\begin{document} \title{On the Sturm-Liouville Operator with Summable Potential} \author{O. A. Veliev \\ %EndAName {\small Depart. of Math, Fen-Ed. Fak, Marmara University., }\\ {\small \ G\"{o}ztepe Kamp\"{u}s\"{u}, 81040 , Kadik\"{o}y Istanbul , Turkey,% }\\ {\small \ e-mail: oveliev@marmara.edu.tr} \and M.Toppamuk Duman \\ %EndAName {\small \ Depart. of Math, Fen-Ed. Fak, Dokuz Eyl\"{u}l Univ., }\\ {\small \ Tinaztepe Kam, Kaynaklar, Buca, 35160, \ Izmir, Turkey,}\\ {\small \ e-mail: melda.duman@deu.edu.tr }} \maketitle \begin{abstract} We investigate the Sturm-Liouville operator \[ L(q)=-\frac{d^{2}}{dx^{2}}+q(x) \] in $L_{2}[0,1]$ with strongly regular boundary conditions and arbitrary Lebesque integrable Potential $q(x)$. We obtain asymptotic formulas of arbitrary order for eigenvalues and eigenfunctions of $L(q).$ Besides we give a simple proof of Riesz basisness of eigenfunctions and associeted functions of this operator. \end{abstract} \section{ Introduction} Consider the boundary value problem \begin{equation} -y^{^{\prime \prime }}(x)+q(x)y(x)=\lambda y(x),\qquad 0\leq x\leq 1 \end{equation} \[ U_{1}(y)=a_{1}y^{^{\prime }}(0)+b_{1}y^{^{\prime }}(1)+a_{0}y(0)+b_{0}y(1)=0, \] \begin{equation} U_{2}(y)=c_{1}y^{^{\prime }}(0)+d_{1}y^{^{\prime }}(1)+c_{0}y(0)+d_{0}y(1)=0, \end{equation} where $q(x)$ is a Lebesgue integrable complex-valued function and the coefficients of the boundary conditions (2)$\ $are complex numbers. We use the following notations, definitions, and results of [3] ( see p. 56-65 of [3]). The normalized boundary conditons of (2) is \[ U_{i}(y)=U_{i0}(y)+U_{i1}(y)=0,\text{ }i=1,2, \] where $U_{i0}(y)=\alpha _{i}y^{k_{i}}(0)+\alpha _{i,0}y(0)$ , $% U_{i1}(y)=\beta _{i}y^{k_{i}}(1)+\beta _{i,0}y(1)$ and $k_{i}$ are the maximal orders such that $0\leq k_{2}\leq k_{1}\leq 1$ and at least one of $% \alpha _{i}$ and $\beta _{i}$ is nonzero for any $i=1,2$. The normalized boundary conditions is said to be regular if the numbers $\theta _{-1}$ and $% \theta _{1}$ defined by \begin{equation} \frac{\theta _{-1}}{s}+\theta _{0}+s\theta _{1}=\left| \begin{array}{cc} (\alpha _{1}+s\beta _{1})w_{1}^{k_{1}} & (\alpha _{1}+\frac{1}{s}\beta _{1})w_{2}^{k_{1}} \\ (\alpha _{2}+s\beta _{2})w_{1}^{k_{2}} & (\alpha _{2}+\frac{1}{s}\beta _{2})w_{2}^{k_{2}} \end{array} \right| \label{charct} \end{equation} are nonzero, where $w_{1}$ and $w_{2}$ are the two distinct roots of -1. If the roots $\zeta _{1},\zeta _{2}$ of (3) (or $\theta _{1}\zeta ^{2}+\theta _{0}\zeta +\theta _{-1}=0)$ are simple, that is, \begin{equation} \theta _{0}^{2}-4\theta _{-1}\theta _{1}\neq 0, \label{discrminant} \end{equation} then we say that the boundary conditions (2) \ are strongly regular. This occurs in only following three cases: \textbf{1. }$a_{1}d_{1}-b_{1}c_{1}\neq 0.$ Then (2) equivalent to the conditions\textbf{\ } \begin{equation} y^{^{\prime }}(0)+\alpha _{11}y(0)+\alpha _{12}y(1)=0,\text{ }y^{^{\prime }}(1)+\alpha _{21}y(0)+\alpha _{22}y(1)=0. \label{bc1} \end{equation} Hense $\theta _{0}=0,$ $\theta _{-1}=\theta _{1}=1,$ $\theta _{0}^{2}-4\theta _{-1}\theta _{1}\neq 0$ \textbf{2. }$a_{1}d_{1}-b_{1}c_{1}=0,$ $\left| a_{1}\right| +\left| b_{1}\right| >0,$ $b_{1}c_{0}+a_{1}d_{0}\neq 0$ , $c_{0}\neq \pm d_{0},$ $% a_{1}\neq \pm b_{1}$. Then (2) equivalent to \begin{eqnarray} U_{1}(y) &=&a_{1}y^{^{\prime }}(0)+b_{1}y^{^{\prime }}(1)+a_{0}y(0)+b_{0}y(1)=0,\text{ } \nonumber \\ \text{\ }U_{2}(y) &=&c_{0}y(0)+d_{0}y(1)=0, \end{eqnarray} and $\theta _{0}=2(a_{1}c_{0}+b_{1}d_{0})w_{1},$ \ $\theta _{-1}=\theta _{1}=(b_{1}c_{0}+a_{1}d_{0})w_{1}\neq 0.$ Hense \begin{equation} \theta _{0}^{2}-4\theta _{-1}\theta _{1}=4(c_{0}^{2}-d_{0}^{2})(a_{1}^{2}-b_{1}^{2})\neq 0 \end{equation} if and only if $c_{0}\neq \pm d_{0},$ $a_{1}\neq \pm b_{1}.$ \textbf{3. } $a_{1}=b_{1}=c_{1}=d_{1}=0,a_{0}d_{0}+b_{0}c_{0}\neq 0.$ Then the \ conditions (2) are equivalent to the Dirichlet conditions and $% \theta _{0}=0,\theta _{-1}=-\theta _{1}=1.$ If the boundary conditions (2) \ are strongly regular then the eigenvalues of the operator $L(q(x))$ , generated by (1), \ (2) consists of the sequences $\{\lambda _{n,1}\},\{\lambda _{n,2}\}$ \ satisfying \begin{equation} \lambda _{n,1}=(2n\pi -i\ln \zeta _{1})^{2}+O(1), \label{sq1} \end{equation} \begin{equation} \lambda _{n,2}=(2n\pi -i\ln \zeta _{2})^{2}+O(1), \label{sq2} \end{equation} for $n\geq N,$ $N>>1,$ where $\zeta _{1}$ and $\zeta _{2}$ are the distinct roots of (3) (see (4) ) . From these formulas, and the same formulas for the eigenvalues $\{\lambda _{n,1}^{0}\},\{\lambda _{n,2}^{0}\}$ \ of $L(0)$ one can easily obtain the following inequalities from which we use essentially: \begin{eqnarray} \left| \lambda _{n,j}-\lambda _{k,j}^{0}\right| &>&\left| 2(n-k)\pi \right| \left| 2(n+k)\pi -i2\ln \zeta _{j}\right| -c_{1}>c_{2}n, \nonumber \\ \text{ }j &=&1,2,\text{\ }\forall k\neq n,\text{ }k=0,1,..., \end{eqnarray} \begin{equation} \left| \lambda _{n,j}-\lambda _{k,s}^{0}\right| \geq \left| J_{1}\right| \left| J_{2}\right| -c_{3}>c_{4}n,\text{ \ }j\neq s,\forall k=0,1,.... \label{dis2} \end{equation} for $n\geq N,$ where $J_{1}=2(n-k)\pi -i(\ln \zeta _{1}-\ln \zeta _{2}),$ $J_{2}=2(n+k)\pi -i(\ln \zeta _{1}+\ln \zeta _{2}).$ \ Here and in forthcaming relations by $N$ we denote a big positive integer, that is, $N\gg 1,$ and by $c_{m},$ for $m=1,2,...,$ the positive, independent on $N,$ constants whose exact value are inessential. Proof of (10) follows from the relations $n-k\neq 0$, $n+k>n>>1$, $\left| \ln \zeta _{j}\right| c_{6}n,$ because $n+k>n>>1$ and $\left| \ln \zeta _{1}+\ln \zeta _{2}\right| c_{8}. \] \ 2. $r_{1}=r_{2}$ . Then $t_{1}-t_{2}\neq 0,$ $\left| t_{1}-t_{2}\right| <2\pi .