Compressed Resolvents and Reduction of Spectral Problems on Star Graphs
- 261 Downloads
Abstract
In this paper a two-step reduction method for spectral problems on a star graph with \(n+1\) edges \(e_{0}, e_{1}, \ldots , e_{n}\) and a self-adjoint matching condition at the central vertex v is established. The first step is a reduction to the problem on the single edge \(e_0\) but with an energy depending boundary condition at v. In the second step, by means of an abstract inverse result for Q-functions, a reduction to a problem on a path graph with two edges \(e_0\), \(\widetilde{e}_1\) joined by continuity and Kirchhoff conditions is given. All results are proved for symmetric linear relations in an orthogonal sum of Hilbert spaces. This ensures wide applicability to various different realizations, in particular, to canonical systems and Krein strings which include, as special cases, Dirac systems and Stieltjes strings. Employing two other key inverse results by de Branges and Krein, we answer e.g. the following question: If all differential operators are of one type, when can the reduced system be chosen to consist of two differential operators of the same type?
Keywords
Compressed resolvent Differential operator Star graph Path graph Canonical system Krein string Stieltjes stringMathematics Subject Classification
47A10 34B45 34L401 Introduction
Direct and inverse spectral problems for differential operators on graphs have attracted rapidly increasing interest over the last two decades. Since it seems impossible to list all relevant literature, we mention the three monographs [5, 26, 33], whose bibliographies reflect the intense research activities in this area very well. The vast majority of these works considers concrete operators on graphs such as Sturm–Liouville and Schrödinger operators [2, 3, 16, 25, 37], Laplace and related operators [29], Krein strings [13], Stieltjes strings [30, 31], canonical systems and Dirac operators [1, 8]. More abstract methods were developed in [32, 35].
The results of this paper are of more conceptual character; they have implications to many differential and difference operators, and corresponding Cauchy problems. We establish abstract results for compressed resolvents of self-adjoint extensions of an orthogonal sum of symmetric relations. They allow us to reduce spectral problems on a star graph with self-adjoint matching condition at the central vertex to problems on a single edge and, using inverse results, to problems on a path graph with two edges, see Fig. 1a–c. We also address the question whether for concrete realizations the reduced system on the path graph can be chosen of the same type as the original system.
The two-step reduction of a star graph to a path graph (color figure online)
To exemplify our results, we consider a star graph \({\mathcal G}\) as in Fig. 1a and we suppose that for each edge \(e_j\) a symmetric linear relation \(T_j\) with equal defect numbers 1 in a Hilbert space \({\mathcal H}_j\) and a boundary triplet \(({\mathbb C},\Gamma _{j1},\Gamma _{j2})\), \(j=0,1,\ldots ,n\). are given. For concrete realizations, e.g. differential or difference operators as considered below, the boundary triplets depend on the values and derivatives of the functions at the central vertex v. The Hilbert space \({\mathcal H}\) on the graph \({\mathcal G}\) is the orthogonal sum of the Hilbert spaces \({\mathcal H}_j\) on the edges \(e_j,\, j=0,1,\ldots ,n\), and in \({\mathcal H}\) the symmetric relation \(\mathbf T\) is defined as the orthogonal sum of the relations \(T_j\). The self-adjoint extensions of \(\mathbf T\) in \({\mathcal H}\) are all described by an interface condition parametrized by two \((n+1)\times (n+1)\)-matrices \({\mathcal A}\), \({\mathcal B}\) (see (2.12), (2.13) below); we denote them by \(\mathbf{T}_{{\mathcal A},{\mathcal B}}\) and we fix such an extension.
An inverse result of de Branges [7, Thm. 7] (see also [38, Thm. 1]) yields that every (scalar) Nevanlinna function is the Titchmarsh-Weyl function of a canonical system. Hence \(\widetilde{T}_1\) can even be chosen to be generated by a trace normed canonical system on the interval \([0,\infty )=\widetilde{e}_1\). Thus, in the graph situation, the compressed resolvent of \(\mathbf{T}_{{\mathcal A},{\mathcal B}}\) on the star graph, see (1.1), coincides with the compressed resolvent of an operator \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) on a path graph with two edges \(e_0\), \(\widetilde{e}_1\), given by \(T_0\) on \(e_0\) and a canonical system on \(\widetilde{e}_1\) joined by continuity and Kirchhoff conditions at v, see Theorem 4.6. If, in particular, \(T_0\) is given by a canonical system while all other relations \(T_j\), \(j=1,2,\ldots ,n\), are arbitrary, then \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) is given by two canonical systems on the path graph \(\widetilde{\mathcal G}\), see Corollary 5.1.
While the mass distribution function of a Krein string is an arbitrary non-decreasing (left-continuous) function, for a Stieltjes string it is a step function with steps accumulating at most at the right end-point. For a star graph of Stieltjes strings as in Fig. 1a this means that the steps on the edges \(e_j\) accumulate at most at the outer vertices \(v_j\), \(j=0,1,\ldots ,n\). Employing asymptotic expansions of Titchmarsh-Weyl functions, it follows that if, in the above situation, inequality (1.5) holds and all \(T_j\), \(j=0,1,\ldots ,n\), are Stieltjes strings, the reduced system \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) consists of two Stieltjes strings, see Corollary 5.4.
If all \(T_j\), \(j=0,1,\ldots ,n\), are Sturm–Liouville operators, Theorem 3.1 yields a reduction to the single Sturm–Liouville operator \(T_0\) on \(e_0\) with an energy depending boundary condition at v. Moreover, Theorem 4.6 based on de Branges’ inverse result applies and yields a reduction to a problem on a path graph of two edges for the Sturm–Liouville operator \(T_0\) and a canonical system \(\widetilde{T}_1\) with standard interface conditions. However, due to a lack of corresponding inverse results for Sturm–Liouville operators, the latter can in general not be replaced by a Sturm–Liouville operator.
The paper is organized as follows. In Sect. 2 we present some general results on Krein’s formula for an orthogonal sum of \(n+1\) symmetric linear relations \(T_j\), \(j=0,1,\ldots ,n\), each with equal defect numbers 1. In Sect. 3 we establish the relation between the compressed resolvent (1.1) and the extension of \(T_0\) determined by an energy depending boundary condition, see (1.2). Section 4 contains the main result where we use the abstract inverse result on Q-functions to ‘replace’ the z-depending boundary condition by a new symmetric linear relation \(\widetilde{T}_1\) or, correspondingly, by attaching a realization on a new edge \(\widetilde{e}_1\), see Fig. 1c. Implications of our reduction result for Cauchy problems of first and second order on star graphs are given in Sect. 4.2. In Sect. 5 we consider concrete operators on star graphs: canonical systems and Krein strings, which cover Dirac operators and Stieltjes strings as special cases, respectively, and Sturm–Liouville operators. Combining our results with the inverse results of de Branges and Krein, we obtain a more complete picture of the structure of the reduced problems.
2 The Abstract Schema
In this section we set up a framework for pasting a finite number of symmetric linear relations with general self-adjoint interface conditions, comp. [35]. For this we need some notation and basic properties of symmetric linear relations, boundary triplets and corresponding Titchmarsh-Weyl functions, see e.g. [34, Chapter 14], [9, 10].
2.1 The Titchmarsh-Weyl Function of an ‘Edge’
2.2 The ‘Interface’ Condition
2.3 A Version of M. G. Krein’s Resolvent Formula
The resolvent of any self-adjoint relation \(\mathbf{T}_{{\mathcal A},{\mathcal B}}\) is related to the resolvent of the particular self-adjoint relation \(\mathbf{T}_{0,{\mathcal I}}\) with the special ‘interface’ condition \(\Gamma _2\mathbf{y}=0\) as follows.
Proposition 2.1
Proof
3 The Compressed Resolvent
Theorem 3.1
Remark 3.2
Proof of Theorem 3.1
Theorem 3.1 may be used to relate the spectra of \(\mathbf{T}_{{\mathcal A},{\mathcal B}}\) in \({\mathcal H}\) and of the z-depending spectral problem (3.2) in \({\mathcal H}_0\). For simplicity, we consider the case that the self-adjoint extensions of the \(T_j\) have discrete spectra, hence all the functions \(m_j\) and also \(n_0\) are meromorphic. Note that, if \(z=\lambda \) is a pole of \(n_0\), then the second relation in (3.2) becomes \(\Gamma _{02}g=0\); if \(z=\lambda \) is a zero of \(n_0\) or if \(n_0 \equiv 0\), then the second relation in (3.2) becomes \(\Gamma _{01}g=0\).
Corollary 3.3
- (i)
If \(z\in \rho (\mathbf{T}_{{\mathcal A},{\mathcal B}})\), then for each \(f_0\in {\mathcal H}_0\) the problem (3.2) has the solution \(g_0\) given by (3.1).
- (ii)If \(\lambda \in \sigma _\mathrm{p}(\mathbf{T}_{{\mathcal A},{\mathcal B}})\) with eigenvector \(\mathbf{g}\) and \(n_0\) is holomorphic in a neighbourhood of \(\lambda \), thenis satisfied for \(z=\lambda \), and \(\lambda \) is an eigenvalue of (3.5) with eigenvector \({P}_0\mathbf{g}\) if \({P}_0\mathbf{g}\ne 0\).$$\begin{aligned} \{{P}_0\mathbf{g},0\} \in (T_0^*-z), \quad \Gamma _{01}{P}_0\mathbf{g}-n_0(z)\Gamma _{02}{P}_0\mathbf{g}=0, \end{aligned}$$(3.5)
Proof
Example 3.4
In the special case of Robin type interface conditions, the properties of \(n_0\) proved in Theorem 3.1 for general interface conditions (2.12) are immediate from the explicit formula (3.8) for \(n_0: n_0\) does not depend on \(m_0\), and since \(m_1, m_2, \ldots , m_n\) are Nevanlinna functions so is \(n_0\) for every \(\tau \in \mathbb R\cup \{\infty \}\).
Remark 3.5
Remark 3.6
4 The Reduced System and its Realization via de Branges’ Inverse Theorem
4.1 The Reduced System
In the following theorem we show that there exists a single linear relation \(\widetilde{T}_1\) with Titchmarsh–Weyl function \(\widetilde{m}_1=n_0\) such that the compressed resolvents on \({\mathcal H}_0\) of the corresponding self-adjoint extensions of \(\mathbf{T} = T_0 \oplus T_1 \oplus \cdots \oplus T_n\) with interface condition (2.12) and of \(\widetilde{\mathbf{T}} = T_0 \oplus \widetilde{T}_1\) with continuity and Kirchhoff type conditions (3.9) coincide. The main tool here is an abstract inverse result [28].
Theorem 4.1
Proof
According to a general inverse result for Q-functions (or Titchmarsh-Weyl functions), see [28, 11, Thm. 5.1], for the given Nevanlinna function \(n_0\) in (3.3) there exists a symmetric linear relation \(\widetilde{T}_1\) with equal defect numbers 1 and a boundary triplet \((\mathbb C,\widetilde{\Gamma }_{11}, \widetilde{\Gamma }_{12})\) for \(\widetilde{T}_1^*\) such that the corresponding Titchmarsh-Weyl function \(\widetilde{m}_1\) coincides with \(n_0\), i.e. \(\widetilde{m}_1=n_0\). On the other hand, by Remark 3.5, the function \(\widetilde{n}_0\) in (3.2) for \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) is given by \(\widetilde{n}_0=\widetilde{m}_1\) and hence \(\widetilde{n}_0=n_0\). Since \(\mathbf{T}\) and \(\widetilde{\mathbf{T}}\) have the same first component \(T_0\), this implies that the problems (3.2) for \(\mathbf{T}_{{\mathcal A},{\mathcal B}}\) and \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) coincide and hence Theorem 3.1 yields the claim. \(\square \)
Remark 4.2
Corollary 4.3
- (i)For \(z\in {\mathbb C}{\setminus }{\mathbb R}\) and \(f_0\in {\mathcal H}_0\), the first components of the solutions \(\mathbf{g}\) and \(\widetilde{\mathbf{g}}\) of the inhomogeneous equationscoincide, that is, \(P_0\mathbf{g}=\widetilde{P}_0\widetilde{\mathbf{g}}\).$$\begin{aligned} (\mathbf{T}_{{\mathcal A},{\mathcal B}}-z)\mathbf{g}=\begin{pmatrix}f_0\\ 0\\ \vdots \\ 0\end{pmatrix},\quad (\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}-z)\widetilde{\mathbf{g}}=\begin{pmatrix}f_0\\ 0\end{pmatrix} \end{aligned}$$
- (ii)
If \(\lambda \in \sigma _\mathrm{p}(\mathbf{T}_{{\mathcal A},{\mathcal B}})\) with eigenvector \(\mathbf{g}\) and \({P}_0\mathbf{g}\ne 0\), then \(\lambda \in \sigma _\mathrm{p}(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0})\) with eigenvector \(\widetilde{\mathbf{g}}\) such that \(\widetilde{P}_0\widetilde{\mathbf{g}}=P_0\mathbf{g}\), and vice versa if \(\widetilde{P}_0\widetilde{\mathbf{g}}\ne 0\).
Proof
The claim in (i) is immediate from Theorem 4.1. The claims in (ii) follow from the fact that for a self-adjoint operator A in a Hilbert space with eigenvalue \(\lambda _0\in {\mathbb R}\) the orthogonal projection \(P_{\lambda _0}\) onto the corresponding eigenspace may be obtained as the strong limit \(P_{\lambda _0}=\text{ s }-\lim \nolimits _{z\rightarrow \lambda _0}(z_0-z)(A-z)^{-1}\) if z tends to \(\lambda _0\) perpendicularly to \({\mathbb R}\).\(\square \)
4.2 Implications for Cauchy Problems
Theorem 4.4
Proof
Remark 4.5
(i) Clearly, the equality (4.5) continues to hold under more general assumptions on \(x_0\) and \(f_0\) if only the solutions can be expressed by a relation of the form (4.6), see e.g. [24, Thm. I.6.5].
ii) A corresponding result can be proved for the solutions of second order Cauchy problems if \(\mathbf{T}_{{\mathcal A},{\mathcal B}}\) and \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) generate cosine families, see e.g. [14] and also [36, Props. 13.2.2 and 2.3.1].
4.3 Realization by a Canonical System
In this subsection we employ a fundamental inverse theorem by de Branges [7, Thm. 7], [38, Thm. 1] to show that the linear relation \(\widetilde{T}_1\) in the reduced system in Theorem 4.1 can always be realized as a canonical system. To this end, we first need to recall some basic results for canonical systems on a single edge, where we omit the subscript j for simplicity.
In the following theorem we make use of a fundamental inverse result of de Branges, see [7, Thm. 7], [38, Thm. 1]. Given any Nevanlinna function m, there exists a Hermitian H on \([0,\infty )\) with the properties mentioned at the beginning of this subsection such that m is the Titchmarsh-Weyl function of the canonical system corresponding to H. This means that the symmetric linear relation in Theorem 4.1 can always be chosen to be a canonical system.
Theorem 4.6
5 Differential Operators on Star Graphs
In this section we investigate the structure of the reduced problem if the symmetric relations \(T_j\), \(j=0,1,\ldots ,n\), (or some of them) are differential or difference operators on the edges of a star graph \({\mathcal G}\). In particular, we ask e.g. the following question: If all differential operators are of one type, when can the reduced system constructed in the previous section be chosen to consist of two differential operators of the same type?
Star graph \({\mathcal G}\) with \(n+1\) edges and distinguished edge \(e_0\) (color figure online)
The symmetric relations \(T_j\), \(j=0,1,\ldots ,n\), in the Hilbert spaces \({\mathcal H}_j\) may be induced by three different types of differential or difference expressions on the edges \(e_j\), which are all supposed to be regular at v: canonical systems, Sturm–Liouville problems, and Krein strings; they include, as special cases, Dirac systems and Stieltjes strings. Here the ‘interface’ condition (2.12) becomes a matching condition at the common vertex v.
5.1 Canonical Systems and Sturm–Liouville Operators
By Theorem 4.6, independently of the type of differential operators inducing the relations \(T_j\), \(j=1,2,\ldots ,n\) on the subgraph of edges different from \(e_0\) (the blue/non-bold part in Fig. 2), the new part \(\widetilde{T}_1\) of the reduced system \(\widetilde{\mathbf{T}}_{{\mathcal A}_0,{\mathcal B}_0}\) extending \(T_0\oplus \widetilde{T}_1\) can always be chosen to be a canonical system on an edge \(\widetilde{e}_1\).
Thus the answer to the question when the whole reduced system consists of two canonical systems is immediate from Theorem 4.6: it suffices that only \(T_0\) is induced by a canonical system.
Corollary 5.1
Reduced path graph of two canonical systems (color figure online)
We mention that two canonical systems coupled at a common end-point, as obtained in Corollary 5.1, were studied in [17, Sect. 6].
For Sturm–Liouville operators the situation is very different. Since there are no suitable inverse results in this case, one cannot hope to be able to reduce a system of Sturm–Liouville operators on a star graph with \(n+1\) edges to a system of two Sturm–Liouville operators on a path graph. Nevertheless the results of the two previous sections can still be applied.
Theorem 3.1 reduces the system of Sturm–Liouville operators \(T_j\), \(j=0,1,\ldots ,n\), on the star graph \({\mathcal G}\) with arbitrary interface conditions (2.12) to the Sturm–Liouville operator \(T_0\) on the single edge \(e_0\) with an energy depending boundary condition at v given by (3.3) with \(m_j\) as in (5.6), \(j=1,2,\ldots ,n\).
According to Theorem 4.6, the energy depending boundary condition at v can be replaced by attaching a canonical system on a new edge \(\widetilde{e}_1\). In order to decide whether the latter can be replaced by a Sturm–Liouville operator would require suitable inverse results for Sturm–Liouville problems, which to the best of our knowledge do not exist.
5.2 Krein Strings and Stieltjes Strings
For a star graph of \(n+1\) Krein strings, the reduced system need not consist of two Krein strings. For the case of Robin type interface conditions, we establish a sufficient condition involving the Titchmarsh–Weyl functions of the n Krein strings of the subsystem and the Robin parameter. In this subsection we strongly rely on [18] (see also [20]).
A Krein string is given by a bounded left-continuous and non-decreasing function M on an (edge) interval \(e=[0,\ell ),\,0<\ell \le \infty \); here, for \(x\in (0,\ell )\), the value M(x) is considered to be the mass of the string on the interval [0, x). We assume that the string has no concentrated mass at 0 and at \(\ell \) so that, in particular, \(M(0)=M(0+)=0\), and we set \(M(\ell ):=\lim _{x\nearrow \ell }M(x)\le \infty \).
(ii) \(\underline{\ell +M(\ell )=\infty }\): In the singular case we have to distinguish two subcases. If \(\int _0^\ell x^2\,\mathrm{d}M(x)=\infty \), then the limit point case prevails at \(\ell \) and no boundary condition at \(\ell \) is needed. The Krein string with the boundary condition \(f'(0)=0\) induces a unique self-adjoint operator S in \(L^2_M(e,{\mathbb C})\) which is non-negative. The corresponding Titchmarsh-Weyl function belongs to the Stieltjes class \({\mathcal S}\), see [18, Thm. 10.1]. If \(\int _0^\ell x^2\,\mathrm{d}M(x)<\infty \), then the limit circle case prevails at \(\ell \). In this case, the Krein string with the boundary condition \(f'(0)=0\) induces infinitely many self-adjoint operators in \(L^2_M(e,{\mathbb C})\), but only one non-negative self-adjoint operator S in \(L^2_M(e,{\mathbb C})\).
As a consequence of this inverse result, if the function \(n_0\) from (3.3) in Theorem 3.1 belongs to the class \({\mathcal S}\), then the operator \(\widetilde{T}_1\) can be chosen to be a Krein string on some interval \(\widetilde{e}_1=[0,\widetilde{\ell }_1)\). In the following theorem we establish a sufficient condition for \(n_0\in {\mathcal S}\) for the case of Robin type interface conditions.
Theorem 5.2
Proof
Finally, if \(T_0\) is also supposed to be given by a Krein string, then the last claim follows from the fact that Kirchhoff conditions on the vertex v joining \(e_0\) and \(\widetilde{e}_1\) amount to the required continuity conditions at v. \(\square \)
Remark 5.3
Corollary 5.4
Notes
Acknowledgements
Open access funding provided by TU Wien (TUW). All authors thank for the support of EPSRC, Grant No. EP/1038217/1, which enabled this work to commence in January 2013. The last author gratefully acknowledges the support of the Swiss National Science Foundation, SNF, Grant No. 169104.
References
- 1.Adamyan, V., Langer, H., Tretter, C., Winklmeier, M.: Dirac–Krein systems on star graphs. Integral Equ. Oper. Theory 86(1), 121–150 (2016)MathSciNetzbMATHGoogle Scholar
- 2.Avdonin, S., Kurasov, P.: Inverse problems for quantum trees. Inverse Probl. Imaging 2(1), 1–21 (2008)MathSciNetzbMATHGoogle Scholar
- 3.Avdonin, S., Kurasov, P., Nowaczyk, M.: Inverse problems for quantum trees II: recovering matching conditions for star graphs. Inverse Probl. Imaging 4(4), 579–598 (2010)MathSciNetzbMATHGoogle Scholar
- 4.Avdonin, S., Nicaise, S.: Source identification problems for the wave equation on graphs. Inverse Probl. 31(9), 095007, 29 (2015)MathSciNetzbMATHGoogle Scholar
- 5.Berkolaiko, G., Kuchment, P.: Introduction to Quantum Graphs, Mathematical Surveys and Monographs, vol. 186. American Mathematical Society, Providence (2013)zbMATHGoogle Scholar
- 6.Calkin, J.W.: Abstract symmetric boundary conditions. Trans. Am. Math. Soc. 45(3), 369–442 (1939)MathSciNetzbMATHGoogle Scholar
- 7.de Branges, L.: Some Hilbert spaces of entire functions IV. Trans. Am. Math. Soc. 105, 43–83 (1962)MathSciNetzbMATHGoogle Scholar
- 8.de Snoo, H., Winkler, H.: Canonical systems of differential equations with self-adjoint interface conditions on graphs. Proc. R. Soc. Edinb. Sect. A 135(2), 297–315 (2005)MathSciNetzbMATHGoogle Scholar
- 9.Derkach, V.: Boundary triplets, Weyl functions, and the Kreĭn formula. In: Alpay, D. (ed.) Operator Theory, pp. 183–218. Springer, Basel (2015)Google Scholar
- 10.Derkach, V.A., Malamud, M.M.: Generalized resolvents and the boundary value problems for Hermitian operators with gaps. J. Funct. Anal. 95(1), 1–95 (1991)MathSciNetzbMATHGoogle Scholar
- 11.Derkach, V.A., Malamud, M.M.: The extension theory of Hermitian operators and the moment problem. J. Math. Sci. 73(2), 141–242 (1995)MathSciNetzbMATHGoogle Scholar
- 12.Dym, H., McKean, H.P.: Gaussian processes, function theory, and the inverse spectral problem. Probability and Mathematical Statistics, vol. 31. Academic Press, New York-London (1976)zbMATHGoogle Scholar
- 13.Eckhardt, J.: An inverse spectral problem for a star graph of Krein strings. J. Reine Angew. Math. 715, 189–206 (2016)MathSciNetzbMATHGoogle Scholar
- 14.Fattorini, H.O.: Second Order Linear Differential Equations in Banach Spaces, North-Holland Mathematics Studies, vol. 108. North-Holland Publishing Co., Amsterdam (1985)Google Scholar
- 15.Fleige, A., Winkler, H.: An indefinite inverse spectral problem of Stieltjes type. Integral Equ. Oper. Theory 87(4), 491–514 (2017)MathSciNetzbMATHGoogle Scholar
- 16.Güneysu, B., Keller, M., Schmidt, M.: A Feynman–Kac–Itô formula for magnetic Schrödinger operators on graphs. Probab. Theory Relat. Fields 165(1–2), 365–399 (2016)zbMATHGoogle Scholar
- 17.Hassi, S., De Snoo, H., Winkler, H.: Boundary-value problems for two-dimensional canonical systems. Integral Equ. Oper. Theory 36(4), 445–479 (2000)MathSciNetzbMATHGoogle Scholar
- 18.Kac, I.S., Kreĭn, M.G.: On the spectral functions of the string. Am. Math. Soc. Transl. (2) 103, 19–102 (1974)zbMATHGoogle Scholar
- 19.Kac, I.S., Kreