the vertex $O$ of the characteristic cone $\Sigma_{2,0}$ and does not propagate along the cone. For the wave equation without lower order terms, we presented the exact behavior of the singular solutions at the point $O$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{definition}[theorem]{Definition} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{remark}[theorem]{Remark} \section{Introduction} Consider the hyperbolic partial differential equation, involving the wave operator in its main part, with lower order terms of the form \begin{equation} Lu\equiv u_{x_1x_1}+u_{x_2x_2}-u_{tt}+b_1u_{x_1}+b_2u_{x_2}+bu_{t}+cu=f \label{eq:0p1} \end{equation} expressed in Cartesian coordinates $x_1,x_2,t$ in a simply connected region $\Omega_0\subset \mathbb{R}^{3}$. The region \begin{equation*} \Omega_0:=\{(x_1,x_2,t):0\varepsilon \},\quad \varepsilon \in \lbrack 0,1), \end{equation*} which in polar coordinates becomes \begin{equation} \Omega_{\varepsilon }=\{(\varrho ,\varphi ,t):t>0,\,0\leq \varphi <2\pi ,\varepsilon +t<\varrho <1-t\}. \label{eq:0p8} \end{equation} We define \textsl{generalized solution} of Problem $P_{\alpha }$ in $\Omega_{\varepsilon },\varepsilon \in (0,1)$, in Definition \ref{def2.1} below. Note that, if a generalized solution $u$ belongs to {$C^{1}({\bar{\Omega}}_{\varepsilon })\cap $}$C^2({\Omega }_{\varepsilon })$, it is called a \textsl{classical solution} of Problem $P_{\alpha }$ in $\Omega_{\varepsilon }$, $\varepsilon \in (0,1)$, and it satisfies the equation \eqref{eq:0p1} in $\Omega_{\varepsilon }$. It should be pointed out that the case $\varepsilon =0$ is totally different from the case $% \varepsilon \neq 0$. This paper is a generalization, extension, and improvement of the results obtained in \cite{GHP}. The paper, besides Introduction, consists of six more sections. In Section 2, using some appropriate techniques, we formulate the 2-D boundary value problems $P_{\alpha ,1}$, $P_{\alpha ,2}$ and $P_{\alpha ,3}$, corresponding to the 3-D Problem $P_{\alpha }$. The aim of Section 3 is to treat Problem $P_{\alpha ,3}$. For this reason, we construct and study the system of integral equations, assigned to the under consideration equation \eqref{eq:0p1}. Also, we present results concerning the classical solutions of Problem $P_{\alpha ,3}$ in $\Omega_{\varepsilon },\varepsilon \in (0,1)$ and give corresponding a priori estimates. We mention also here Lemma \ref{lm3.1}, which is actually a maximum principle for Problem $P_{\alpha ,3}$. In Section 4 we prove Theorems \ref{thm4.1} and \ref{thm4.2}, which ensure the existence and uniqueness of a generalized solution of Problem $P_{\alpha ,1}$ in 2-D domain. Using the results of the previous section, in Section 5 we study the existence and uniqueness of a generalized solution of 3-D Problem $P_{\alpha }$. More precisely, Theorem \ref{thm5.1} ensures the uniqueness of a generalized solution for Problem $P_{\alpha }$ in $\Omega _{\varepsilon },\varepsilon \in \lbrack 0,1)$, while Theorems \ref{thm5.2}, \ref{thm5.3} and \ref{thm5.4} ensure the existence of a generalized solution for problem $P_{\alpha }$ in $\Omega_0$, which is a classical one in each domain $\Omega _{\varepsilon },\varepsilon \in \lbrack 0,1)\ $and satisfies some a priori estimates in $C^2(\Omega_{\varepsilon })$. Comparing these estimates with such ones of \cite{GHP}, we see that the new estimates are better even in the case of the wave equation \eqref{eq:0p30} without lower order terms. In Theorems \ref{thm6.1}, \ref{thm6.2} and \ref{thm6.3} under different conditions, imposed on the coefficients of the equation \eqref{eq:0p1}, we present some singular \textsl{generalized solutions} which are smooth enough away from the point $% O$, while at the point $O$ they have power type singularity of the type $% (x_1^2+x_2^2+t^2)^{-n/2}$. More precisely, we formulate and prove the following theorem. \begin{theorem} \label{thm6.3} Let the coefficients $b_1,b_2,b$ be constants, $c(x_1,x_2,t)=c(|x|,t)\in C^{1}(\bar{\Omega}_0)$, $4c\leq b_1^2+b_2^2-b^2$ in $\bar{\Omega}_0$ and $\alpha =\alpha (|x|)\in C^{1}(0,1]\cap C[0,1]$. Then for each $n\in \mathbb{N}$ there exists a function $f_{n}(x_1,x_2,t)\in C^{n-2}(\bar{\Omega}_0)\cap C^{\infty }(\Omega _0)$, for which the corresponding generalized solution $u_{n}$ of problem $P_{\alpha }$ belongs to $C^2(\bar{\Omega}_0\backslash O)$ and satisfies the estimate \begin{equation} |u_{n}(x_1,x_2,|x|)|\geq \frac{1}{2}|u_{n}(2x_1,2x_2,0)|+|x|^{-n}|% \cos n(\arctan \frac{x_2}{x_1})|. \label{eq:0p101a} \end{equation} \end{theorem} On the other hand, in Theorems \ref{thm6.1} and \ref{thm6.2} the coefficients of the equation \eqref{eq:0p1} are nonconstant. Finally, in Theorem \ref{thm7.1} for the wave equation \eqref{eq:0p30} we present two-sided estimates for the singularities of the \textsl{generalized solution }of Problem $P_{\alpha }$. In particular, the exact behavior for the singular solution $u_{n}(x_1,x_2,t)$ around $O$ is $(x_1^2+x_2^2+t^2)^{-n/2}\cos n(\arctan \frac{x_2}{x_1})$. \begin{remark} \label{rmk1.2} \rm Actually, all these results state some conditions on the coefficients of the equation \eqref{eq:0p1}, under which we do not have a positive answer to Open Question 1. For example, Theorem \ref{thm7.1} ensures that for any parameter $\alpha (x)$, involved to the boundary condition \eqref{eq:0p2} on $\Sigma_0$, there are infinitely many singular solutions of the wave equation \eqref{eq:0p30}. That means that, it is impossible to give a positive answer to Open Question 1, by using the wave operator only. Possibly, it is necessary to ask some of the nonzero lower order perturbations of the wave equation to be involved to the more general equation \eqref{eq:0p1}. This is one of the reasons of the existence of the present paper, where we use and developed further the ideas of \cite{GHP}. Note also that, each one of \ the singular solutions has a strong singularity at the vertex $O$ of the cone $% \Sigma_{2,0}$. The singularities of the generalized solutions do not propagate in the direction of the bicharacteristics on the characteristic cone. It is traditionally assumed that the wave equation, with right-hand side sufficiently smooth in $\bar{\Omega}_0$, cannot have a solution with an isolated singular point. For results, concerning the propagation of singularities for second order operators, see H\"{o}rmander \cite[Chapter 24.5]{H}. \end{remark} We conclude this section with the following four more questions. \noindent\textbf{Open Questions:} \begin{enumerate} \item Find the exact behavior of all singular solutions at the point $O$, different from those ones which appear in Theorems \ref{thm6.1}, \ref{thm6.2} and \ref{thm7.1}. \item Find appropriate conditions for the function $f$ under which the Problem $% P_{\alpha }$, even for the wave equation, has only classical solutions. We do not know any kind of such results even for Problem $P2$. \item In all results, concerning the existence of singular solutions (except Theorem \ref{thm6.3}), we assume that $a_2\equiv 0$. Is it possible to find any singular solution, when $a_2\neq 0?$ Even in the case $a_2\neq 0$, Theorem \ref{thm5.3} ensures the existence of a generalized solution for any function $f$, but we do not know the behavior of a such solution at $% (0,0,0)$. \item From the a priori estimates, obtained in Theorems \ref{thm5.2}--\ref{thm5.4} for all solutions of Problem $P_{\alpha }$, including singular ones, it follows that, as $\rho \to 0$, none of these solutions can grow up faster than exponential one. The arising question is: are there singular solutions of Problem $P_{\alpha }$ with exponential growth as $\rho \to 0$ or any such solution is of polynomial growth satisfying \eqref{eq:0p101a}? \end{enumerate} In the case of Problem $P1$, for the wave equation \eqref{eq:0p1} the answer to Open Questions 1, 2 and 4 above can be found in \cite{NT}. \section{Preliminaries} In this section we consider \eqref{eq:0p1} in polar coordinates $x_1=\varrho \cos \varphi $, $x_2=\varrho \sin \varphi $ and $t$ \begin{equation} Lu=\frac{1}{\varrho }(\varrho u_{\varrho })_{\varrho }+\frac{1}{\varrho ^2}% u_{\varphi \varphi }-u_{tt}+a_1u_{\varrho }+a_2u_{\varphi }+bu_{t}+cu=f \label{eq:1p1} \end{equation} in a simply connected region \begin{equation} \Omega_{\varepsilon }:=\big\{(\varrho ,\varphi ,t):00$, if: \begin{enumerate} \item $u\in C^{1}({\bar{\Omega}}_{\varepsilon })$, $u\bigr\rvert_{\Sigma _1\cap \partial \Omega_{\varepsilon }}=0$; $[u_{t}+\alpha (\varrho )u] \bigr\rvert_{\Sigma_0\cap \partial \Omega_{\varepsilon }}=0$; \item The equality \begin{equation} \label{eq:1p6} \begin{aligned} &\int_{\Omega_{\varepsilon }}[u_{t}v_{t}-u_{\varrho }v_{\varrho } -\frac{1}{\varrho ^2}u_{\varphi }v_{\varphi }+(a_1u_{\varrho } +a_2u_{\varphi}+bu_{t}+cu-f)v]\varrho \,d\varrho \,d\varphi \,dt \\ &=\int_{\Sigma_0\cap \partial \Omega_{\varepsilon }}\alpha (\varrho )uv\varrho \,d\varrho d\varphi \end{aligned} \end{equation} holds for all \begin{equation*} v\in V_{\varepsilon }:=\{v\in C^{1}({\bar{\Omega}}_{\varepsilon }):[v_{t}+(\alpha +b)v]\bigr\rvert_{\Sigma_0\cap \partial \Omega _{\varepsilon }}=0,v\bigr\rvert_{\Sigma_{2,\varepsilon }}=0\}. \end{equation*} \end{enumerate} \end{definition} The following lemma describes the properties of generalized solutions of Problem $P_{\alpha }$ in $\Omega_{\varepsilon }$. \begin{lemma} \label{lm2.1} Each \textsl{generalized solution} of Problem $P_{\alpha }$ in $\Omega_0$ is also a \textsl{generalized solution} of the same problem in $\Omega_{\varepsilon }$ for $\varepsilon >0$. \end{lemma} The proof of this lemma follows from the proof in \cite[Lemma 2.1]{GHP}. In the special, but main case, when \begin{equation} f(\varrho ,\varphi ,t)=f_{n}^{(1)}(\varrho ,t)\cos n\varphi +f_{n}^{(2)}(\varrho ,t)\sin n\varphi , \label{eq:1p7} \end{equation} we ask the generalized solution to be of the form \begin{equation*} u(\varrho ,\varphi ,t)=u_{n}^{(1)}(\varrho ,t)\cos n\varphi +u_{n}^{(2)}(\varrho ,t)\sin n\varphi . \end{equation*} Then, in view of \eqref{eq:1p1}, we obtain the 2-D system \begin{equation} \label{eq:1p9} \begin{gathered} \frac{1}{\varrho }(\varrho u_{n,\varrho }^{(1)})_{\varrho }-u_{n,tt}^{(1)}+a_1u_{n,\varrho }^{(1)}+bu_{n,t}^{(1)}+(c-\frac{n^2} {\varrho ^2})u_{n}^{(1)}+na_2u_{n}^{(2)}=f_{n}^{(1)}, \\ \frac{1}{\varrho }(\varrho u_{n,\varrho }^{(2)})_{\varrho }-u_{n,tt}^{(2)}+a_1u_{n,\varrho }^{(2)}+bu_{n,t}^{(2)}+(c-\frac{n^2} {\varrho ^2})u_{n}^{(2)}-na_2u_{n}^{(1)}=f_{n}^{(2)}. \end{gathered} \end{equation} We consider this system in the domain \begin{equation*} G_{\varepsilon }=\{(\varrho ,t):t>0,\varepsilon +t<\varrho <1-t\} \end{equation*} which is bounded by the sets: \begin{equation}\label{eq:1p10} \begin{gathered} S_0 =\{(\varrho ,t):t=0,0<\varrho <1\}, \\ S_1=\{(\varrho ,t):\varrho =1-t\},\\ S_{2,\varepsilon }=\{(\varrho ,t):\varrho =t+\varepsilon \}. \end{gathered} \end{equation} In this case, for $u=(u^{(1)},u^{(2)})(\varrho ,t)$, the 2-D problem corresponding to $P_{\alpha }$ is $P_{\alpha ,1}$: \begin{equation} \begin{gathered} \frac{1}{\varrho }(\varrho u_{\varrho }^{(1)})_{\varrho }-u_{tt}^{(1)}+a_1u_{\varrho }^{(1)}+bu_{t}^{(1)}+(c-\frac{n^2}{\varrho ^2})u^{(1)}+na_2u^{(2)}=f^{(1)}\text{ \quad in }G_{\varepsilon }, \\ \frac{1}{\varrho }(\varrho u_{\varrho }^{(2)})_{\varrho }-u_{tt}^{(2)}+a_1u_{\varrho }^{(2)}+bu_{t}^{(2)}+(c-\frac{n^2}{\varrho ^2})u^{(2)}-na_2u^{(1)}=f^{(2)}\text{ \quad in }G_{\varepsilon }, \\ u^{(i)}\bigr\rvert_{S_1\cap \partial G_{\varepsilon }}=0,\quad \lbrack u_{t}^{(i)}+\alpha (\varrho )u^{(i)}]\bigr\rvert_{S_0\cap \partial G_{\varepsilon }}=0,\quad i=1,2. \end{gathered} \label{eq:1p12} \end{equation} The generalized solution of the Problem $P_{\alpha ,1}$ is as follows. \begin{definition} \label{def2.2} \rm A function $u=(u^{(1)},u^{(2)})(\varrho ,t)$ is called a \textsl{generalized solution} of Problem $P_{\alpha ,1}$ in $G_{\varepsilon }$, $\varepsilon >0$, if: \begin{enumerate} \item $u\in C^{1}({\bar{G}}_{\varepsilon })$, $[u_{t}^{(i)}+\alpha (\varrho )u^{(i)}]\bigr\rvert_{S_0\cap \partial G_{\varepsilon }}=0$, $u^{(i)}% \bigr\rvert_{S_1\cap \partial G_{\varepsilon }}=0,i=1,2;$ \item The equalities \begin{equation} \begin{gathered} \int_{G_{\varepsilon }}\big[u_{t}^{(1)}v_{1,t}-u_{\varrho }^{(1)}v_{1,\varrho }+\big( a_1u_{\varrho }^{(1)}+bu_{t}^{(1)}+(c -\frac{n^2}{\varrho ^2})u^{(1)}+na_2u^{(2)} -f^{(1)}\big) v_1\big] \varrho d\varrho \,dt \\ =\int_{S_0\cap \partial G_{\varepsilon }}\alpha (\varrho )u^{(1)}v_1\varrho \,d\varrho ,\\ \int_{G_{\varepsilon }}[u_{t}^{(2)}v_{2,t}-u_{\varrho }^{(2)}v_{2,\varrho }+\big( a_1u_{\rho }^{(2)}+bu_{t}^{(2)} +(c-\frac{n^2}{\varrho ^2})u^{(2)}-na_2u^{(1)} -f^{(2)}\big) v_2\big] \varrho d\varrho \,dt\\ =\int_{S_0\cap \partial G_{\varepsilon }}\alpha (\varrho )u^{(2)}v_2\varrho \,d\varrho \end{gathered} \label{eq:1p14} \end{equation} hold for all \begin{equation*} v_1,v_2\in V_{\varepsilon }^{(1)}=\{{v\in C^{1}({\bar{G}}_{\varepsilon }):[v_{t}+(\alpha +b)v]\bigr\rvert_{S_0\cap \partial G_{\varepsilon }}=0,v% \bigr\rvert_{S_{2,\varepsilon }\cap \partial G_{\varepsilon }}=0\}.} \end{equation*} \end{enumerate} \end{definition} Introducing a new function \begin{equation} z^{(i)}(\varrho ,t)=\varrho ^{\frac{1}{2}}u^{(i)}(\varrho ,t),\ i=1,2, \label{eq:1p15} \end{equation} we transform the system \eqref{eq:1p12} to the system \begin{equation} \begin{gathered} z_{\varrho \varrho }^{(1)}-z_{tt}^{(1)}+a_1z_{\varrho }^{(1)}+bz_{t}^{(1)}+\big( c-\frac{1}{2\varrho }a_1-\frac{4n^2-1}{% 4\varrho ^2}\big) z^{(1)}+na_2z^{(2)}=\varrho ^{\frac{1}{2}}f^{(1)}, \\ z_{\varrho \varrho }^{(2)}-z_{tt}^{(2)}+a_1z_{\varrho}^{(2)}+bz_{t}^{(2)}+\big( c-\frac{1}{2\varrho }a_1-\frac{4n^2-1}{% 4\varrho ^2}\big) z^{(2)}-na_2z^{(1)}=\varrho ^{\frac{1}{2}}f^{(2)}. \end{gathered} \label{eq:1p16} \end{equation} with the string operator in the main part. Substituting the new coordinates \begin{equation} \xi =1-\varrho -t,\quad \eta =1-\varrho +t, \label{eq:1p17} \end{equation} from \eqref{eq:1p16} we derive \begin{equation} \begin{gathered} U_{\xi \eta }^{(1)}-A_1U_{\xi }^{(1)}-B_1U_{\eta }^{(1)}-C_1U^{(1)}-D_1U^{(2)}=F^{1}(\xi ,\eta )\text{ \quad in }% D_{\varepsilon }, \\ U_{\xi \eta }^{(2)}-A_2U_{\xi }^{(2)}-B_2U_{\eta }^{(2)}-C_2U^{(2)}-D_2U^{(1)}=F^2(\xi ,\eta )\text{ \quad in }% D_{\varepsilon }, \end{gathered} \label{eq:1p18} \end{equation} where $D_{\varepsilon }=\{(\xi ,\eta ):0<\xi <\eta <1-\varepsilon \}$ and for $i=1,2$: \begin{equation} \begin{gathered} U^{(i)}(\xi ,\eta )=z^{(i)}(\varrho (\xi ,\eta ),t(\xi ,\eta )), \\ F^{(i)}(\xi ,\eta )=\frac{1}{4\sqrt{2}}(2-\eta -\xi )^{\frac{1% }{2}}f^{(i)}(\varrho (\xi ,\eta ),t(\xi ,\eta )). \end{gathered} \label{eq:1p19} \end{equation} Note that, in the case of the equation \eqref{eq:1p1}, the corresponding coefficients of the system \eqref{eq:1p18} are \begin{equation} \begin{gathered} A_1=A_2=\frac{1}{4}(a_1+b),B_1=B_2=\frac{1}{4} (a_1-b),D_2=-D_1=\frac{1}{4}na_2, \\ C_1=C_2=\frac{1}{4}\big( \frac{4n^2-1}{(2-\xi -\eta )^2% }+\frac{a_1}{2-\xi -\eta }-c\big) . \end{gathered} \label{eq:1p101} \end{equation} As we see, Problem $P_{\alpha ,1}$ is reduced to the Darboux-Goursat problem for the system \eqref{eq:1p18} with the same boundary conditions. That is, we consider the following question. \noindent\textbf{Problem $P_{\alpha ,2}$.} Find a solution $(U^{1},U^2)$ of the system \eqref{eq:1p18} in $D_{\varepsilon }$ with the boundary conditions \begin{equation} \ U^{(i)}(0,\eta )=0,\quad (U_{\eta }^{(i)}-U_{\xi }^{(i)})(\xi ,\xi )+\alpha (1-\xi )U^{(i)}(\xi ,\xi )=0,\ i=1,2. \label{eq:1p20} \end{equation} To investigate the smoothness or the singularity of a solution of the original 3-D problem $P_{\alpha }$ on $\Sigma_{2,0}$, we are seeking a classical solution of the corresponding 2-D problem $P_{\alpha ,2}$ not only in the domain $D_{\varepsilon }$, but also in the domain \begin{equation} D_{\varepsilon }^{(1)}:=\{(\xi ,\eta ):0<\xi <\eta <1,0<\xi <1-\varepsilon \},\quad \varepsilon >0. \end{equation} Clearly, $D_{\varepsilon }\subset D_{\varepsilon }^{(1)}$. Consider now an appropriate boundary value problem for the system of equations whose coefficients are continuous in ${\bar{D}}_{\varepsilon}^{(1)}$, $\varepsilon >0$: \noindent\textbf{Problem $P_{\alpha ,3}$.} Find a solution $(U^{1},U^2)$ of the system \begin{equation} \begin{gathered} U_{\xi \eta }^{(1)}-A_1U_{\xi }^{(1)}-B_1U_{\eta }^{(1)}-C_1U^{(1)}-D_1U^{(2)}=F^{1}(\xi ,\eta )\text{ \quad in }{\bar{D}}% _{\varepsilon }^{(1)}, \\ U_{\xi \eta }^{(2)}-A_2U_{\xi }^{(2)}-B_2U_{\eta }^{(2)}-C_2U^{(2)}-D_2U^{(1)}=F^2(\xi ,\eta )\text{ \quad in }{\bar{D}}% _{\varepsilon }^{(1)}, \end{gathered} \label{eq:1p21} \end{equation} with the boundary conditions \begin{equation} U^{(i)}(0,\eta )=0,(U_{\eta }^{(i)}-U_{\xi }^{(i)})(\xi ,\xi )+\alpha (1-\xi )U^{(i)}(\xi ,\xi )=0, \label{bound} \end{equation} for $i=1,2$ and $\xi \in (0,1-\varepsilon )$. \section{The system of integral equations corresponding to Problem $P_{\protect\alpha ,3}$} First of all, we construct an equivalent system of integral equations in such a way that any solution to Problem $P_{\alpha ,3}$ is a solution of the constructed system and vice-verca. For this reason, following the ideas of [9], concerning the representation of the solutions of the Protter's problem for the wave equation, we will try to find a corresponding representation of the solutions in the case, where the equation \eqref{eq:0p1} involves lower order terms. In this case, because of the appearance in the equation % \eqref{eq:1p1} of the term $a_2u_{\varphi }$, we have to deal not with a single scalar equation, but with a system of equations. In order to realize these ideas, for any $(\xi_0,\eta_0)\in D_{\varepsilon }^{(1)}$, we consider the sets \begin{equation*} \Pi :=\{(\xi ,\eta ):0<\xi <\xi_0,\xi_0<\eta <\eta_0\},\quad T:=\{(\xi ,\eta ):0<\xi <\xi_0,\xi <\eta <\xi_0\} \end{equation*} and the following integrals: \begin{gather*} I_0^{(i)}:=\iint_{\Pi }U_{\xi \eta }^{(i)}(\xi ,\eta )\,d\xi d\eta \,=\int_0^{\xi_0}\int_{\xi_0}^{\eta_0}U_{\xi \eta }^{(i)}(\xi ,\eta )\,d\eta \,d\xi\,, \\ I_1^{(i)}:=\iint_{T}U_{\xi \eta }^{(i)}(\xi ,\eta )\,d\xi d\eta \,=\int_0^{\xi_0}\int_{\xi }^{\xi_0}U_{\xi \eta }^{(i)}(\xi ,\eta )\,d\eta \,d\xi \,. \end{gather*} As it has been shown in \cite{GHP}, \begin{equation*} I_0^{(i)}+2I_1^{(i)}=U^{(i)}(\xi_0,\eta_0)-\int_0^{\xi _0}\alpha (1-\xi )U^{(i)}(\xi ,\xi )\,d\xi . \end{equation*} Set $p^{(i)}:=U_{\xi }^{(i)}$, $q^{(i)}:=U_{\eta }^{(i)}$. Then, in view of the last relation and \eqref{eq:1p21}, we obtain \begin{equation} \label{eq:2p80} \begin{aligned} U^{(1)}(\xi_0,\eta_0)=&\int_0^{\xi_0}\int_{\xi_0}^{\eta _0}[F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi ,\eta )\,d\eta \,d\xi\\ &+2\int_0^{\xi_0}\int_0^{\eta }[F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi ,\eta )\,d\xi \,d\eta \\ &+\int_0^{\xi_0}\alpha (1-\xi )U^{(1)}(\xi ,\xi )\,d\xi ,\quad \text{for }(\xi_0,\eta_0)\in {\bar{D}}_{\varepsilon }^{(1)}, \end{aligned} \end{equation} \begin{equation}\label{eq:2p801} \begin{aligned} U^{(2)}(\xi_0,\eta_0)=&\int_0^{\xi_0}\int_{\xi_0}^{\eta _0}[F^2+A_2p^{(2)}+B_2q^{(2)}+C_2U^{(2)}+D_2U^{(1)}](\xi ,\eta )\,d\eta \,d\xi \\ &+2\int_0^{\xi_0}\int_0^{\eta }[F^2+A_2p^{(2)}+B_2q^{(2)}+C_2U^{(2)}+D_2U^{(1)}](\xi ,\eta )\,d\xi \,d\eta \\ &+\int_0^{\xi_0}\alpha (1-\xi )U^{(2)}(\xi ,\xi )\,d\xi ,\text{ \ \ \ for }(\xi_0,\eta_0)\in {\bar{D}}_{\varepsilon }^{(1)}, \end{aligned} \end{equation} which is the first couple of desired integral equations. From $\eqref{eq:2p80}$ and \eqref{eq:2p801} we derive for the first derivatives $p^{(i)}$ and $q^{(i)}$ ($i=1,2$) the next four integral equations: \begin{equation} \begin{aligned} p^{(1)}(\xi_0,\eta_0)=&\int_0^{\xi _0}[F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi ,\xi _0)\,d\xi \\ &+\int_{\xi_0}^{\eta _0}[F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi _0,\eta )\,d\eta \\ +\alpha (1-\xi_0)U^{(1)}(\xi_0,\xi_0), \end{aligned} \label{eq:2p91} \end{equation} \begin{equation} q^{(1)}(\xi_0,\eta_0)=\int_0^{\xi _0}[F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi ,\eta _0)\,d\xi , \label{eq:2p92} \end{equation} \begin{equation} \begin{aligned} p^{(2)}(\xi_0,\eta_0)=&\int_0^{\xi _0}[F^2+A_2p^{(2)}+B_2q^{(2)}+C_2U^{(2)}+D_2U^{(1)}](\xi ,\xi _0)\,d\xi \\ &+\int_{\xi_0}^{\eta _0}[F^2+A_2p^{(2)}+B_2q^{(2)}+C_2U^{(2)}+D_2U^{(1)}](\xi _0,\eta )\,d\eta \\ &+\alpha (1-\xi_0)U^{(2)}(\xi_0,\xi_0), \end{aligned} \label{eq:2p911} \end{equation} \begin{equation} q^{(2)}(\xi_0,\eta_0)=\int_0^{\xi _0}[F^2+A_2p^{(2)}+B_2q^{(2)}+C_2U^{(2)}+D_2U^{(1)}](\xi ,\eta _0)\,d\xi . \label{eq:2p921} \end{equation} Now, we set \begin{equation} \begin{gathered} M:=\max \left( \sup_{D_{\varepsilon }^{(1)}}\lvert F^{1}\rvert ,\sup_{D_{\varepsilon }^{(1)}}\lvert F^2\rvert \right) ,\quad M_{\alpha}:=\sup_{[0,1-\varepsilon ]}\lvert \alpha (\xi )\rvert \\ c(\varepsilon ):=\max_{i=1,2}\left\{ \sup_{D_{\varepsilon }^{(1)}}\lvert A_{i}\rvert ,\ \sup_{D_{\varepsilon }^{(1)}}\lvert B_{i}\rvert ,\ \sup_{D_{\varepsilon }^{(1)}}\lvert C_{i}\rvert ,\ \sup_{D_{\varepsilon }^{(1)}}\lvert D_{i}\rvert \right\} , \end{gathered} \label{eq:2p5} \end{equation} and formulate the following results. \begin{theorem} \label{thm3.1} Let $F^{i},A_{i},B_{i},C_{i},D_{i}\in C({\bar{D}}_{\varepsilon }^{(1)})$, $i=1,2$, $\varepsilon >0$. Then there exists a classical solution $(U^{(1)},U^{(2)})\in C^{1}({\bar{D}}_{\varepsilon }^{(1)})$ of the Problem $P_{\alpha ,3}$ for which $U_{\xi \eta }^{(i)}\in C({\bar{D}}_{\varepsilon }^{(1)}),i=1,2$ and \begin{equation} \begin{gathered} \lvert U^{(i)}(\xi_0,\eta_0)\rvert \leq M[4c(\varepsilon )+M_{\alpha }]^{-2}\exp \left\{ 8c(\varepsilon )+2M_{\alpha }\right\} \quad {\text{in }} D_{\varepsilon }^{(1)},\ i=1,2, \\ \sup_{D_{\varepsilon }^{(1)}}\{\lvert U_{\xi }^{(i)}\rvert ,\lvert U_{\eta }^{(i)}\rvert \}\leq M[4c(\varepsilon )+M_{\alpha }]^{-1}\exp \left\{ 8c(\varepsilon )+2M_{\alpha }\right\} ,\ i=1,2. \end{gathered} \label{eq:2p7} \end{equation} \end{theorem} \noindent\textbf{Proof.} To get our results, we will solve the system of integral equations \eqref{eq:2p80}--\eqref{eq:2p921}. For this reason we use sequence of successive approximations $(U_{m}^{(i)},p_{m}^{(i)},q_{m}^{(i)}),m=1,2,\dots$, defined by the formula{e} \begin{equation} \begin{aligned} U_{m+1}^{(i)}(\xi_0,\eta_0) =&\int_0^{\xi_0}\int_{\xi _0}^{\eta_0}E_{m}^{(i)}(\xi ,\eta ) \,d\eta \,d\xi +2\int_0^{\xi_0}\int_0^{\eta }E_{m}^{(i)}(\xi ,\eta ) \,d\xi \,d\eta \\ &+\int_0^{\xi_0}\alpha (1-\xi )U_{m}^{(i)}(\xi ,\xi )\,d\xi ,\quad \ i=1,2;\quad m=0,1,2\dots \\ p_{m+1}^{(i)}(\xi_0,\eta_0) =&\int_0^{\xi_0}E_{m}^{(i)}(\xi ,\xi_0)\,d\xi +\int_{\xi_0}^{\eta _0}E_{m}^{(i)}(\xi_0,\eta )\,d\eta \\ &+\alpha (1-\xi_0)U_{m}^{(i)}(\xi_0,\xi_0),\ \quad i=1,2;\quad m=0,1,2\dots \\ q_{m+1}^{(i)}(\xi_0,\eta_0)=&\int_0^{\xi _0}E_{m}^{(i)}(\xi ,\eta_0)\,d\xi ,\quad i=1,2;\quad m=0,1,2\dots \\ U_0^{(i)}(\xi_0,\eta_0)=&0,\quad p_0^{(i)}(\xi_0,\eta_0)=0,\\ q_0^{(i)}(\xi_0,\eta_0)=&0,\quad i=1,2,\quad \text{in $D_{\varepsilon}^{1},$} \end{aligned} \label{eq:2p820} \end{equation} where \begin{align*} E_{m}^{(1)}(\xi ,\eta ) &:=[F^{1}+A_1p_{m}^{(1)}+B_1q_{m}^{(1)}+C_1U_{m}^{(1)}+D_1U_{m}^{(2)}](\xi ,\eta ), \\ E_{m}^{(2)}(\xi ,\eta ) &:=[F^2+A_2p_{m}^{(2)}+B_2q_{m}^{(2)}+C_2U_{m}^{(2)}+D_2U_{m}^{(1)}](\xi ,\eta ). \end{align*} We will show that each of the functions $U_{m}^{(i)},p_{m}^{(i)}$ and $q_{m}^{(i)},i=1,2$, is continuous in ${\bar{D}}_{\varepsilon }^{(1)}$ and for any $(\xi_0,\eta_0)\in {\bar{D}}_{\varepsilon }^{(1)}$ and $m\in \mathbb{N}$ \begin{equation} {\lvert (U_{m}^{(i)}-U_{m-1}^{(i)})(\xi_0,\eta_0)\rvert \leq M\frac{% [4c(\varepsilon )+M_{\alpha }]^{m-1}}{(m+1)!}}\text{ }{(\xi_0+\eta_0)}% ^{m+1}{,} \label{eq:2p8} \end{equation} \begin{equation} \begin{aligned} &\max \left\{ {\lvert (p_{m}^{(i)}-p_{m-1}^{(i)})|(\xi_0,\eta_0),\lvert (q_{m}^{(i)}-q_{m-1}^{(i)})|(\xi_0,\eta_0)}\right\}\\ &\leq {M\frac{[4c(\varepsilon )+M_{\alpha }]^{m-1}}{m!}} (\xi_0+\eta_0)^{m} \end{aligned}\label{eq:2p9} \end{equation} Indeed, by induction: 1) For ${m=1}$ \begin{equation*} {U_1^{(i)}(\xi_0,\eta_0)=\int_0^{\xi_0}\int_{\xi }^{\eta_0}}% \text{ }{F}^{(i)}{(\xi ,\eta )\,d\eta \,d\xi +\int_0^{\xi_0}\int_{\xi }^{\xi_0}}\text{ }{F}^{(i)}{(\xi ,\eta )\,d\eta \ d\xi \,,} \end{equation*} and hence{\ \begin{equation*} \lvert U^{(1)}(\xi_0,\eta_0)\rvert \leq M\xi_0\eta_0\leq M(\xi _0+\eta_0)^2/2. \end{equation*} Similarly one can estimate }$p_1^{(i)}$ and $q_1^{(i)}$. {2) Let now, by the induction hypothesis, \eqref{eq:2p8} and \eqref{eq:2p9} be satisfied for} some{\ }$m\in \mathbb{N}$. Then for $i=1,2${\ \begin{equation*} \lvert (U_{m}^{(i)}-U_{m-1}^{(i)})(\xi_0,\eta_0)\rvert \leq M\frac{% [4c(\varepsilon )+M_{\alpha }]^{m-1}}{m!}(\xi_0+\eta_0)^{m}:=Q_{m}(\xi _0+\eta_0)^{m}. \end{equation*} It follows that \begin{align*} &\lvert (U_{m+1}^{(i)}-U_{m}^{(i)})(\xi_0,\eta_0)\rvert\\ & \leq Q_{m}\Big[4c(\varepsilon )\Big( \int_0^{\xi_0}\int_{\xi_0}^{\eta_0}(\xi +\eta )^{m}\,d\eta \,d\xi +2\int_0^{\xi_0}\int_0^{\eta }(\xi +\eta )^{m}\,d\xi\,d\eta \Big) \\ &\quad +\frac{M_{\alpha }}{m+1}\int_0^{\xi_0}(2\xi )^{m+1}\,d\xi \Big]\\ &\leq \frac{Q_{m}}{(m+1)(m+2)}\big[4c(\varepsilon )\big((\xi_0+\eta _0)^{m+2}-\eta_0^{m+2}-\xi_0^{m+2}\big)+M_{\alpha }(2\xi _0)^{m+2}\big] \\ &\leq \frac{Q_{m+1}}{(m+2)}(\xi_0+\eta_0)^{m+2}. \end{align*} By \eqref{eq:2p91} and \eqref{eq:2p911}, we have} \begin{align*} &\lvert (p_{m+1}^{(i)}-p_{m}^{(i)})|(\xi_0,\eta_0) \\ &\leq \frac{Q_{m}}{m+1}\big[4c(\varepsilon )\big((\xi _0+\eta_0)^{m+1}-\xi_0{}^{m+1}\,\big)+M_{\alpha }(2\xi_0)^{m+1}% \big]\leq Q_{m+1}(\xi_0+\eta_0)^{m+1}. \end{align*} A similar estimate holds for (${q_{m+1}^{(i)}-q_{m}^{(i)})}$. So that \eqref{eq:2p8} and \eqref{eq:2p9} hold and hence the uniform convergence of the sequences $\{U_{m}^{(i)}(\xi ,\eta )\}_{m\in \mathbb{N}}$, $\{p_{m}^{(i)}(\xi ,\eta )\}_{m\in \mathbb{N}}$ and $\{q_{m}^{(i)}(\xi ,\eta )\}_{m\in \mathbb{N}}$ in ${\bar{D}}_{\varepsilon }^{(1)}$ follows. For the limit functions $U^{(i)},p^{(i)},q^{(i)}\in C({\bar{D}}_{\varepsilon }^{(1)})$ we obtain the integral equalities \eqref{eq:2p80}--\eqref{eq:2p921} with the obvious condition $U^{(i)}(0,\eta_0)=0$. From the integral equalities \eqref{eq:2p80}--\eqref{eq:2p921} it follows that $p^{(i)}=U_{\xi }^{(i)}$ and $q^{(i)}=U_{\eta }^{(i)}$ in ${\bar{D}}_{\varepsilon }^{(1)}$. Therefore, $U^{(i)}\in C^{1}({\bar{D}}_{\varepsilon}^{(1)}),i=1,2$. Also, in view of \eqref{eq:2p8}, we see that \begin{align*} \lvert (U^{(i)}(\xi_0,\eta_0)\rvert & =\big| \sum_{m=0}^{\infty}(U_{m+1}^{(i)}-U_{m}^{(i)})(\xi_0,\eta_0)\big| \leq M\sum_{m=0}^{\infty }\frac{[4c(\varepsilon )+M_{\alpha }]^{m}}{(m+2)!} (\xi_0+\eta_0)^{m+2} \\ & \leq M[4c(\varepsilon )+M_{\alpha }]^{-2}\exp \left\{ 8c(\varepsilon )+2M_{\alpha }\right\}, \quad i=1,2. \end{align*} So, using {\eqref{eq:2p9}, }for the derivatives $U_{\xi_0}^{(i)}(\xi _0,\eta_0)$ and $U_{\eta_0}^{(i)}(\xi_0,\eta_0)$ we get the estimates \begin{align*} \lvert U_{\xi_0}^{(i)}(\xi_0,\eta_0)\rvert &=\big|\sum_{m=0}^{\infty }(p_{m+1}^{(i)}-p_{m}^{(i)})(\xi_0,\eta_0)\big| \leq M\sum_{m=0}^{\infty }\frac{[4c(\varepsilon )+M_{\alpha }]^{m}}{(m+1)!}(\xi_0+\eta_0)^{m+1} \\ &\leq M[4c(\varepsilon )+M_{\alpha }]^{-1}\exp \left\{ 8c(\varepsilon )+2M_{\alpha }\right\} ,\quad i=1,2 \end{align*} and \begin{equation*} \lvert U_{\eta_0}^{(i)}(\xi_0,\eta_0)\rvert \leq M[4c(\varepsilon )+M_{\alpha }]^{-1}\exp \left\{ 8c(\varepsilon )+2M_{\alpha }\right\} ,\quad i=1,2, \end{equation*} which shows \eqref{eq:2p7}. Also, by \eqref{eq:2p91}--\eqref{eq:2p921}, it follows that \begin{align*} U_{\xi_0\eta_0}^{(1)} &\equiv U_{\eta_0\xi _0}^{(1)}=F^{1}+A_1U_{\xi_0}^{(1)}+B_1U_{\eta _0}^{(1)}+C_1U^{(1)}+D_1U^{(2)}, \\ U_{\xi_0\eta_0}^{(2)} &\equiv U_{\eta_0\xi_0}^{(2)}=F^2\text{ }+A_2U_{\xi_0}^{(2)}+B_2U_{\eta_0}^{(2)}+C_2U^{(2)}+D_2U^{(1)}. \end{align*} Thus, the vector-valued function {$U(\xi_0,\eta_0)$ }is a solution to the system \eqref{eq:1p21} and{\ $U_{\xi \eta }\in C({\bar{D}}_{\varepsilon }^{(1)})$. Finally, using representations \eqref{eq:2p91}--\eqref{eq:2p921} for the first derivatives of }$U^{(i)}${, we conclude that each function $U^{(i)}(\xi_0,\eta_0)$ satisfies the boundary condition \eqref{bound} of the Problem }${P}_{\alpha ,3}$ for $\eta =\xi $. \hfill$\square$ The next lemma is very important for the investigation of the singularity of a generalized solution of Problem $P_{\alpha }$. \begin{lemma} \label{lm3.1} Let $F^{i},A_{i},B_{i},C_{i},D_{i}\in C({\bar{D}}_{\varepsilon }^{(1)})$, $% i=1,2,$ \begin{equation} A_{i}\geq 0,B_{i}\geq 0,C_{i}\geq 0,D_{i}\geq 0,\text{ }\alpha (1-\xi )\geq 0% \text{ in ${\bar{D}}_{\varepsilon }^{(1)}$, $\ i=1,2$} \label{eq:2p101} \end{equation} and \begin{equation} \mathbf{(a)}\text{\ }p_1^{(i)}\geq 0\text{ \ and \ \ }q_1^{(i)}\geq 0, \quad\text{or}\quad(\mathbf{b})\text{\ }F^{(i)}\geq 0\text{\ in }{\bar{D}% }_{\varepsilon }^{(1)},\quad \text{$i=1,2.$} \label{eq:2p101a} \end{equation} Then for the solution $(U^{(1)},U^{(2)})$ of Problem $P_{\alpha ,3}$ (already found in Theorem \ref{thm3.1}) we have \begin{equation} U^{(i)}(\xi ,\eta )\geq 0,\quad U_{\eta }^{(i)}(\xi ,\eta )\geq 0,\quad U_{\xi }^{(i)}(\xi ,\eta )\geq 0\quad \text{for }(\xi ,\eta )\in \text{${% \bar{D}}_{\varepsilon }^{(1)}$},\ i=1,2\text{.} \end{equation} \end{lemma} \noindent\textbf{Proof.} First, suppose that the condition $\mathbf{(b)}$ is satisfied. Then, in view of \eqref{eq:2p820}, for $(\xi_0,\eta_0)\in {\bar{D}}_{\varepsilon }^{(1)}$ we have \begin{gather} U_1^{(i)}(\xi_0,\eta_0)=\int_0^{\xi_0}\int_{\xi_0}^{\eta _0}F^{(i)}(\xi ,\eta )\,d\eta \,d\xi +2\int_0^{\xi_0}\int_0^{\eta }F^{(i)}(\xi ,\eta )\,d\xi \,d\eta \geq 0, \label{eq:2p201} \\ p_1^{(i)}(\xi_0,\eta_0)=\int_0^{\xi_0}F^{(i)}(\xi ,\xi _0)\,\,d\xi +\int_{\xi_0}^{\eta_0}F^{(i)}(\xi_0,\eta )\,\,d\eta \geq 0, \label{eq:2p202} \\ q_1^{(i)}(\xi_0,\eta_0)=\int_0^{\xi_0}F^{(i)}(\xi ,\eta _0)\,d\xi \geq 0,\quad i=1,2. \label{eq:2p203} \end{gather} Thus, the condition $\mathbf{(b)}$\ is {stronger, than }$\mathbf{(a).}$ {% Assume now that} $\mathbf{(a)}$ $p_1^{(i)}\geq 0$ and $q_1^{(i)}\geq 0$ in ${\bar{D}}_{\varepsilon }^{(1)}$. Then, using {\eqref{eq:2p203}}, we find that $U_1^{(i)}\geq 0$. Thus, in both cases $\mathbf{(a)}$\textbf{\ }or $% \mathbf{(b)}$, the inequalities {\eqref{eq:2p201} - \eqref{eq:2p203}} hold. Suppose next that {for some }$m\in \mathbb{N}$% \begin{equation*} {(U_{m}^{(i)}-U_{m-1}^{(i)})\geq 0,\ (p_{m}^{(i)}-p_{m-1}^{(i)})\geq 0,\ (q_{m}^{(i)}-q_{m-1}^{(i)})\geq 0}\text{ \ \ in}\ {\bar{D}}_{\varepsilon }^{(1)},\ i=1,2. \end{equation*} Then \begin{align*} E_{m}^{(i)}-E_{m-1}^{(i)} =&A_{i}{(p_{m}^{(i)}-p_{m-1}^{(i)})}+B_{i}{(q_{m}^{(i)}-q_{m-1}^{(i)})} +C_{i}{(U_{m}^{(i)}-U_{m-1}^{(i)})} \\ &{+D_{i}{(U_{m}^{(i+1)}-U_{m-1}^{(i+1)})}\geq 0} \quad\text{in } {\bar{D}}_{\varepsilon }^{(1)},\quad i=1,2, \end{align*} where we denote ${{U_{m}^{(3)}:=U_{m}^{(1)}}}$. Therefore, we see that \begin{align*} (U_{m+1}^{(i)}-U_{m}^{(i)})(\xi_0,\eta_0) =&\int_0^{\xi_0}\int_{\xi_0}^{\eta_0}(E_{m}^{(i)}-E_{m-1}^{(i)})(\xi ,\eta )\,d\eta \,d\xi \\ &+2\int_0^{\xi_0}\int_0^{\eta }(E_{m}^{(i)}-E_{m-1}^{(i)})(\xi ,\eta )\,d\xi \,d\eta \\ &+\int_0^{\xi_0}\alpha (1-\xi )({U_{m}^{(i)}-U_{m-1}^{(i)}})(\xi ,\xi )\,d\xi \geq 0. \end{align*} In the same manner, \begin{align*} (p_{m+1}^{(i)}-p_{m}^{(i)})(\xi_0,\eta_0) =&\int_0^{\xi_0}(E_{m}^{(i)}-E_{m-1}^{(i)})(\xi ,\xi_0)\,d\xi +\int_{\xi_0}^{\eta_0}(E_{m}^{(i)}-E_{m-1}^{(i)})(\xi_0,\eta )\,\,d\eta \\ &+\alpha (1-\xi_0)({U_{m}^{(i)}-U_{m-1}^{(i)}})(\xi _0,\xi_0)\geq 0, \end{align*} \[ (q_{m+1}^{(i)}-q_{m}^{(i)})(\xi_0,\eta_0)=\int_0^{\xi _0}(E_{m}^{(i)}-E_{m-1}^{(i)})(\xi ,\eta_0)\,d\xi \geq 0. \] Finally, by induction, we conclude that \begin{gather*} U^{(i)}(\xi_0,\eta_0)=\sum_{m=0}^{\infty }(U_{m+1}^{(i)}-U_{m}^{(i)})(\xi_0,\eta_0)\geq 0, \\ p^{(i)}(\xi_0,\eta_0)\geq 0,\quad q^{(i)}(\xi_0,\eta_0)\geq 0,\quad (\xi_0,\eta_0)\in {\bar{D}}_{\varepsilon }^{(1)}. \end{gather*} \hfill$\square$ \begin{remark} \label{rmk3.1} \rm Note here that Lemma \ref{lm3.1}, which is actually a maximum principle for Problem $% P_{\alpha ,3}$, describes the behavior of the system {\eqref{eq:1p21} around the point }$(1,1)$. Thus, this lemma becomes particularly useful in Sections 6 and 7 in finding singular solutions of the equation {\eqref{eq:1p1}. None the less, when the equation \eqref{eq:1p1}} transforms to the system \eqref{eq:1p18}, by \eqref{eq:1p101}, we see that $D_2=-D_1=na_2/4$. Since, in view of Lemma \ref{lm3.1}, $D_1\geq 0$ and $D_2\geq 0$, it should be $% a_2\equiv 0$. Because of this fact, we are able to find singular solutions only when $a_2\equiv 0$ (see also Introduction, Open Questions, 3). \end{remark} As a consequence of Theorem \ref{thm3.1} and representations \eqref{eq:2p91}--\eqref{eq:2p921}, we have the following smoothness result: \begin{theorem} \label{thm3.2} Let $F^{i},A_{i},B_{i},C_{i},D_{i}\in C^{1}({\bar{D}}_{\varepsilon }^{(1)})$, $i=1,2$, $\varepsilon >0$. Then there exists a classical solution $U\in C^2({\bar{D}}_{\varepsilon }^{(1)})$ of Problem $P_{\alpha ,3}$. \end{theorem} \noindent\textbf{Proof.} Since we have already shown that \begin{equation} p_{\eta }^{(1)}(\xi_0,\eta_0)\equiv q_{\xi }^{(1)}(\xi_0,\eta_0) =[F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi_0,\eta_0) \label{eq:2p93} \end{equation} and that similar representations for $p_{\eta }^{(2)}$ and $q_{\xi }^{(2)}$ hold, we have to prove only the fact that $p_{\xi }^{(i)}$ and $q_{\eta }^{(i)}$ exist and belong to $C({\bar{D}}_{\varepsilon }^{(1)})$. Indeed, to do this, we observe the following: 1. For fixed $\eta_0$ the equality {\eqref{eq:2p93}} is a linear ODE for the function $q^{(1)}(\xi_0,\eta_0)$. So, using the well known formula for the solution with the initial Cauchy data $q^{(1)}(0,\eta_0)=0$ from {% \eqref{eq:2p92}, we find that} \begin{equation} q^{(1)}(\xi_0,\eta_0)= \int_0^{\xi_0}[F^{1}+A_1p^{(1)}+C_1U^{(1)}+D_1U^{(2)}](\xi ,\eta_0)\exp \Big( \int_{\xi }^{\xi_0}B_1(\tau ,\eta_0)d\tau \Big) d\xi . \label{eq:2p94} \end{equation} Since $F^{1},A_1,B_1,C_1,D_1,U^{(1)},U^{(2)}\in C^{1}({\bar{D}}% _{\varepsilon }^{(1)})$ and $p^{(1)},p_{\eta }^{(1)}\in C({\bar{D}}% _{\varepsilon }^{(1)})$, by {\eqref{eq:2p94}, }we conclude that $q^{(1)}\in C^{1}({\bar{D}}_{\varepsilon }^{(1)})${.} 2. For fixed $\xi_0$ the equality {\eqref{eq:2p93}} is a linear ODE for the function $p^{(1)}(\xi_0,\eta_0)$. So, arguments similar to those above lead to \begin{equation} \label{eq:2p95} \begin{aligned} p^{(1)}(\xi_0,\eta_0)=& G_1(\xi_0)\exp \Big(\int_{\xi_0}^{\eta_0}A_1(\xi_0,\eta )d\eta \Big)\\ &+\int_{\xi_0}^{\eta_0}\left[ F^{1}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}\right] (\xi_0,\eta )\\ &\times\exp\Big( \int_{\eta }^{\eta_0}A_1(\xi_0,\tau )d\tau \Big) d\eta . \end{aligned} \end{equation} The function $G_1(\xi_0)$, which is defined implicitly by \eqref{eq:2p91}, is of the form \begin{align*} &G_1(\xi_0)\\ &=\int_0^{\xi_0}\big[ F^{1}+A_1p^{(1)}+B_1q^{(1)}+C_1U^{(1)}+D_1U^{(2)}\big] (\xi ,\xi _0)d\xi +\alpha (1-\xi_0)U^{(1)}(\xi_0,\xi_0). \end{align*} Obviously $G_1\in C^{1}({\bar{D}}_{\varepsilon }^{(1)})$, because $F^{1}$, $A_1$, $B_1$, $C_1$, $D_1$, $\alpha $, $U^{(1)}$, $U^{(2)}$, $q^{(1)}\in C^{1}({\bar{D}}_{\varepsilon }^{(1)})$ and $p^{(1)},p_{\eta}^{(1)}\in C({\bar{D}}_{\varepsilon }^{(1)})$. Finally, by \eqref{eq:2p95}, we see that $p^{(1)}\in C^{1}({\bar{D}}_{\varepsilon }^{(1)})$. \hfill$\square$ \begin{remark} \label{rmk3.2} \rm By studying a solution of the Problem $P_{\alpha ,3}$ in the domain ${\bar{D}}_{\varepsilon }^{(1)}$, we are actually investigating the behavior of the solution of Problem $P_{\alpha ,2}$ in the domain ${\bar{D}}_{\delta }$, when $\delta \to 0$, around the line $\eta =1$. It is easy to show that, for any $\varepsilon \in (0,1)$ and $\delta \in (0,1)$, the solutions of these two problems coincide in their common domain ${\bar{D}}_{\varepsilon }^{(1)}\cap {\bar{D}}_{\delta }$. \end{remark} \section{Existence and uniqueness theorems for the 2-D Problem $\mathbf{P_{\alpha ,1}}$} %4 Consider the 2-D problem $P_{\alpha ,1}:$ \begin{equation} \begin{gathered} \frac{1}{\varrho }(\varrho u_{\varrho }^{(1)})_{\varrho }-u_{tt}^{(1)}+a_1u_{_{\varrho }}^{(1)}+bu_{t}^{(1)}+(c-\frac{n^2}{% \varrho ^2})u^{(1)}+na_2u^{(2)}=f^{(1)}\text{ in }G_{\varepsilon }, \\ \frac{1}{\varrho }(\varrho u_{\varrho }^{(2)})_{\varrho }-u_{tt}^{(2)}+a_1u_{\varrho }^{(2)}+bu_{t}^{(2)}+(c-\frac{n^2}{\varrho ^2})u^{(2)}-na_2u^{(1)}=f^{(2)}\text{ in }G_{\varepsilon }, \\ u^{(i)}\bigr\rvert_{S_1\cap \partial G_{\varepsilon }}=0,\quad \lbrack u_{t}^{(i)}+\alpha (\varrho )u^{(i)}]\bigr\rvert_{S_0\cap \partial G_{\varepsilon }}=0,\quad i=1,2. \end{gathered} \label{eq:3p1} \end{equation} Note that, the generalized solution of the problem $P_{\alpha ,1}$ in the domain $G_{\varepsilon }$ , $\varepsilon \in (0,1)$, was defined by Definition \ref{def2.2}. \begin{theorem} \label{thm4.1} Let $a_1,a_2,b,c,f^{(1)},f^{(2)}\in C^{1}(\bar{G}_0\setminus (0,0))$. Then there exists a generalized solution $u=(u^{(1)},u^{(2)})\in C^2(\bar{G}_0\setminus (0,0))$ of problem $P_{\alpha ,1}$ in $G_0$, which is a classical solution of the problem $P_{\alpha ,1}$ in any domain $G_{\varepsilon }$, $\varepsilon \in (0,1)$. \end{theorem} \noindent\textbf{Proof.} In view of \eqref{eq:1p15} and \eqref{eq:1p17}, i.e. $z(\varrho ,t)=\varrho ^{1/2}u(\varrho ,t)$ and $\xi =1-\varrho -t$, $\eta =1-\varrho +t$, we introduce the function \begin{equation*} U^{(i)}(\xi,\eta)=z^{(i)}(\varrho(\xi,\eta),t(\xi ,\eta )). \end{equation*} Then Problem $P_{\alpha ,1}$, in the new terms, becomes $P_{\alpha ,2}$, i.e. \begin{equation} \begin{gathered} U_{\xi \eta }^{(1)}-A_1U_{\xi }^{(1)}-B_1U_{\eta }^{(1)}-C_1U^{(1)}-D_1U^{(2)}=F^{1}(\xi ,\eta )\text{ in }D_{\varepsilon }, \\ U_{\xi \eta }^{(2)}-A_2U_{\xi }^{(2)}-B_2U_{\eta }^{(2)}-C_2U^{(2)}-D_2U^{(1)}=F^2(\xi ,\eta )\text{ in }D_{\varepsilon }, \end{gathered} \label{eq:3p5} \end{equation} \begin{equation} U^{(i)}(0,\eta )=0,(U_{\eta }^{(i)}-U_{\xi }^{(i)})(\xi ,\xi )+\alpha (1-\xi )U^{(i)}(\xi ,\xi )=0,\quad i=1,2, \label{eq:3p3} \end{equation} where the connection between the coefficients is given by \eqref{eq:1p101}. For each fixed $\varepsilon \in (0,1)$ Theorem \ref{thm3.2} ensures the existence of a classical solution $(U^{(1)},U^{(2)})\in C^2(\bar{D}_{\varepsilon }^{(1)})$ of the problem $P_{\alpha ,3}$. More precisely, for any fixed $% \varepsilon_1,\varepsilon_2$ with $0<\varepsilon_1<\varepsilon _2<1$ the corresponding vector-valued solution $U_{\varepsilon_2}$ is a restriction of $U_{\varepsilon_1}$ in the region $D_{\varepsilon_2}$. So, essentially we have a function of class $C^2\left( \overset{\_}{D}% _0\backslash (0,0)\right) $, which in any region $D_{\varepsilon }$ coincides with the corresponding solution $U_{\varepsilon }$ and is a classical solution of Problem $P_{\alpha ,3}$. We remark that the inverse transformations {\eqref{eq:1p15} and \eqref{eq:1p17} }lead to a vector-valued function $(u^{(1)},u^{(2)})\in C^2\left( \bar{G}% _0\backslash (0,0)\right) $, which is a classical solution of Problem $% P_{\alpha ,1}$\ in each $G_{\varepsilon }$. This solution is also a generalized solution of the same problem in $G_0$, because for each concrete test function $v\in $\ $V_0$ there is an $\varepsilon_{v}>0$ for which $v\equiv 0$ in $G_0\backslash G_{\varepsilon_{v}}$ and {% \eqref{eq:0p7}} coincides with \eqref{eq:1p6}. The proof of the theorem is complete. \hfill$\square$ \begin{theorem} \label{thm4.2} Let $a_1,a_2,b,c\in C^{1}(\bar{G}_0\setminus (0,0))$. Then for each fixed $\varepsilon \in (0,1)$ there exists at most one generalized solution of the problem $P_{\alpha ,1}$ in $G_{\varepsilon }$. \end{theorem} \noindent\textbf{Proof.} Let $(u_1^{(1)},u_1^{(2)})$ and $(u_2^{(1)},u_2^{(2)})$ be two generalized solutions of $P_{\alpha ,1}$ in $G_{\varepsilon}$. Then for $u^{(i)}:=u_1^{(i)}-u_2^{(i)}$, $i=1,2$, we see that \begin{enumerate} \item $u^{(i)}\in C^{1}({\bar{G}}_{\varepsilon })$, $[u_{t}^{(i)}+\alpha (\varrho )u^{(i)}]\bigr\rvert_{S_0\cap \partial G_{\varepsilon }}=0$, $u^{(i)}\bigr\rvert_{S_1\cap \partial G_{\varepsilon }}=0$, $i=1,2$; \item The equalities \begin{equation} \begin{aligned} &\int_{G_{\varepsilon }}\Big[ u_{t}^{(1)}v_{t}^{(1)}-u_{% \varrho }^{(1)}v_{\varrho }^{(1)}+\big( a_1u_{\varrho }^{(1)}+bu_{t}^{(1)}+(c-\frac{n^2}{\varrho ^2})u^{(1)}+na_2u^{(2)}% \big) v^{(1)}\Big] \varrho d\varrho \,dt \\ &=\int_{S_0\cap \partial G_{\varepsilon }}\alpha (\varrho )u^{(1)}v^{(1)}\varrho \,d\varrho , \\ &\int_{G_{\varepsilon }}\Big[ u_{t}^{(2)}v_{t}^{(2)} -u_{\varrho }^{(2)}v_{\varrho }^{(2)}+\big( a_1u_{\varrho }^{(2)}+bu_{t}^{(2)}+(c-\frac{n^2}{\varrho ^2})u^{(2)}-na_2u^{(1)} \big) v^{(2)}\Big] \varrho d\varrho \,dt \\ &=\int_{S_0\cap \partial G_{\varepsilon }}\alpha (\varrho )u^{(2)}v^{(2)}\varrho \,d\varrho \end{aligned} \label{eq:3p101} \end{equation} hold for all {functions }$v^{(1)},v^{(2)}\in V_{\varepsilon }^{(1)}$. \end{enumerate} If the functions $v^{(i)}\in C^2({\bar{G}}_{\varepsilon })$, then from \eqref{eq:3p101} we conclude that \begin{equation} \begin{aligned} \int_{G_{\varepsilon }}\Big[ \big( \frac{1}{\varrho } (\rho v_{\varrho }^{(1)})_{\varrho }-v_{tt}^{(1)}-\frac{1}{\varrho }(\varrho a_1v^{(1)})_{\varrho }-(bv^{(1)})_{t}&\\ +(c-\frac{n^2}{\varrho ^2})v^{(1)}\big) u^{(1)} +na_2v^{(1)}u^{(2)}\Big] \varrho d\varrho \,dt&=0, \\ \int_{G_{\varepsilon }}\Big[ \big( \frac{1}{\rho } (\rho v_{\varrho }^{(2)})_{\varrho }-v_{tt}^{(2)}-\frac{1}{\varrho }(\varrho a_1v^{(2)})_{\varrho }-(bv^{(2)})_{t} &\\ +(c-\frac{n^2}{\varrho ^2})v^{(2)}\big) u^{(2)} -na_2v^{(2)}u^{(1)}\Big] \varrho d\varrho \,dt&=0. \end{aligned} \label{eq:3p120} \end{equation} For $h^{(1)},h^{(2)}\in C^{1}\left( \bar{G}_0\backslash (0,0)\right) $ we state the following problem. \textbf{Problem }$P_{\alpha ,1}^{\ast }$. Find a solution $v^{(1)},v^{(2)}\in V_{\varepsilon }^{(1)}\cap C^2({\bar{G}}_{\varepsilon }) $ of the system \begin{gather*} \frac{1}{\varrho }(\rho v_{\varrho }^{(1)})_{\varrho }-v_{tt}^{(1)}-\frac{1}{\varrho }(\varrho a_1v^{(1)})_{\varrho }-(bv^{(1)})_{t}+(c-\frac{n^2}{\varrho ^2})v^{(1)}-na_2v^{(2)}=h^{(1)}, \\ \frac{1}{\varrho }(\varrho v_{\varrho }^{(2)})_{\varrho }-v_{tt}^{(2)}-\frac{1}{\varrho }(\varrho a_1v^{(2)})_{\varrho }-(bv^{(2)})_{t}+(c-\frac{n^2}{\varrho ^2})v^{(2)}+na_2v^{(1)}=h^{(2)}. \end{gather*} For $z^{(i)}=\varrho ^{1/2}v^{(i)}$, $\xi_1=1-\varepsilon -\eta $, $\eta _1=1-\varepsilon -\xi $, and \begin{equation} V^{(i)}(\xi_1,\eta_1)=z^{(i)}(1-\varepsilon -\eta_1,1-\varepsilon -\xi_1), \end{equation} the domain $G_{\varepsilon }$ maps into $D_{\varepsilon }$, and for appropriate coefficients $A_{i},B_{i},C_{i},D_{i}$ and $\beta =\alpha +b$ the above Problem $P_{\alpha ,1}^{\ast }$ transforms to the Darboux--Goursat Problem $P_{\beta ,3}$. But for this problem Theorem \ref{thm3.2} ensures the solvability in $C^2({\bar{D}}_{\varepsilon })$. Consequently, there exists a classical solution $(V^{(1)},V^{(2)})\in C^2({\bar{D}}_{\varepsilon })$ and so the inverse transformations\emph{\ }{\eqref{eq:1p15} and % \eqref{eq:1p17} }lead to a classical solution $(v^{(1)},v^{(2)})\in C^2(% \bar{G}_{\varepsilon })$ of Problem $P_{\alpha ,1}^{\ast }$. Moreover, with these functions $v^{(1)},v^{(2)}$ the system \eqref{eq:3p120} becomes \begin{equation} \begin{gathered} \int_{G_{\varepsilon }}\left[ \left( h^{(1)}+na_2v^{(2)}\right) u^{(1)}+na_2v^{(1)}u^{(2)}\right] \varrho d\varrho \,dt=0, \\ \int_{G_{\varepsilon }}\left[ \left( h^{(2)}-na_2v^{(1)}\right) u^{(2)}-na_2v^{(2)}u^{(1)}\right] \varrho d\varrho \,dt=0. \end{gathered} \label{eq:3p102} \end{equation} Since the functions $h^{(1)}(\varrho ,t),h^{(2)}(\varrho ,t)\in C^{1} \left( {\bar{G}}_0\backslash (0,0)\right) $ are arbitrary, \eqref{eq:3p102} gives $u^{(1)}(\varrho ,t)=u^{(2)}(\varrho ,t)=0$ in $G_{\varepsilon }$, i.e. $(u_1^{(1)},u_1^{(2)})\equiv (u_2^{(1)},u_2^{(2)})$. The proof is complete. \hfill$\square$ \section{Existence and uniqueness theorems for the 3-D Problem $P_{\protect\alpha }$} %5 In this section we consider the following 3-D boundary value problem. \noindent\textbf{Problem $P_{\alpha }$.} Find a solution to the equation \begin{equation} Lu=\frac{1}{\varrho }(\varrho u_{\varrho })_{\varrho }+\frac{1}{\varrho ^2}% u_{\varphi \varphi }-u_{tt}+a_1u_{\varrho }+a_2u_{\varphi }+bu_{t}+cu=f(\varrho ,\varphi ,t)\text{ in }\Omega_{\varepsilon }, \label{eq:4p1} \end{equation} which satisfies the boundary conditions \begin{equation} \quad u\bigr\rvert_{\Sigma_1\cap \partial \Omega_{\varepsilon }}=0,\quad \lbrack u_{t}+\alpha (\varrho )u]\bigr\rvert_{\Sigma_0\cap \partial \Omega_{\varepsilon }}=0. \label{eq:4p2} \end{equation} For this problem we formulate the following theorems. \begin{theorem} \label{thm5.1} Let $a_1,a_2,b,c\in C^{1}(\bar{\Omega}_0\backslash O)$. Then for $0\leq \varepsilon <1$ there exists at most one generalized solution of Problem $P_{\alpha }$ in\ $\Omega_{\varepsilon }$. \end{theorem} \noindent\textbf{Proof.} \textit{Case} $0<\varepsilon <1$. If $u_1,u_2$ are two generalized solutions of $P_{\alpha }${\ in\ $\Omega_{\varepsilon }$, then $u:=u_1-u_2\in C^{1}(\bar{\Omega}_{\varepsilon })$} satisfies \eqref{eq:4p2} and \begin{equation} \begin{aligned} &\int_{\Omega_{\varepsilon }}[u_{t}v_{t}-u_{\varrho }v_{\varrho }-\frac{1}{\varrho ^2}u_{\varphi }v_{\varphi }+(a_1u_{\varrho }+a_2u_{\varphi }+bu_{t}+cu)v]\varrho \,d\varrho \,d\varphi \,dt\\ &=\int_{\Sigma_0\cap \partial \Omega_{\varepsilon }}\alpha (\varrho )uv\varrho \,d\varrho d\varphi \end{aligned}\label{eq:4p3} \end{equation} holds for all $v\in V_{\varepsilon }$. We will show that in the Fourier expansion\ \begin{equation} u(\varrho ,\varphi ,t)=\sum_{n=0}^{\infty }\left\{ u_{n}^{(1)}(\varrho ,t)\cos n\varphi +u_{n}^{(2)}(\varrho ,t)\sin n\varphi \right\} \label{eq:4p4a} \end{equation} the coefficients satisfy $u_{n}^{(i)}(\varrho ,t)$ $\equiv 0$ in $\Omega_{\varepsilon }$, $i=1,2$, i.e. $u\equiv 0$ in $\Omega_{\varepsilon}$. Since $u\in C^{1}(\bar{\Omega}_{\varepsilon })$, using the substitution \begin{equation*} v_1(\varrho ,\varphi ,t)=w_1(\varrho ,t)\cos n\varphi \in V_{\varepsilon }\text{ \ \ \ or \ \ \ }v_2(\varrho ,\varphi ,t)=w_2(\varrho ,t)\sin n\varphi \in V_{\varepsilon } \end{equation*} in \eqref{eq:4p3}, we derive the system \begin{equation} \begin{aligned} &\int_{G_{\varepsilon }}\Big[ u_{n,t}^{(1)}w_{1,t}-u_{n, \varrho }^{(1)}w_{1,\varrho }+\big( a_1u_{n,\varrho }^{(1)}+bu_{n,t}^{(1)}+(c-\frac{n^2}{\varrho ^2} )u_{n}^{(1)}+na_2u_{n}^{(2)}\big) w_1\Big] \varrho d\varrho \,dt \\ &=\int_{S_0\cap \partial G_{\varepsilon }}\alpha (\varrho )u_{n}^{(1)}w_1\varrho \,d\varrho ,\\ &\int_{G_{\varepsilon }}\left[ u_{n,t}^{(2)}w_{2,t}-u_{n, \varrho }^{(2)}w_{2,\varrho }+\left( a_1u_{n,\varrho }^{(2)}+bu_{n,t}^{(2)}+(c-\frac{n^2}{\varrho ^2} )u_{n}^{(2)}-na_2u_{n}^{(1)}\right) w_2\right] \varrho \,d\varrho dt \\ &=\int_{S_0\cap \partial G_{\varepsilon }}\alpha (\varrho )u_{n}^{(2)}w_2\varrho \,d\varrho \end{aligned} \label{eq:4p5} \end{equation} for all $w_1,w_2\in V_{\varepsilon }^{(1)}$ and $n\in \mathbb{N}\cup \{0\}$. By Definition \ref{def2.2}, the function $(u_{n}^{(1)},u_{n}^{(2)})(\varrho ,t)$ is a generalized solution of the homogeneous problem $P_{\alpha ,1}$. Clearly, Theorem \ref{thm4.2} implies $u_{n}^{(1)}(\varrho ,t)\equiv u_{n}^{(2)}(\varrho ,t)\equiv 0$ in $\Omega_{\varepsilon }$ for $n\in \mathbb{N}\cup \{0\}$ and so $u^{(1)}=u_1-u_2\equiv 0$ in $\Omega _{\varepsilon }$. {\textit{Case }}$\varepsilon =0$. Let $\varepsilon_0$ be an arbitrary fixed number of $(0,1)$. Then, by Lemma \ref{lm2.1}, it follows that the generalized solution $u\in C^{1}(\bar{\Omega}_0\setminus (0,0,0))$ of Problem $% P_{\alpha }$ in $\Omega_0$ is also a generalized solution of the homogeneous problem $P_{\alpha }$ in $\Omega_{\varepsilon_0}$. Since, by the previous case, $u\equiv 0$ in $\Omega_{\varepsilon_0}$ and $% \varepsilon_0>0$ is arbitrary, we see that $u=u_1-u_2\equiv 0$ in $% \Omega_0$. This completes the proof of the theorem.\hfill $\square$ \begin{theorem} \label{thm5.2} Let $a_1,a_2,b,c\in C^{1}\left( \bar{\Omega}_0\backslash O\right) $ and the function $f\in C(\overset{-}{\Omega }_0)\cap C^{1}( \overset{-}{\Omega }_0\backslash O) $ be of the form \begin{equation} f(\varrho ,\varphi ,t)=\sum_{n=0}^{k}\big\{ f_{n}^{(1)}(\varrho ,t)\cos n\varphi +f_{n}^{(2)}(\varrho ,t)\sin n\varphi \big\} ,\; k\in \mathbb{N}\cup \{0\}. \label{eq:4p6} \end{equation} Then there exists one and only one generalized solution \begin{equation} u(\varrho ,\varphi ,t)=\sum_{n=0}^{k}\left\{ u_{n}^{(1)}(\varrho ,t)\cos n\varphi +u_{n}^{(2)}(\varrho ,t)\sin n\varphi \right\} \label{eq:4p7} \end{equation} of the problem $P_{\alpha }$ in $\Omega_0$. This solution $u\in C^2\left( \bar{\Omega}_0\backslash O\right) $ is a classical solution of the problem $P_{\alpha }$ in each domain $\Omega_{\varepsilon },\varepsilon \in (0,1)$. Moreover, if \begin{equation*} |a_1|\leq d\varrho ^{-1},\quad |a_2|\leq d\varrho ^{-2}, \quad |b|\leq d\varrho^{-2},\quad |c|\leq d\varrho ^{-2}, \quad |\alpha |\leq d\varrho ^{-2}\quad\text{in }\bar{\Omega}_0\backslash O, \end{equation*} then, in view of \eqref{eq:4p7}, for a fixed $n$, the corresponding trigonometric polynomial $u_{n}$ of degree $n$, satisfies the following a priori estimates: For $n=0$, \begin{equation} \| u_0(x_1,x_2,t)\|_{C^{1}(\Omega_{\varepsilon }^{(1)})} =\sum_{|\alpha |\leq 1}{\sup_{\Omega_{\varepsilon }^{(1)}}} |D^{\alpha }u_0| \leq 6\varepsilon ^{3/2} \exp\big( \frac{32d+2}{\varepsilon ^2}\big) \| f_0^{(1)} \|_{C^0(\bar{G}_0)}; \label{eq:4p8} \end{equation} while for $n\in \mathbb{N}$, \begin{equation} \| u_{n}(x_1,x_2,t)\|_{C^{1}(\Omega_{\varepsilon }^{(1)})} \\ \leq \frac{6\varepsilon ^{3/2}}{n(n+2d)}\exp \big( \frac{8n(n+3d)}{\varepsilon ^2}\big) \big( \| f_{n}^{(1)}\|_{C^0(\bar{G}_0)}+\| f_{n}^{(2)} \|_{C^0(\bar{G}_0)}\big) , \label{eq:4p9} \end{equation} where $\Omega_{\varepsilon }^{(1)}\ =\Omega_0\cap \{(\varrho ,t):\varrho +t>\varepsilon \}$. \end{theorem} \noindent\textbf{Proof.} It suffices to consider the case of a fixed number $n$. As in Section 2, we make the substitutions \begin{equation} \xi =1-\varrho -t,\quad \eta =1-\varrho +t, \label{eq:4p11a} \end{equation} and introduce the new function \begin{equation} U^{(i)}(\xi ,\eta )=\varrho ^{1/2}u^{(i)}(\varrho (\xi ,\eta ),t(\xi ,\eta )). \label{eq:4p11} \end{equation} Denote \begin{equation*} F^{(i)}(\xi ,\eta ):=\frac{1}{4\sqrt{2}}(2-\eta -\xi )^{1/2}f_{n}^{(i)}(\xi ,\eta )\in C^{1}\left( \bar{D}_0\backslash (1,1)\right) , \end{equation*} and use the notation of \eqref{eq:1p101}. Then the problem reduces to Problem $P_{\alpha ,3}$. Thus, we can use Theorems \ref{thm3.1} and \ref{thm3.2} to ensure the existence of a classical solution $(U^{(1)},U^{(2)})(\xi ,\eta )$ of this problem with the estimates \eqref{eq:2p7}. \noindent\textit{Case} $n\in \mathbb{N}$. In view of \eqref{eq:2p5}, \eqref{eq:1p101}, it is easy to see that we can chose \begin{gather*} c(\varepsilon ):= \frac{n(n+2d)}{\varepsilon ^2},\quad M_{\alpha }:=\frac{4d}{\varepsilon ^2} \\ M \leq \frac{1}{4}\max \left\{ \|f_{n}^{(1)}\|_{C^0(\bar{G}_0)}, \|f_{n}^{(2)}\|_{C^0(\bar{G}_0)}\right\} :=M_{n}, \end{gather*} Hence, Theorems \ref{thm3.1} and \ref{thm3.2} ensure the smoothness of the solution $U$ of Problem $P_{\alpha ,3}$ in $D_{\varepsilon }^{(1)}=\{(\xi ,\eta ):0<\xi <\eta <1,\;0<\xi <1-\varepsilon \},\;\varepsilon >0$, i.e. \begin{equation} (U_{n}^{(1)},U_{n}^{(2)})(\xi ,\eta ):=U(\xi ,\eta )\in C^2(\bar{D}% _{\varepsilon }^{(1)})\,. \label{eq:4p14} \end{equation} On the other hand, these theorems ensure the a priori estimates \begin{align*} \sup_{D_{\varepsilon }^{(1)}}|U_{n}^{(i)}(\xi ,\eta )| &\leq M_{n}\left[ 4c(\varepsilon )+M_{\alpha }\right] ^{-2}\exp \left\{ 8c(\varepsilon )+2M_{\alpha }\right\} \\ &\leq M_{n}\varepsilon ^{4}[4n(n+2d)]^{-2}\exp \left\{ (8n(n+3d)\varepsilon ^{-2}\right\} ,\\ \sup_{D_{\varepsilon }^{(1)}}\{|U_{n,\xi }^{(i)}|,\;|U_{n,\eta}^{(i)}|\} &\leq M_{n}\varepsilon ^2[4n(n+2d)]^{-1}\exp \left\{ (8n(n+3d)\varepsilon ^{-2}\right\} . \end{align*} Also, by \eqref{eq:4p11a} and \eqref{eq:4p11}, we have $u_{n}^{(i)}(\varrho ,t)=\varrho ^{-\frac{1}{2}}U_{n}^{(i)}(\xi ,\eta )$. Since $\varrho \geq \varepsilon /2$ for $(\xi ,\eta )\in D_{\varepsilon }^{(1)}$, by the inverse transformation we see that \begin{equation} \begin{gathered} |u_{n}^{(i)}(\varrho ,t)| \leq M_{n}\frac{\varepsilon ^{7/2}}{8n^2(n+2d)^2}\exp \big( \frac{8n(n+3d)}{\varepsilon ^2} \big) , \\ |u_{n,t}^{(i)}(\varrho ,t)| \leq M_{n}\frac{\varepsilon ^{3/2}}{n(n+2d)}\exp \big( \frac{8n(n+3d)}{\varepsilon ^2}\big) , \\ |u_{n,\varrho }^{(i)}(\varrho ,t)| \leq M_{n}\frac{ \varepsilon ^{3/2}}{n(n+2d)}\exp \big( \frac{8n(n+3d)}{\varepsilon ^2} \big) . \end{gathered} \label{eq:4p16} \end{equation} Therefore, in view of \eqref{eq:4p7} and \eqref{eq:4p16}, for the trigonometrical polynomial \begin{equation} u_{n}(\varrho ,\varphi ,t)=u_{n}^{(1)}(\varrho ,t)\cos n\varphi +u_{n}^{(2)}(\varrho ,t)\sin n\varphi \label{eq:4p201} \end{equation} we derive \begin{equation} \| \frac{1}{\varrho }u_{n,\varphi }(\varrho ,\varphi ,t)\|_{C(\Omega_{\varepsilon }^{(1)})}\leq M_{n}\frac{% \varepsilon ^{5/2}}{4n(n+2d)^2}\exp \left( \frac{8n(n+3d)}{\varepsilon ^2% }\right) . \label{eq:5p20} \end{equation} Since $ u_{n}(\varrho \cos \varphi ,\varrho \sin \varphi ,t)=u_{n}^{(1)}(\varrho ,\varphi ,t)$, obviously one has \begin{equation*} |u_{n,x_{i}}(x_1,x_2,t)|\leq 2M_{n}\frac{\varepsilon ^{3/2}}{n(n+2d)}% \exp \left( \frac{8n(n+3d)}{\varepsilon ^2}\right) ,\quad i=1,2. \end{equation*} So, the estimate \eqref{eq:4p9} holds in $\Omega_{\varepsilon }^{(1)}$. \noindent\textit{Case $n=0$.} By \eqref{eq:4p6} and \eqref{eq:4p7}, we have $f_0(\varrho ,\varphi ,t)=f_0^{(1)}(\varrho ,t)$ and $u_0(x_1,x_2,t)=u_0(\varrho ,\varphi ,t)=u_0^{(1)}(\varrho ,t)$. Take \begin{equation*} c(\varepsilon ):=\frac{8d+1}{4\varepsilon ^2},\quad M_{\alpha }:=\frac{4d}{\varepsilon ^2},\quad M:=\frac{1}{4}\| f_0^{(1)}\|_{C^0(\bar{G}_0)}\,. \end{equation*} Then, as in the previous case, we obtain \eqref{eq:4p8}. \hfill$\square$ \begin{theorem} \label{thm5.3} Let the conditions of Theorem \ref{thm5.2} be fulfilled. Also, for the sake of simplicity, suppose that $a_1,a_2,b,c\in C^{1}(\bar{\Omega}_0)$ and $% |\alpha ^{\prime }(\varrho )|\leq d_1/\varrho ^{-3}$. Then for a fixed $% n\in \mathbb{N}$ \ the corresponding trigonometric polynomial $u_{n}$ of degree $n$ from \eqref{eq:4p201} satisfies the following a priori estimate \begin{equation} \| u_{n}(x_1,x_2,t)\|_{C^2(\bar{\Omega}_{\varepsilon}^{(1)})} \leq C_1\varepsilon ^{-1/2}\exp \big( \frac{8n(n+3d)}{\varepsilon ^2}\big) \big( \| f_{n}^{(1)}\|_{C^0(\bar{G}_0)}+\| f_{n}^{(2)}\|_{C^0(\bar{G}_0)}\big) , \end{equation} where the constant $C_1$ does not depend on $n$ and $\varepsilon $. \end{theorem} \noindent\textbf{Proof.} We will use the estimates of Theorem \ref{thm5.2} and the representations of the second derivatives of Theorem \ref{thm3.2}. Following the same arguments, as in Theorem \ref{thm5.2}, we obtain the estimates \begin{equation*} \sup_{D_{\varepsilon }^{(1)}}\{|U_{n,\xi \eta }^{(i)}|,|U_{n,\eta \eta }^{(i)}|,\;|U_{n,\xi \xi }^{(i)}|\}\leq C_1M_{n}\exp \left\{ 8n(n+3d)\varepsilon ^{-2}\right\} ,\text{ }i=1,2 \end{equation*} and conclude that \begin{equation*} \sup_{D_{\varepsilon }^{(1)}}\{|u_{n,x_{i}x_{j}}|,|u_{n,tx_{i}}|\}\leq C_1M_{n}\varepsilon ^{-1/2}\exp \big( \frac{8n(n+3d)}{\varepsilon ^2}\big). \end{equation*} \hfill$\square$ The next theorem is an immediate consequence of Theorems \ref{thm5.1}, \ref{thm5.2} and \ref{thm5.3}. \begin{theorem} \label{thm5.4} Let the conditions of Theorem \ref{thm5.3} be fulfilled and let $f\in C^{1}(\bar{\Omega}_0)$ be of the form \begin{equation} f(\varrho ,\varphi ,t)=\sum_{n=0}^{\infty }\{f_{n}^{(1)}(\varrho ,t)\cos n\varphi +f_{n}^{(2)}(\varrho ,t)\sin n\varphi \}. \label{eq:4p17} \end{equation} Suppose that the Fourier coefficients $f_{n}^{(1)}(\varrho ,t)$ and $% f_{n}^{(2)}(\varrho ,t)$ satisfy \begin{equation} \begin{aligned} \| f\|_{\exp \,(\varepsilon )} :=&\exp \big( \frac{32d+2}{\varepsilon ^2}\big) \| f_0^{(1)}\|_{C^0(\bar{G}_0)} +\sum_{n=1}^{\infty }\frac{1}{n(n+2d)}\exp \big(\frac{8n(n+3d)}{\varepsilon ^2}\big)\\ &\times\big( \| f_{n}^{(1)}\|_{C^0(\bar{G}_0)}+\| f_{n}^{(2)} \|_{C^0(\bar{G}_0)}\big) <\infty \,. \end{aligned} \label{eq:4p18} \end{equation} Then there exists one and only one generalized solution $u\in C^{1}\left( \bar{\Omega}_{\varepsilon }^{(1)}\right) $ of the problem $P_{\alpha }$ in $% \Omega_{\varepsilon }$ and satisfies the a priori estimate \begin{equation} \| u\|_{C^{1}(\Omega_{\varepsilon }^{(1)})}\leq 6\varepsilon ^{3/2}\| f\|_{\exp \,(\varepsilon )}. \label{eq:4p19} \end{equation} Moreover, if \begin{equation} \begin{aligned} \| f\|_{\exp_1\,(\varepsilon )}:=&\exp \big( \frac{32d+2}{\varepsilon ^2}\big) \| f_0^{(1)}\|_{C^0(\bar{G}_0)} +\sum_{n=1}^{\infty }\exp \big( \frac{8n(n+3d)}{\varepsilon ^2}\big) \\ &\times \big( \| f_{n}^{(1)}\|_{C^0(\bar{G}_0)} +\| f_{n}^{(2)}\|_{C^0(\bar{G}_0)}\big) <\infty , \end{aligned} \label{eq:4p20} \end{equation} then $u\in C^2\left( \bar{\Omega}_{\varepsilon }^{(1)}\right) $, $u(x,t)$ is a classical solution of the problem $P_{\alpha }$ in $\Omega _{\varepsilon }$ and satisfies the a priori estimate \begin{equation} \| u\|_{C^2(\Omega_{\varepsilon }^{(1)})}\leq C_2\varepsilon ^{-1/2}\| f\|_{\exp_1\,(\varepsilon )}. \label{eq:4p21} \end{equation} \end{theorem} \begin{remark} \label{rmk5.1} \rm It is obvious that the estimates \eqref{eq:4p19} or \eqref{eq:4p21} hold, if the series \eqref{eq:4p18} and \eqref{eq:4p20} are finite. In this case we have a solution, which is of class $C^{1}(\bar{\Omega}_0\backslash O)$ or of class $C^2(\bar{\Omega}_0\backslash O)$. For example, the condition % \eqref{eq:4p20} is valid for each $\varepsilon \in (0,1)$, if there exists a sequence $a_{n}$ $\to +\infty $ as $n\to +\infty $ such that \begin{equation} \sum_{n=1}^{\infty }\exp \left( n^2a_{n}\right) \left( \| f_{n}^{(1)}\|_{C^0(\bar{G}_0)}+\| f_{n}^{(2)}\|_{C^0(\bar{G}% _0)}\right) <\infty . \label{eq:4p22} \end{equation} \end{remark} To show this, it is enough to see that the inequality $n(\varepsilon ^2a_{n}-8)\geq 24d$ holds, for all large enough $n\in \mathbb{N}$. \section{On the singularity of solutions of Problem $P_{\alpha }$} %6 For the the equation \begin{equation} Lu=\frac{1}{\varrho }(\varrho u_{\varrho })_{\varrho }+\frac{1}{\varrho ^2}% u_{\varphi \varphi }-u_{tt}+a_1u_{\varrho }+a_2u_{\varphi }+bu_{t}+cu=f(\varrho ,\varphi ,t)\text{ in }\Omega_0, \label{eq:5p1} \end{equation} we consider the boundary conditions of Problem $P_{\alpha }$, i.e. \begin{equation} P_{\alpha }:\quad u\bigr\rvert_{\Sigma_1}=0,\quad \lbrack u_{t}+\alpha (\varrho )u]\bigr\rvert_{\Sigma_0\backslash O}=0 \label{eq:5p2} \end{equation} and prove the following result. \begin{theorem} \label{thm6.1} Let $\alpha (\varrho )\geq 0;$ $a_1$, $b$, $c\in C^{1}({\bar{\Omega}}_0\backslash O)$, $a_2\equiv 0$ and \begin{equation} a_1(\varrho ,t)\geq |b|(\varrho ,t),\quad a_1(\varrho ,t)\geq 2\varrho c(\varrho ,t),\quad (\varrho ,t)\in {\Omega }_0. \label{eq:5p90} \end{equation} Then for each function \begin{equation*} f_{n}(\varrho ,\varphi ,t)=\varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}\cos n\varphi \in C^{n-2}({\bar{\Omega}}_0)\cap C^{\infty }({\Omega }% _0),\quad n\in \mathbb{N}, \end{equation*} the corresponding generalized solution $u_{n}$ of the problem $P_{\alpha }$ belongs to $C^2({\bar{\Omega}}_0\backslash O)$ and satisfies the estimate \begin{equation} |u_{n}(\varrho ,\varphi ,\varrho )|\geq \frac{1}{2}|u_{n}(2\varrho ,\varphi ,0)|+\varrho ^{-n}|\cos n\varphi |\geq \varrho ^{-n}|\cos n\varphi |,\quad 0<\varrho <1\,. \label{eq:5p3} \end{equation} \end{theorem} \noindent\textbf{Proof.} Note that, by Theorem \ref{thm1.1}, the functions \begin{equation*} w_{n}(\varrho ,\varphi ,t)=\varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}(a_{n}\cos n\varphi +b_{n}\sin n\varphi ),\quad n\geq 4, \end{equation*} are classical solutions of the homogeneous Problem $P_{\alpha }^{\ast }$ for the wave equation, when $\alpha \equiv 0$. Now consider the special case of Problem $P_{\alpha }$: \begin{equation} Lu=f_{n}\equiv \varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}\cos n\varphi \quad\text{in $\Omega_0.$} \label{eq:5p5a} \end{equation} Observe also that \begin{equation*} f_{n}(x_1,x_2,t)=(x_1^2+x_2^2)^{-n}(x_1^2+x_2^2-t^2)^{n-1/2}% \mathop{\rm Re}(x_1+ix_2)^{n} \end{equation*} and obviously $f_{n}\in C^{n-2}({\bar{\Omega}}_0)\cap C^{\infty }({\Omega } _0)$, $n\in \mathbb{N}$. Theorem \ref{thm5.1} states that the equation \eqref{eq:5p5a} with boundary conditions \eqref{eq:5p2} has at most one generalized solution. On the other hand, from Theorem \ref{thm5.2} it is known that, for the above right-hand side, there exists a generalized solution in $\Omega_0$ of the form \begin{equation*} u_{n}(\varrho ,\varphi ,t)=u_{n}^{(1)}(\varrho ,t)\cos n\varphi \in C^2({% \bar{\Omega}}_0\backslash O), \end{equation*} which is classical solution in $\Omega_{\varepsilon }$, $\varepsilon \in (0,1)$. By setting $u_{n}^{(2)}(\varrho ,t)=\varrho ^{\frac{1}{2}% }u_{n}^{(1)}(\varrho ,t)$ and substituting \begin{equation} \xi =1-\varrho -t,\quad \eta =1-\varrho +t, \label{eq:5p6} \end{equation} the equation \eqref{eq:5p5a}, with boundary conditions \eqref{eq:5p2}, in view of \begin{equation} U(\xi ,\eta )=u_{n}^{(2)}(\varrho (\xi ,\eta ),t(\xi ,\eta )), \label{eq:5p7} \end{equation} becomes a Darboux-Goursat problem $P_{\alpha ,3}$: \begin{gather} U_{\xi \eta }-AU_{\xi }-BU_{\eta }-CU=F(\xi ,\eta ), \label{eq:5p8} \\ U(0,\eta )=0,\quad (U_{\eta }-U_{\xi })(\xi ,\xi )+\alpha (1-\xi )U(\xi ,\xi )=0. \label{eq:5p8a} \end{gather} Note that, because of the condition $a_2\equiv 0$ and the special right-hand side of \eqref{eq:5p5a}, we do not obtain a system as in the general case of Section 3, but a single equation \eqref{eq:5p8}. According to \eqref{eq:1p101}, the coefficients of \eqref{eq:5p8} are defined as follows: \begin{equation} \begin{gathered} A=\frac{1}{4}(a_1+b)\geq 0,\quad B=\frac{1}{4}(a_1-b)\geq 0, \\ C(\xi ,\eta )=\frac{1}{4}\Big( \frac{4n^2-1}{(2-\eta -\xi )^2}+\frac{a_1(\xi ,\eta )}{2-\eta -\xi }-c(\xi ,\eta )\Big) \geq 0,\ n\in \mathbb{N}, \end{gathered} \label{eq:5p89} \end{equation} \begin{equation} F(\xi ,\eta )=2^{n-\frac{5}{2}}\Big[ \frac{(1-\xi )(1-\eta )}{2-\eta -\xi }% \Big] ^{n-\frac{1}{2}}\in C^{n-1}({\bar{D}}_{\varepsilon }^{(1)}),\quad F(\xi ,\eta )\geq 0, \label{eq:5p10} \end{equation} where we preserve the same notations for $a_1,b$ and $c$ in the new coordinates $(\xi ,\eta )$. Next, in view of Theorem \ref{thm3.2} and Lemma \ref{lm3.1}, we formulate the following result. \begin{proposition} \label{prop6.1} There exists a classical solution $U(\xi,\eta )\in C^2(\bar{D}_0\setminus (1,1))$ for the problem \eqref{eq:5p8}, \eqref{eq:5p8a} for which \begin{equation*} U(\xi ,\eta )\geq 0,\quad U_{\xi }(\xi ,\eta )\geq 0,\quad U_{\eta } (\xi ,\eta )\geq 0 \quad \text{in } \bar{D}_{\varepsilon }^{(1)}. \end{equation*} \end{proposition} Set \begin{equation} K=\int_{D_{\frac{1}{2}}^{(1)}}F^2(\xi ,\eta )\,\,d\xi d\eta >0. \label{eq:5p10a} \end{equation} Then from \eqref{eq:5p8} for $0<\varepsilon <1/2$ it follows that \begin{equation} \begin{aligned} 00$ large enough. Following the proof of Theorem \ref{thm6.1} and using the transformations \eqref{eq:5p6} and \eqref{eq:5p7}, we lead to the function $W(\xi ,\eta )$ of \eqref{eq:5p7}, for which the equation % \eqref{eq:5p8} reduces to \begin{equation} W_{\xi \eta }-\frac{1}{2}(a_1+2\lambda ^{\prime })W_{\eta }-CW=F(\xi ,\eta ). \label{eq:5p103} \end{equation} Here $C(\xi ,\eta )$ and $F(\xi ,\eta )$ are functions from \eqref{eq:5p10} and \eqref{eq:5p89}. We formulate now a result to be used in the proof of Theorem \ref{thm7.1}. \begin{proposition} \label{prop6.2} There exists a classical solution $W(\xi,\eta )\in C^2(\bar{D}_0\backslash (1,1))$ of the problem \eqref{eq:5p103}, \eqref{eq:5p8a} for which \begin{equation} W(\xi ,\eta )\geq 0,\quad W_{\xi }(\xi ,\eta )\geq 0,\quad W_{\eta }(\xi ,\eta )\geq 0 \quad \mbox{in }\bar{D}_{\varepsilon}^{(1)}. \label{eq:5p104} \end{equation} \end{proposition} For the function \begin{equation} g_{n}(\varrho ,\varphi ,t)=\varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}\cos n\varphi , \label{eq:6p301} \end{equation} as the right-hand side of the equation $L_1w=g$, Theorem \ref{thm6.1} gives a singular solution $w_{n}$ satisfying \eqref{eq:5p3}; that is, \begin{equation} |w_{n}(\varrho ,\varphi ,\varrho )|\geq \frac{1}{2}|w_{n}(2\varrho ,\varphi ,0)|+\varrho ^{-n}|\cos n\varphi |\geq \varrho ^{-n}|\cos n\varphi |,\ 0<\varrho <1. \end{equation} Now, the inverse transform $u_{n}=\exp \left\{ \Lambda (\varrho +t)\right\} w_{n}$ gives \begin{equation} |u_{n}(\varrho ,\varphi ,\varrho )|\geq \frac{1}{2}|u_{n}(2\varrho ,\varphi ,0)|+\varrho ^{-n}|\cos n\varphi |\geq \varrho ^{-n}|\cos n\varphi |,\ 0<\varrho <1, \end{equation} where the function $u_{n}(\varrho ,\varphi ,t)$ is a solution of the problem % \eqref{eq:5p101}, \eqref{eq:5p2} with \begin{equation*} f(\varrho ,\varphi ,t)=\exp \left\{ \Lambda (\varrho +t)\right\} \text{ }% \varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}\cos n\varphi .\text{ } \end{equation*} The proof is complete. \hfill$\square$\smallskip Next, we find singular solutions for the original Problem $P_{\alpha}$, formulated in Section 1. \noindent\textbf{Proof of Theorem \ref{thm6.3}} Recall that we are looking for a suitable right-hand side function $f_{n}$ of \eqref{eq:0p1} for which singular solutions exist. Set \begin{equation} u(x_1,x_2,t)=\exp \left\{ (bt-b_1x_1-b_2x_2)/2\right\} v(x_1,x_2,t)\,. \label{eq:6p305} \end{equation} Then equation \eqref{eq:1p1} becomes \begin{equation} v_{x_1x_1}+v_{x_2x_2}-v_{tt}+c_1v=h\equiv \exp \left\{ (b_1x_1+b_2x_2-bt)/2\right\} f\text{\quad in }\Omega_0,\text{\ } \label{eq:6p302} \end{equation} where $c_1:=c+(b^2-b_1^2-b_2^2)/4$. In order to apply Theorem \ref{thm6.2}, we rewrite \eqref{eq:6p302} in polar coordinates and obtain the problem $P_{\gamma }$: \begin{equation*} \begin{gathered} \frac{1}{\varrho }(\varrho v_{\varrho })_{\varrho }+\frac{1}{\varrho ^2} v_{\varphi \varphi }-v_{tt}+c_1v=h(\varrho ,\varphi ,t)\quad\text{in } \Omega_0, \\ v\bigr\rvert_{\Sigma_1}=0,\quad \lbrack v_{t}+\gamma (\varrho )v]% \bigr\rvert_{\Sigma_0\backslash O}=0 \end{gathered} \end{equation*} with $\gamma (\varrho ):=\alpha (\varrho )+b/2$. In order to apply Theorem \ref{thm6.2} to the Problem $P_{\gamma }$, the only condition we need is $c_1\leq 0, $ i.e. $c+(b^2-b_1^2-b_2^2)/4\leq 0$, which is satisfied. If we now choose \begin{equation*} h_{n}(\varrho ,\varphi ,t)=\exp \left\{ \Lambda (\varrho +t)\right\} \varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}\cos n\varphi \end{equation*} according to Theorem \ref{thm6.2} with the constant $\Lambda $ large enough, then the Problem $P_{\gamma }$ has a corresponding singular solution $v_{n}\in C^2(% \bar{\Omega}_0\backslash O)$ with the estimate \begin{equation*} |v_{n}(\varrho ,\varphi ,\varrho )|\geq \frac{1}{2}|v_{n}(2\varrho ,\varphi ,0)|+\varrho ^{-n}|\cos n\varphi |,\quad 0<\varrho <1. \end{equation*} Now the inverse transform of \eqref{eq:6p305} gives a singular \textsl{% generalized solution} $u_{n}\in C^2(\bar{\Omega}_0\backslash O)$ of the Problem $P_{\alpha }$ with the right-hand side \begin{align*} f_{n}=&\exp \left\{ (bt-b_1x_1-b_2x_2)/2+\Lambda (\varrho +t)\right\}\\ &\times |x|^{-n}(x_1^2+x_2^2-t^2)^{n-1/2}\cos n(\arctan \frac{x_2}{x_1}). \end{align*} The proof is complete. \hfill$\square$ \begin{remark} \label{rmk6.1} Aldashev in \cite{Ald98} considered \eqref{eq:1p1} and studied the \textit{homogeneous} Problems $P_{\alpha }$ and $P_{\alpha }^{\ast }$. Unfortunately, as it is easy to check, the procedure which he follows leads to a correct conclusion only in the case of the wave equation, i.e. only in the case where all the lower order terms in \eqref{eq:1p1} are identically zero. Otherwise, this procedure leads to systems of differential equations which are not equivalent to those which should be solved (see \eqref{eq:1p9}). This is due to the fact that, in the systems obtained in \cite{Ald98} by integration with respect to $\varphi $, the Fourier coefficients $u_k$ of degree $k$ depend on the coefficients $u_{k-1}$ of degree $k-1$. \end{remark} \section{Applications to the wave equation, singular solutions} In this section we consider the wave equation \begin{equation} \square u=u_{x_1x_1}+u_{x_2x_2}-u_{tt}=f(x_1,x_2,t) \label{eq:6p1} \end{equation} subject to the boundary-value problem $P_{\alpha }$, i.e. \begin{equation} \square u=f\text{ in\ $\Omega_0$,}\quad u\bigr\rvert% _{\Sigma_1}=0,\quad \lbrack u_{t}+\alpha (|x|)u]\bigr\rvert_{\Sigma _0\backslash O}=0. \label{eq:6p2} \end{equation} As an application to the wave equation of the results of the previous section, we have the following statement. \begin{theorem} \label{thm7.1} Let $\alpha \in C^{\infty }(0,1]$ $\cap C[0,1]$ be an arbitrary function. Then: \begin{itemize} \item[(i)] For each $n\in \mathbb{N}$, $n\geq 4$, there exists a function\ $f_{n}\in C^{n-2}({\bar{\Omega}}_0)\cap C^{\infty }(\Omega _0)$, \ for which the corresponding generalized solution $u_{n}$ of the problem $P_{\alpha }$ belongs to $C^{n}({\bar{\Omega}}_0\backslash O)$ and satisfies the estimate \begin{equation} |u_{n}(x_1,x_2,|x|)|\geq \frac{1}{2}|u_{n}(2x_1,2x_2,0)|+|x|^{-n}|% \cos n(\arctan \frac{x_2}{x_1})|. \label{eq:6p3} \end{equation} \item[(ii)] In the case $\alpha (\varrho )\leq 0$ an upper estimate of the singular solution $u_{n}$ is \begin{equation} |u_{n}(x_1,x_2,t)|\leq C_{\mu }|x|^{-1/2}\Big( \frac{|x|}{% x_1^2+x_2^2-t^2}\Big) ^{n-\frac{1}{2}}|\cos n(\arctan \frac{x_2% }{x_1})|,\quad (|x|,t)\in D_1^{\mu }, \label{eq:6p4a} \end{equation} where $C_{\mu }$ is a constant. and \begin{equation*} D_1^{\mu }:=\left\{ (\varrho ,t):0<\varrho -t\leq \varrho +t\leq \mu (\varrho -t)\right\} ,\mu <2^{\frac{2n+1}{2n-1}}-1. \end{equation*} \end{itemize} Thus, for\ $\alpha (\varrho )\leq 0$ we have two-sided estimates, which in the limit cases $t=|x|$ and $t=0$ are: \begin{equation} \begin{gathered} |x|^{-n}|\cos n(\arctan \frac{x_2}{x_1})|\leq |u_{n}(x_1,x_2,|x|)|, \\ |u_{n}(x_1,x_2,0)|\leq C|x|^{-n}|\cos n(\arctan \frac{x_2}{x_1})|, \end{gathered} \label{eq:6p4} \end{equation} with $C$ a constant. That is, in the case of $\alpha (\varrho )\leq 0$ the exact behavior of $u_{n}(x_1,x_2,t)$ around $O$ is $(x_1^2+x_2^2+t^2)^{-n/2}\cos n(\arctan \frac{x_2}{x_1})$. \end{theorem} \noindent\textbf{Proof.} \textbf{(i)} Note that, the wave equation \eqref{eq:6p1} is of the form \eqref{eq:5p101} and so the first part of Theorem \ref{thm7.1} follows from Theorem \ref{thm6.2} and \cite{GHP}. Actually, according to this theorem we choose the function $f_{n}$ to be of the special form \begin{equation} \square u=f_{n}=\exp \left\{ \Lambda (\varrho +t)\right\} \text{ }\varrho ^{-n}(\varrho ^2-t^2)^{n-1/2}\cos n\varphi \quad\text{in } \Omega _0, \label{eq:6p85a} \end{equation} where $\Lambda >0$ is large enough and such that $\Lambda +\alpha (\varrho )\geq 0,\varrho \in \lbrack 0,1]$. Then by Theorems \ref{thm5.1} and \ref{thm5.2} there exists a unique generalized solution $u_{n}(\varrho ,\varphi ,t)$ of the equation \eqref{eq:6p85a}, satisfying the boundary conditions \eqref{eq:6p2} and the estimates \eqref{eq:6p3} (see Theorem \ref{thm6.2}). On the other hand, by \cite[Theorem 5.2]{GHP}, for the equation \eqref{eq:6p85a} there exists a generalized solution in $\Omega_0$ of the form \begin{equation*} u_{n}(\varrho ,\varphi ,t)=u_{n}^{(1)}(\varrho ,t)\cos n\varphi \in C^{n}({% \bar{\Omega}}_0\backslash O), \end{equation*} which is a classical solution of Problem $P_{\alpha }$ in $\Omega _{\varepsilon }$, $\varepsilon \in (0,1)$. \\ \textbf{(ii)} By setting $u_{n}^{(2)}(\varrho ,t)=\varrho ^{\frac{1}{2}% }u_{n}^{(1)}(\varrho ,t)$ and substituting \begin{equation} \xi =1-\varrho -t,\quad \eta =1-\varrho +t, \label{eq:6p86} \end{equation} the problem \eqref{eq:6p85a}, \eqref{eq:6p2}, in view of \begin{equation} U_{n}(\xi ,\eta )=u_{n}^{(2)}(\varrho (\xi ,\eta ),t(\xi ,\eta )), \label{eq:6p87} \end{equation} becomes a Darboux-Goursat problem $P_{\alpha ,3}:$% \begin{gather} U_{n,\xi \eta }-C(\xi ,\eta )U_{n}=G(\xi ,\eta )\equiv \exp \left\{ \Lambda (1-\xi )\right\} F(\xi ,\eta ), \label{eq:6p88} \\ U_{n}(0,\eta )=0,\quad (U_{n,\eta }-U_{n,\xi })(\xi ,\xi )+\alpha (1-\xi )U_{n}(\xi ,\xi )=0. \label{eq:7p201} \end{gather} Here, the coefficients \begin{equation} C(\xi ,\eta )=\frac{4n^2-1}{4(2-\eta -\xi )^2}\in C^{\infty }({\bar{D}} _{\varepsilon }^{(1)}),\quad n\geq 4, \label{eq:6p89} \end{equation} and \begin{equation} F(\xi ,\eta )=2^{n-\frac{5}{2}}\Big[ \frac{(1-\xi )(1-\eta )}{2-\eta -\xi } \Big] ^{n-\frac{1}{2}}\in C^{n-1}({\bar{D}}_{\varepsilon }^{(1)}) \label{eq:6p90} \end{equation} are defined by \eqref{eq:1p101} and \eqref{eq:1p19}. Now, we need some information about the behavior of the function $U_{n}(\xi ,\eta )$. Since, by Theorem \ref{thm6.2}, \begin{equation*} U_{n}(\xi ,\eta )=\exp \left\{ \Lambda (\varrho +t)\right\} \ W(\xi ,\eta )=\exp \left\{ \Lambda (1-\xi )\right\} \ W(\xi ,\eta ), \end{equation*} $W(\xi ,\eta )\geq 0$ and $W_{\eta }(\xi ,\eta )\geq 0$\ in $\bar{D}% _{\varepsilon }^{(1)}$, in view of Proposition \ref{prop6.2}, we formulate the following result. \begin{proposition} \label{prop7.1} There exists a classical solution $U_{n}(\xi ,\eta )\in C^{n}(\bar{D}_0\setminus (1,1))$ for \eqref{eq:6p88}, \eqref{eq:7p201} for which \begin{equation} U_{n}(\xi ,\eta )\geq 0,\quad U_{n,\eta }(\xi,\eta )\geq 0, \quad (\xi ,\eta )\in \bar{D}_{\varepsilon }^{(1)}. \label{eq:6p93} \end{equation} \end{proposition} Put \begin{equation} K_1=\int_{D_0}G^2(\xi ,\eta )\,d\xi \,d\eta >0. \label{eq:6p90a} \end{equation} Then, by \eqref{eq:6p88}, for $0<\varepsilon <1$ it follows that \begin{equation} \begin{aligned} K_1&\geq \int_{D_{\varepsilon }^{(1)}}G^2(\xi ,\eta )\,d\xi \,d\eta \geq \int_{D_{\varepsilon }^{(1)}}G(\xi ,\eta )\,F(\xi ,\eta )\,d\xi\,d\eta \\ &=\int_{D_{\varepsilon }^{(1)}}U_{n,\xi \eta }F(\xi ,\eta )\,d\xi \,d\eta -\int_{D_{\varepsilon }^{(1)}}C(\xi ,\eta )U_{n}(\xi ,\eta )F(\xi ,\eta )\,d\xi \,d\eta \\ &=:I_1-I_2. \end{aligned} \label{eq:6p91} \end{equation} Then \begin{equation} \begin{aligned} I_1=&\int_0^{1-\varepsilon }\int_{\xi }^{1}(U_{n,\xi \eta }F)(\xi ,\eta )\,d\eta \,d\xi \\ =&-\int_0^{1-\varepsilon }U_{n,\xi }(\xi ,\xi )F(\xi ,\xi )\,d\xi -\int_{D_{\varepsilon }^{(1)}}(U_{n,\xi }F_{\eta })(\xi ,\eta )\,d\xi \,d\eta , \end{aligned}\label{eq:6p12} \end{equation} and, by \eqref{eq:6p90}, $F(\xi ,1)=0$. Since \begin{equation} \begin{aligned} &\int_{D_{\varepsilon }^{(1)}}(U_{n,\xi }F_{\eta })(\xi ,\eta)\,d\xi \,d\eta \\ &=\int_0^{1-\varepsilon }(U_{n}F_{\eta })(\eta ,\eta )\,d\eta +\int_{1-\varepsilon }^{1}(U_{n}F_{\eta })(1-\varepsilon ,\eta )\,d\eta -\int_{{D}_{\varepsilon }^{(1)}}(U_{n}F_{\xi \eta })(\xi ,\eta )\,d\xi \,d\eta , \end{aligned} \end{equation} equation \eqref{eq:6p12} becomes \begin{equation} \begin{split} I_1& =-\int_0^{1-\varepsilon }[U_{n,\xi }(\xi ,\xi )F(\xi ,\xi )+U_{n}(\xi ,\xi )F_{\eta }(\xi ,\xi )]\,d\xi \\ & \quad -\int_{1-\varepsilon }^{1}(U_{n}F_{\eta })(1-\varepsilon ,\eta )\,d\eta +\int_{D_{\varepsilon }^{(1)}}(U_{n}F_{\xi \eta })(\xi ,\eta )\,d\xi \,d\eta . \end{split} \label{eq:6p13} \end{equation} From \eqref{eq:6p13} and \eqref{eq:6p91} it follows that \begin{equation} \begin{split} K_1\geq& I_1-I_2=-\int_0^{1-\varepsilon }[U_{n,\xi }(\xi ,\xi )F(\xi ,\xi )+U_{n}(\xi ,\xi )F_{\xi }(\xi ,\xi )]\,d\xi \\ &-\int_{1-\varepsilon }^{1}(U_{n}F_{\eta })(1-\varepsilon ,\eta )\,d\eta +\int_{{D}_{\varepsilon }^{(1)}}U_{n}[F_{\xi \eta }-CF](\xi ,\eta )\,d\xi \,d\eta . \end{split}\label{eq:6p14} \end{equation} Because of \eqref{eq:5p92}, the last integral vanishes. Thus, using the boundary conditions for the functions $U_{n}$ and $F$, when $\eta =\xi $, we see that \begin{equation} \begin{aligned} K_1&\geq I_1-I_2\\ &=-\int_0^{1-\varepsilon }[U_{n,\xi }(\xi ,\xi)F(\xi,\xi ) +U_{n}(\xi ,\xi )F_{\xi }(\xi ,\xi )]\,d\xi -\int_{1-\varepsilon }^{1}(U_{n}F_{\eta })(1-\varepsilon ,\eta )\,d\eta \\ &=-\frac{1}{2}(FU_{n})(1-\varepsilon ,1-\varepsilon ) -\frac{1}{2}\int_0^{1-\varepsilon }\alpha (1-\xi )U_{n}(\xi ,\xi )F(\xi ,\xi )\,d\xi \\ &\quad -\int_{1-\varepsilon }^{1}(U_{n}F_{\eta })(1-\varepsilon ,\eta )\,d\eta . \end{aligned} \label{eq:6p22} \end{equation} Since $\alpha (\xi )$, \quad $F_{\eta }\leq 0$ and $F,U_{n},U_{n,\eta }\geq 0$, for $0<\delta <\varepsilon <1$, we have \begin{equation} \begin{split} K_1& \geq I_1-I_2 \\ & \geq -\frac{1}{2}(U_{n}F)(1-\varepsilon ,1-\varepsilon )+\int_{1-\varepsilon }^{1}U_{n}(1-\varepsilon ,\eta )\lvert F_{\eta }(1-\varepsilon ,\eta )\rvert \,d\eta \\ & \geq -\frac{1}{2}(U_{n}F)(1-\varepsilon ,1-\varepsilon )+\int_{1-\delta }^{1}U_{n}(1-\varepsilon ,\eta )\lvert F_{\eta }(1-\varepsilon ,\eta )\rvert \,d\eta \\ & \geq -\frac{1}{2}(U_{n}F)(1-\varepsilon ,1-\varepsilon )+\int_{1-\delta }^{1}U_{n}(1-\varepsilon ,1-\delta )\lvert F_{\eta }(1-\varepsilon ,\eta )\rvert \,d\eta \\ & \geq -\frac{1}{2}(U_{n}F)(1-\varepsilon ,1-\varepsilon )+(U_{n}F)(1-\varepsilon ,1-\delta ) \\ & \geq U_{n}(1-\varepsilon ,1-\delta )\left[ F(1-\varepsilon ,1-\delta )-% \frac{1}{2}F(1-\varepsilon ,1-\varepsilon )\right] \\ & \geq \nu (U_{n}F)(1-\varepsilon ,1-\delta ), \end{split} \label{eq:6p27} \end{equation} provided taht the constant $v>0$ satisfies \begin{equation} 2(1-\nu )F(1-\varepsilon ,1-\delta )\geq F(1-\varepsilon ,1-\varepsilon ). \label{eq:6p50} \end{equation} Using the explicit formula \eqref{eq:6p90} for the function $F(\xi ,\eta )$, we see that the above inequality is equivalent to \begin{equation} 2(1-\nu )\big( \frac{\delta }{\varepsilon +\delta }\big) ^{n-\frac{1}{2}% }\geq 2^{-n+\frac{1}{2}}, \label{eq:6p51} \end{equation} which implies \begin{equation} 0<\nu \leq 1-\frac{1}{2}\big( \frac{\varepsilon +\delta }{2\delta }\big) ^{n-\frac{1}{2}}. \label{eq:6p52} \end{equation} A necessary condition, for \eqref{eq:6p52} to be satisfied is that \begin{equation} 1\leq \frac{\varepsilon }{\delta }<2^{\frac{2n+1}{2n-1}}-1. \label{eq:6p53} \end{equation} In this concrete case, using \eqref{eq:6p53}, we can find an upper estimate for the generalized solution $u_{n}$. To do this, we consider the domain \begin{equation} D^{\mu }:=\{(\xi ,\eta ):1-\eta \leq 1-\xi \leq \mu (1-\eta )\}, \label{eq:6p54} \end{equation} where $1\leq \mu <2^{\frac{2n+1}{2n-1}}-1$. Observe that \begin{equation*} \inf_{D^{\mu }}\Big\{ 1-\frac{1}{2}\Big( \frac{1-\xi +1-\eta }{2(1-\eta )}% \Big) ^{n-\frac{1}{2}}\Big\} =1-\frac{1}{2}\Big( \frac{1+\mu }{2}% \Big) ^{n-\frac{1}{2}}=:C_{\mu }>0. \end{equation*} For $\nu =C_{\mu }$ the inequalities \eqref{eq:6p51} and \eqref{eq:6p50} are satisfied and so, by \eqref{eq:6p27}, we see that \begin{equation} U(\xi ,\eta )\leq 2^{-n+5/2}K_1C_{\mu }^{-1}\Big( \frac{2-\xi -\eta }{% (1-\xi )(1-\eta )}\Big) ^{n-\frac{1}{2}},\quad (\xi ,\eta )\in D^{\mu }. \label{eq:6p54a} \end{equation} By \eqref{eq:5p7} and \eqref{eq:5p6}, the inequality \eqref{eq:6p54a} transforms to \begin{equation} u_{n}^{(2)}(\varrho ,t)\leq 4K_1C_{\mu }^{-1}\Big( \frac{\varrho }{ \varrho ^2-t^2}\Big) ^{n-\frac{1}{2}}, \label{eq:6p55} \end{equation} which is satisfied for $(\varrho ,t)\in D_1^{\mu }:=\big\{ 0<\varrho -t\leq \varrho +t\leq \mu (\varrho -t)\big\}$. Finally, \eqref{eq:6p55} implies \begin{equation} u_{n}^{(1)}(\varrho ,t)\leq 4K_1C_{\mu }^{-1}\varrho ^{-1/2} \Big( \frac{\varrho }{\varrho ^2-t^2}\Big) ^{n-\frac{1}{2}} \quad\text{for }(\varrho ,t)\in D_1^{\mu }, \label{eq:6p56} \end{equation} which coincides with the estimate \eqref{eq:6p4a}. Note that $C_{\mu }=1/2$ on $\{ t=0\} $ and so \begin{equation} u_{n}^{(1)}(\varrho ,0)\leq 8K_1\varrho ^{-n},\quad 0<\varrho <1, \label{eq:6p57} \end{equation} which is the upper estimate in \eqref{eq:6p4}. The proof of Theorem \ref{thm7.1} is complete.\hfill $\square$ \begin{remark} \label{rmk7.1} \rm Since in Theorems \ref{thm6.1}, \ref{thm6.2} and \ref{thm7.1} the conditions imposed upon lower terms of \eqref{eq:5p1} are not invariant with respect to substitution of the independent variables \begin{equation} v(\varrho ,\varphi ,t)=u(\varrho ,\varphi ,t)\ \exp \lambda (\varrho ,t), \label{eq:6p1000} \end{equation} for various functions $\lambda (\varrho ,t)$, we can find a series of singular solutions of Problem $P_{\alpha }$ for different classes of equations of the form \eqref{eq:5p1}. This procedure is interesting by itself and is demonstrated by the following simple example \end{remark} \noindent\textbf{Example.} Consider the special form of the equation \eqref{eq:5p1}, with constant lower order terms, that is \begin{equation} Lu\equiv u_{x_1x_1}+u_{x_2x_2}-u_{tt}+b_1u_{x_1}+b_2u_{x_2}+bu_{t}+% \frac{1}{4}(b_1^2+b_2^2-b^2)u=f,\text{ in }\Omega_0, \label{eq:6p1a} \end{equation} with the boundary conditions \eqref{eq:5p2}. Obviously, the equation \eqref{eq:6p1a} satisfies the conditions of Theorems \ref{thm6.3} for $\alpha \in C^{1}([0,1])$ and we obtain a singular solution $u_{n}$ By using the transform \eqref{eq:6p305}, the equation \eqref{eq:6p1a} becomes the wave equation \eqref{eq:6p1}. Then Theorem \ref{thm7.1} becomes useful and in the case, when $\alpha (|x|)\leq -b/2$, in addition, we have two-sided estimates \eqref{eq:6p4} of the generalized solution $u_{n}(x_1,x_2,t)$, whose exact behavior around the point $O\ $ is $(x_1^2+x_2^2+t^2)^{-n/2}\cos n(\arctan \frac{x_2}{x_1})$. \smallskip \noindent \textbf{Acknowledgments.} The authors would like to thank the anonymous referee for making several helpful suggestions for improving the presentation of this paper. The research of Hristov was partially supported by the Bulgarian NSC under Grant MM-904/99. The essential part of the present work was finished and prepublished in \cite{GHPTR}, while Popivanov was visiting the University of Ioannina during 2001. Popivanov would like to thank the Ministry of National Economy of Helenic Republic for providing the NATO Science Fellowship (Ref. No.169/DOP/01) and the University of Ioannina for its hospitality \begin{thebibliography}{99} \bibitem{Ald93} S.A. Aldashev, Correctness of multidimensional Darboux problems for the wave equation, Ukrainian Math. J., \textbf{45 }(1993), 1456-1464. \bibitem{Ald98} S.A. Aldashev, On Darboux problems for a class of multidimensional hyperbolic equations, Differ. Equations, \textbf{34} (1998), 65--69. \bibitem{Ald00} S.A. Aldashev, Some problems for a multidimensional hyperbolic integro-differential equation. Ukrainian Math. J., \textbf{52} (2000), no. 5, 590--595. \bibitem{AS} A.K. Aziz, M. Schneider, Frankl-Morawetz problems in $R^{3}$, SIAM J. Math. Anal., \textbf{10} (1979), 913-921. \bibitem{BB} Ar.B. Bazarbekov, Ak.B. Bazarbekov, Goursat and Darboux problems for the two-dimensional wave equation, I, Differ. Equations, \textbf{30 }(1994), 741-748. \bibitem{Bits} A.V. Bitsadze, Some classes of partial differential equations, Gordon and Breach Science Publishers, New York, 1988. \bibitem{EP98} D.E. Edmunds, N.I. Popivanov, A nonlocal regularization of some over-determined boundary value problems I, SIAM J. Math. Anal., \textbf{29} (1998), No1, 85-105. \bibitem{Gar} P.R. Garabedian, Partial differential equations with more than two variables in the complex domain, J. Math. Mech., \textbf{9} (1960), 241-271. \bibitem{GHP} M.K. Grammatikopoulos, T.D. Hristov and N.I. Popivanov, Singularities of the 3-D Protter's problem for the wave equation, Electron. J. Diff. Eqns., \textbf{2001}(2001), No. 01, pp. 1-26 (URL: http://ejde.math.swt.edu). \bibitem{GHPTR} M.K. Grammatikopoulos, T.D. Hristov and N.I. Popivanov, Singularities of 3-D Protter's problem for the wave equation, involving lower order terms, Technical Reports of the University of Ioannina, Greece, \textbf{5 }(June, 2001), No. 02, pp. 1-31. \bibitem{H} L. H\"{o}rmander, The Analysis of Linear Partial Differential Operators III. Springer-Verlag, Berlin-Heidelberg-New York-Tokyo (1985). \bibitem{Kor} Jong Duek Jeon, Khe Kan Cher, Ji Hyun Park, Yong Hee Jeon, Jong Bae Choi, Protter's conjugate boundary value problems for the two dimensional wave equation, J. Korean. Math. Soc. \textbf{33 }(1996), 857-863. \bibitem{Kar82} G.D. Karatoprakliev, Uniqueness of solutions of certain boundary-value problems for equations of mixed type and hyperbolic equations in space, Differ. Equations, \textbf{18 }(1982), 49-53. \bibitem{Kh95} S. Kharibegashvili, On the solvability of a spatial problem of Darboux type for the wave equation, Georgian Math. J., \textbf{2} (1995), 385-394. \bibitem{Kan95} Khe Kan Cher, Darboux-Protter problems for the multidimensional wave equation in the class of unbounded functions, Math. Notices of Jacutsk State Univ, \textbf{2 }(1995), 105-109. \bibitem{Kan98} Khe Kan Cher, On nontrivial solutions of some homogeneous boundary value problems for the multidimensional hyperbolic Euler-Poisson-Darboux equation in an unbounded domain, Differ. Equations, \textbf{34} (1998), 139--142. \bibitem{Kan981} Khe Kan Cher, On the conjugate Darboux-Protter problem for the two-dimensional wave equation in the special case. Nonclassical equations in mathematical physics (Russian) (Novosibirsk, 1998), 17--25, Izdat. Ross. Akad. Nauk Sib. Otd. Inst. Mat., Novosibirsk, 1998. \bibitem{Nak70} A.M. Nakhushev, A multidimensional analogue of the problem of Darboux for hyperbolic equations, Sov. Math. Dokl., \textbf{11 }(1970), 1162--1165. \bibitem{Nak92} A.M. Nakhushev, Criteria for continuity of the gradient of the solution to the Darboux problem for the Gellerstedt equation, Differ. Equations, \textbf{28} (1992), 1445-1457. \bibitem{PS88} N.I. Popivanov and M. Schneider, The Darboux problem in $% R^{3}$ for a class of degenerated hyperbolic equations, Comptes Rend. de l'Acad.~Bulg.~Sci., \textbf{41, }11 (1988), 7--9. \bibitem{PS95} N.I. Popivanov, M. Schneider, On M.H.\ Protter problems for the wave equation in $R^{3}$, J. Math. Anal. Appl., \textbf{194} (1995), 50-77. \bibitem{NT} N.I. Popivanov, T.P. Popov, Exact behavior of the singularities for the 3-D Protter's problem for the wave equation, In: ''Inclusion Methods for Nonlinear Problems'' with Applications in Engineering, Economics and Physics, Ed. Herzberger, J, \ ''Computing'', Supplement 16 (to appear). \bibitem{Prot} M.H. Protter, New boundary value problem for the wave equation and equations of mixed type, J. Rat. Mech. Anal, \textbf{3} (1954), 435-446. \bibitem{Tong} Tong Kwang-Chang, On a boundary value problem for the wave equation, Science Record, New Series, \textbf{1} (1957), 1-3. \end{thebibliography} \end{document}