$ Therefore $\left| J_{1}\right| \geq \left| t_{1}-t_{2}\right| \geq c_{9}$ whether $% n-k=0$ or not. Hence in all cases we have $\left| J_{1}\right| >c_{10}.$ So $% \left| J_{1}J_{2}\right| >c_{10}c_{6}n.$ Moreover from \ (10), (11) it follows that if $k>2n,n\geq N,$ $N>>1$ then \begin{equation} \left| \lambda _{n,j}-\lambda _{k,i}^{0}\right| \geq k^{2},\text{ \ }\forall i,j=1,2, \end{equation} since \ $\left| 2(n-k)\pi \right| >2k,\left| 2(n+k)\pi -i2\ln \zeta _{1}\right| >2k,\left| J_{1}\right| >2k,\left| J_{2}\right| >2k.$ The eigenfunctions $F_{n,i}(x)$ belonging to $\lambda _{n},_{j}$ satisfy the formula \begin{eqnarray} F_{n,j}(x) &=&(-i)^{k_{2}}(\alpha _{2}+\frac{1}{\zeta _{j}}\beta _{2}+O(% \frac{1}{n}))e^{i\rho _{n,j}x}- \\ &&(-i)^{k_{2}}(\alpha _{2}+\zeta _{j}\beta _{2}+O(\frac{1}{n}))e^{-i\rho _{n,j}x}. \nonumber \end{eqnarray} (see [3, p.77]) , where $(\rho _{n,j})^{2}=\lambda _{n,j}$ , $\rho _{n,j}=2n\pi -i\ln \zeta _{1}+O(\frac{1}{n}),$ and in Case 1 ( see (5)) $k_{2}=1,\alpha _{2}=\beta _{2}=1,\xi _{j}=\pm i.$ in Case 2 ( see (6)) $k_{2}=0,\alpha _{2}=c_{0},\beta _{2}=d_{0},\xi _{j}\neq 0,c_{0}\neq \pm d_{0},$ in Case 3 $k_{2}=0,\alpha _{2}=\beta _{2}=1,\xi _{j}=\pm 1$ Using these relations one can easily verify that the normalized eigenfunction $\Psi _{n,j}(x)$ of $L(q)$ corresponding to the eigenvalue $% \lambda _{n,i}$ satisfy \begin{equation} \Psi _{n,j}(x)=A_{j}e^{i(2\pi n+\gamma _{j})x}+B_{j}e^{-i(2\pi n+\gamma _{j})x}+O(\frac{1}{n}),j=1,2, \end{equation} where $A_{j},B_{j}$ are constants and $\gamma _{j}=-i\ln \zeta _{j}.$ Using these \i n Section 2 we give a simple proof of Riesz basisness of eigenfunctions and associeted functions of the operator $L(q)$ with strongly regular boundary conditions and arbitrary Lebesque integrable Potential $% q(x) $. For the first time the basisness of eigenfunctions and associeted functions of \ ordinary differential operators with strongly regular boundary conditions is investigated in [5,6,7] ( see also [3,4]). In section 3 we obtain the main results of this paper, naimly we obtain asymptotic formulas of arbitrary order for eigenvalues and eigenfunctions of $L(q)$ when $q(x)\in L_{1}[0,1],$ that is, when there is not any condition about smoothness of the\ potential $q(x)$. \section{On the Riesz Basisness} First we consider the adjoint boundary conditions. Using the formula \begin{equation} y^{^{\prime }}(1)\overline{z(1)}-y^{^{\prime }}(0)\overline{z(0)}-y(1)% \overline{z^{^{\prime }}(1)}+y(0)\overline{z^{^{\prime }}(0)}=0 \end{equation} one can compute the adjoint boundary conditions of (2). In the case 1, (see (5)), the adjoint boundary conditions to (2) is \begin{equation} y^{^{\prime }}(0)+\overline{\alpha }_{11}y(0)-\overline{\alpha }_{21}y(1)=0,% \text{ }y^{^{\prime }}(1)-\overline{\alpha }_{12}y(0)+\overline{\alpha }% _{22}y(1)=0. \label{normbc1} \end{equation} Clearly, for this conditions, we have $\theta _{0}=0,$ $\theta _{-1}=\theta _{1}=1,$ $\theta _{0}^{2}-4\theta _{-1}\theta _{1}\neq 0,$ that is, this is strongly regular boundary conditions. The case 3, i.e., the Dirichlet boundary conditions are self adjoint. To find the adjoint boundary conditions of the case 2 ( see (6)) we supplement the conditions \[ U_{3}(y)=a_{1}y(0)-b_{1}y(1)=0,U_{4}(y)=\text{ }y^{^{\prime }}(0)+y^{^{\prime }}(1)=0. \] The determinant of the obtained sistem is $% (a_{1}-b_{1})(b_{1}c_{0}+a_{1}d_{0})\neq 0.$ Finding $y(0),y(1),$ $% y^{^{\prime }}(0),y^{^{\prime }}(1)$ in term $U_{1},U_{2},U_{3},U_{4}$ and then putting they in (15) we see that the adjoint conditions of (6) is \[ \overline{d}_{0}y^{^{\prime }}(0)+\overline{c}_{0}y^{^{\prime }}(1)+\frac{% \overline{a_{0}d_{0}-b_{0}c_{0}}}{\overline{b_{1}-a_{1}}}(y(1)-y(0))=0,\text { \ }\overline{a}_{1}y(1)+\overline{b}_{1}y(0)=0, \] which is strongly regular boundary condition ( see (7)). Hence in all cases the adjoint boundary conditions of the strongly regular boundary conditions are also strongly regular boundary condition. Denote by $\Psi _{n,i,0}(x)$ $% \equiv \Psi _{n,i}(x)$ the eigenfunction of $L(q)$ corresponding to the eigenvalue $\lambda _{n,i}$ and by $\Psi _{n,i,p}(x),$ for $% p=1,2,...,t(n,i), $ the associated functions of $L(q)$ corresponding to the eigenfunction $\Psi _{n,i}(x),$ that is, \begin{equation} (L(q)-\lambda _{n,i})\Psi _{n,i}=0, \end{equation} \begin{equation} (L(q)-\lambda _{n,i})\Psi _{n,i,p}=\Psi _{n,i,p-1},\text{ \ }% p=1,2,...,t(n,i). \end{equation} The eigenfunctions and the associeted functions of $L(0)$\ are denoted by $\Psi _{n,i}^{0}(x)\equiv \Psi _{n,i,0}^{0}(x)$ and $\Psi _{n,i,p}^{0}(x)$ respectively. \begin{theorem} If the conditions (2) are strongly regular then the system of eigenfunctions and associated functions of \ $L(q)$ form a Riesz basis of $L_{2}[0,1].$ \end{theorem} \begin{proof} Since the strongly regular boundary conditions are nondegenerated boundary conditions the system of eigenfunctions and associated functions of \ $L(q)$ is complete ( see [2, Chap. 1, Sec.3] ). The system of eigenfunctions and associated functions of \ $L^{\ast }(q)$ is also complete in $L_{2}[0,1]$ becouse, as we noted above, the \ adjoint boundary conditions are strongly regular. Now we prove that the biortogonal system of the system $\{\Psi _{n,i,p}(x)\}$ consist of eigenfunctions and associated functions of \ the operator $% L^{\ast }(q)$ genereted by the adjoint expression $-y^{^{\prime \prime }}(x)+% \overline{q(x)}y(x)$ \ and the adjoint boundary conditions. Let $\Phi _{m,i}(x)$ $\equiv \Phi _{m,i,0}(x)$ be the eigenfunction of $L^{\ast }(q)$ corresponding to the eigenvalue $\mu _{m,i}.$ Denote by $\Phi _{m,i,p}(x),$ for $p=1,2,...,$ the associated functions corresponding to the eigenfunction $\Phi _{m,i}(x),$ that is, \begin{equation} (L^{\ast }(q)-\mu _{m,i})\Phi _{m,i}=0, \end{equation} \begin{equation} (L^{\ast }(q)-\mu _{m,i})\Phi _{m,i,p}=\Phi _{m,i,p-1},\text{ \ }p=1,2,...,. \end{equation} Multiplying both side of \ (19), (20) by \ $\Psi _{n,i,p}(x),$ for $% p=0,1,2,...,t(n,i)$ \ and using (17), (18) one can easily verify that if $% \overline{\mu _{m,i}}\neq \lambda _{n,i}$ then \ \begin{equation} (\Psi _{n,i,p}(x),\Phi _{m,i,j})=0,\text{ }\forall p,j=0,1,2,..., \end{equation} Since the system of eigenfunctions and associated functions of $L^{\ast }(q)$% , is complete in $L_{2}[0,1]$ \ the function $\Psi _{n,i,p}(x)$ can not be orthogonal to all elements of this system. Therefore no there exists eigenvalue of $L^{\ast }(q)$ different from all eigenvalues of $L(q)$ and similarly no there exists eigenvalue of $L(q)$ different from all eigenvalues of $L^{\ast }(q),$ that is, the eigenvalues of $L^{\ast }(q)$ consists of $\overline{\lambda _{n,i}}.$\ \ Denote by $E$ and $E^{\ast }$ the linear spaces genereted by the eigenfunction and the associated function of $L(q)$ and $L^{\ast }(q)$ corresponding to $\lambda _{n,i}$ and $% \overline{\lambda _{n,i}}$ respectively. Consider the operators \begin{equation} P=\int_{\gamma }(L(q)-\lambda I)^{-1}d\lambda ,\text{ }T=\int_{\gamma ^{\ast }}(L^{\ast }(q)-\lambda I)^{-1}d\lambda , \end{equation} where $\gamma $ is the circle belonging to the resolvent set of $L(q)$ and containing only the eigenvalue $\lambda _{n,i}$ \ and $\gamma ^{\ast }=\{\lambda \in C:\overline{\lambda }\in \gamma \}.$ Since for $\lambda \in \gamma $ the operator $(L(q)-\lambda I)^{-1}$ is bounded and continuously depend on $\lambda $ the operator $P$ is bounded operator. Similarly $T$ \ is bounded. For every $f\in $ $L_{2}[0,1]$ there exists sequence $f_{n}$ consisting of finite lineer combinations of eigenfunctions and associated functions of $L(q)$ and converging to $f.$ Clearly $Pf_{n}\in E,Pf\in E$ and $P^{2}f=Pf,$ that is $P$ is projector of $% L_{2}[0,1]$ onto $E.$ Similarly $T$ is projector of $L_{2}[0,1]$ onto $% E^{\ast }.$ Considering the integrals in (22) as the limit of finite integral sums one can easily verify that $T=P^{\ast }.$ Therefore ( see [4, p. 2313]) for the basis $\Psi _{n,i,p}(x),$ $p=0,1,2,...,t(n,i)$ of $E$ one can find a unique basis $\Phi _{n,i,p}(x),$ $p=0,1,2,...,t(n,i)$ of $E^{\ast }$ biorthogonal to the above basis of $E$. Thus \ the biortogonal system of the system $\{\Psi _{n,i,p}(x)\}$ is $\{\Phi _{n,i,p}(x)\}.$\ Now to prove the theorem, it is enough to show that for every function $f(x)\in L_{2}[0,1] $ the inequalities \begin{equation} \sum\limits_{n=N}^{\infty }\sum\limits_{j=1}^{2}\left| (f,\Psi _{n,j})\right| ^{2}<\infty \text{ , }\sum\limits_{n=N}^{\infty }\sum\limits_{j=1}^{2}\left| (f,\Phi _{n,j})\right| ^{2}<\infty \end{equation} hold (see [1, p.310, Theorem 2.1]). By the asymptotic formula (14), we have \begin{eqnarray*} \sum\limits_{n\geq N}\sum\limits_{j=1,2}\left| (f,\Psi _{n,j})\right| ^{2} &\leq &c_{11}(\sum\limits_{n\geq N,j=1,2}\text{ }\mid (e^{-i\gamma _{j}x}f,e^{i2\pi nx})\mid ^{2}+ \\ &\mid &(e^{i\gamma _{j}x}f,e^{-i2\pi nx})\mid ^{2}+\sum\limits_{n\geq N}\left| (f,O(\frac{1}{n}))\right| ^{2})N,j=1,2, \end{equation} Therefore by (14), (24) we have \begin{equation} (G_{n,j}^{0},\Psi _{n,j}^{0})=h_{j}+O(\frac{1}{n}),\text{ }(G_{n,j},\Psi _{n,j})=h_{j}+O(\frac{1}{n}), \end{equation} where $h_{j}\neq 0.$ Thus from (24) we get \begin{equation} \Phi _{n,j}(x)=\frac{1}{h_{j}}(C_{j}e^{i(2\pi n+\eta _{j})x}+D_{j}e^{-i(2\pi n+\eta _{j})x})+O(\frac{1}{n}),j=1,2, \end{equation} from which follows the second inequality in (23) \end{proof} \section{The Asymptotic Formulas for Eigenfunctions and Eigenvalues} Let $\Phi _{k,j,p}^{0}(x),$ for $p=1,2,...,t(k,j),$ be the associated function of $L^{\ast }(0)$ corresponding to the eigenfunction $\Phi _{k,j}^{0}(x)=\Phi _{k,j,0}^{0}(x).$ Since only finite number of eigenvalues of $L(0)$ and $L(q)$ are multiple there is $n_{0}$ such that $\lambda _{n,1}^{0}$ and $\lambda _{n,1}$ are simple eigenvalues for $n\geq n_{0}.$ Multiplying the equation \begin{equation} \lambda _{n,i}\Psi _{n,i}=(L(0)+q(x))\Psi _{n,i} \end{equation} (see (17)) by $\Phi _{k,j}^{0}(x)$ we get \begin{equation} (\lambda _{n},_{i}-\lambda _{k}^{0},_{j})(\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))=(q(x)\Psi _{n,i}(x),\Phi _{k,j}^{0}(x)), \label{main1} \end{equation} \begin{equation} (\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))=\frac{(q(x)\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))}{\lambda _{n},_{i}-\lambda _{k}^{0},_{j}}, \end{equation} for\ $i,j=1,2.$ Now multiplying (28) by $\Phi _{k,j,1}^{0}(x)$ and using (19), (20) we get \[ (\lambda _{n},_{i}-\lambda _{k}^{0},_{j})(\Psi _{n,i},\Phi _{k,j,1}^{0})=(q(x)\Psi _{n,i},\Phi _{k,j,1}^{0})+\frac{(q(x)\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))}{\lambda _{n},_{i}-\lambda _{k}^{0},_{j}}, \] \[ (\Psi _{n,i}(x),\Phi _{k,j,1}^{0}(x))=\frac{(q(x)\Psi _{n,i}(x),\Phi _{k,j,1}^{0}(x))}{\lambda _{n},_{i}-\lambda _{k}^{0},_{j}}+\frac{(q(x)\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))}{(\lambda _{n},_{i}-\lambda _{k}^{0},_{j})^{2}}% . \] In this way one can redues the formulas \begin{equation} (\Psi _{n,i}(x),\Phi _{k,j,m}^{0}(x))=\sum_{p=0}^{m}\frac{(q(x)\Psi _{n,i}(x),\Phi _{k,j,p}^{0}(x))}{(\lambda _{n},_{i}-\lambda _{k}^{0},_{j})^{m+1-p}} \end{equation} for $m=0,1,2,...,t(k,j).$ Clearly \begin{equation} n_{0}N,$ where $N\gg 1.$ For $i=1,2$ the equality \[ (q(x)\Psi _{n,i},\Phi _{k,j,m}^{0})=\sum\limits_{j_{1}=1}^{2}\sum\limits_{k_{1}=0}^{n_{0}-1}% \sum_{m_{1}=1}^{t(k_{1},j_{1})}c(k,j,m,k_{1},j_{1},m_{1})(\Psi _{n,i},\Phi _{k_{1},j_{1},m_{1}}^{0}) \] \begin{equation} +\sum\limits_{j_{1}=1}^{2}\sum\limits_{k_{1}=n_{0}}^{\infty }c(k,j,m,k_{1},j_{1},0)(\Psi _{n,i}(x),\Phi _{k_{1},j_{1},0}^{0}(x)) \end{equation} holds, where $\overline{c(k,j,m,k_{1},j_{1},m_{1})}=(\overline{q}(x)\Phi _{k,j,m}^{0}(x),\Psi _{k_{1},j_{1},m_{1}}^{0}(x)).$ Moreover \begin{equation} \mid c(k,j,m,k_{1},j_{1},m_{1})\mid N\gg 1$. It is well known that (see [2, Chap.1, Sec.3]), $q(x)\Psi _{n,i}(x)\in L_{1}[0,1]$ and $(q(x)\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))\rightarrow 0$ as $k\rightarrow \infty .$ Therefore there is a constant $C(n,i)$ and $k_{0},j_{0},m_{0}$ such that \begin{equation} \stackunder{k,j,m}{\max }\left| (q(x)\Psi _{n,i}(x),\Phi _{k,j,m}^{0}(x))\right| =\left| (q(x)\Psi _{n,i}(x),\Phi _{k_{0},j_{0},m_{0}}^{0}(x))\right| =C(n,i). \label{max1} \end{equation} Then by (30), (34), and (10), (11) we have \[ \left| (\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))\right| <\frac{C(n,i)}{\left| \lambda _{n,i}-\lambda _{k,j}^{0}(x)\right| },\forall k\geq n_{0},(n,i)\neq (k,j). \] Therefore using (12) we get \[ \sum\limits_{j=1}^{2}\sum\limits_{k>2n_{1}}\left| (\Psi _{n,i}(x),\Phi _{k,j}^{0}(x))\right| <\sum\limits_{k>2n_{1}}\frac{2C(n,i)}{k^{2}}<\frac{% c_{18}C(n,i)}{n_{1}}, \] where $n_{1}>n,$ hence $k>2n_{1}>2n.$ This with (38) imply that the decomposition of $\Psi _{n,i}(x)$ by the basis $\{\Psi _{k,j,m}^{0}(x)\}$ is of the form \begin{equation} \Psi _{n,i}(x)=\sum\limits_{j_{1}=1}^{2}\sum\limits_{k_{1}\leq 2n_{1}}\sum_{m_{1}=0}^{t(k_{1}j_{1})}(\Psi _{n,i}(x),\Phi _{k_{1},j_{1},m_{1}}^{0}(x))\Psi _{k_{1},j_{1},m_{1}}^{0}(x)+g_{i}(x), \label{decomp} \end{equation} where$\stackunder{x\in \lbrack 0,1]}{\sup }\left| g_{i}(x)\right| <\frac{% c_{17}c_{18}C(n,i)}{n_{1}}.$ Using this in $(q(x)\Psi _{n,i}(x),\Phi _{k,j,m}^{0}(x))$ and tending $n_{1}\rightarrow \infty $, we obtain (33). Now we prove (36). Using (39), (33), (34), (35) and (32) we get \[ C(n,i)\leq \sum\limits_{j_{1}=1}^{2}\sum\limits_{k_{1}=n_{0}}^{\infty }\mid c(k_{0},j_{0},m_{0},k_{1},j_{1},0)(\Psi _{n,i}(x),\Phi _{k_{1},j_{1},0}^{0}(x))\mid +c_{19} \] Isolating the term with multiplicant $(\Psi _{n,i}(x),\Phi _{n,i}^{0}(x))$ and using (30) for $(\Psi _{n,i}(x),\Phi _{k_{1},j_{1}}^{0}(x))$ with $% (k_{1},j_{1})\neq (n,i)$ we obtain \ \[ C(n,i)\leq c_{20}+\sum_{k_{1},j_{1}:(k_{1},j_{1})\neq (n,i)}\frac{% c_{21}C(n,i)}{\mid \lambda _{n,i}-\lambda _{k_{1},j_{1}}^{0}\mid }. \] Using (10), (11) it is not hard to verify that \begin{equation} \sum_{k,j,(k,j)\neq (n,i)}\frac{1}{\left| \lambda _{n,j}-\lambda _{k,j}^{0}\right| }=O(\frac{\ln n}{n}). \end{equation} Since $n\gg 1,$ the last two relations give $C(n,1)1$, $S_{1}(\lambda _{n},_{i})$ is defined in (49) and \[ S_{l}(\lambda _{n}{}_{,i})=\sum\limits_{j_{1},j_{2},...,j_{l}}\sum% \limits_{k_{1},k_{2},...,k_{l}}\sum% \limits_{p_{1},p_{2},...,p_{l}}a(n,i,0,k_{1},j_{1},p_{1})a(k_{1},j_{1},p_{1} ,k_{2},j_{2},p_{2})\times \] \begin{equation} a(k_{l-1},j_{l-1},p_{l-1},k_{l},j_{l},p_{l})c(k_{l},j_{l},p_{l},n,i,0) \end{equation} \[ C_{l}=\sum\limits_{j_{1},j_{2},...,j_{l}}\sum\limits_{k_{1},k_{2},...,k_{l}} % \sum% \limits_{p_{1},p_{2},...,p_{l}}a(n,i,0,k_{1},j_{1},p_{1})a(k_{1},j_{1},p_{1} ,k_{2},j_{2},p_{2})\times \] \begin{equation} a(k_{l-2},j_{l-2},p_{l-2},k_{l-1},j_{l-1},p_{l-1})b(_{l-1},j_{l-1},p_{l-1},k _{l},j_{l},p_{l}) \end{equation} for $l>1.$ Here the sums are taken under the conditions $% j_{1},j_{2},...,j_{l}=1,2;$ $k_{1},k_{2},...,k_{l}=1,2,...,\infty ;$ and $% p_{s}=0,1,...,t(k_{s},j_{s})$ for $s=1,2,....$ Moreover $(k_{s},j_{s})$ $% \neq (n,i).$ From the formula (51) we obtain the following \begin{theorem} The eigenvalue $\lambda _{n,i}$ satisfies \begin{equation} \lambda _{n,i}=\lambda _{n,i}^{0}+c(n,i)+F_{m}+O\left( (\frac{\ln n}{n}% )^{m}\right) ,\ \forall m=1,2,.... \end{equation} where $F_{1}=0,$ $F_{k}=A_{k}(\lambda _{n,i}^{0}+c(n,i)+F_{k-1}),$ \ $% \forall k=2,3,....$ Moreover for $m=2$ the formula (54)\ can be written in the form \[ \lambda _{n,i}=\lambda _{n,i}^{0}+c(n,i)+\sum\limits\Sb (k,j)\neq (n,i), \\ % k\geq n_{0} \endSb \frac{c(n,i,0,k,j,0)c(k,j,0,n,i,0)}{\lambda _{n},_{i}-\lambda _{k,,j}^{0}}+O\left( (\frac{\ln n}{n})^{2}\right) \] \end{theorem} \begin{proof} \bigskip It follows from (45)- (48), (32), (34), and (36) that \begin{eqnarray*} &\mid &a(k,j,m,k_{1},j_{1},p_{1})\mid \leq \mid \frac{c_{22}}{\mid \lambda _{n}{}_{,i}-\lambda _{k_{1},j_{1}}^{0}\mid }, \\ &\mid &b(k,j,m,k_{1},j_{1},p_{1})\mid \leq \frac{c_{22}}{\mid \lambda _{n}{}_{,i}-\lambda _{k_{1},j_{1}}^{0}\mid }, \end{eqnarray*} for all $k,j,m,k_{1},j_{1},p_{1},$ $(k_{1},j_{1})\neq (n,i).$ Therefore using (41) we get \ \begin{equation} S_{m}=O\left( (\frac{\ln n}{n})^{m}\right) ,C_{m}=O\left( (\frac{\ln n}{n}% )^{m}\right) ,\forall m=1,2,... \end{equation} Since both $\Psi _{n,i}(x)$ and $\Psi _{n,i}^{0}(x)$ satisfy the formula (14) we have \begin{equation} \Psi _{n,i}(x)=\Psi _{n,i}^{0}(x)+O(\frac{1}{n}),(\Psi _{n,i},\Phi _{n,i}^{0})=1+O(\frac{1}{n}) \end{equation} Hence dividing both part of (43) by $(\Psi _{n,i},\Phi _{n,i}^{0})$ and using (55) we get (54) for $m=1.$ Similarly dividing both part of (51) by $% (\Psi _{n,i},\Phi _{n,i}^{0})$ \ we obtain \begin{equation} (\lambda _{n}{}_{,i}-\lambda _{n}^{0}{}_{,i})=c(n,i)+A_{m}(\lambda _{n}{}_{,i})+O\left( (\frac{\ln n}{n})^{m}\right) . \end{equation} First we consider \ this formula for $m=2.$ Clierly for $k_{1}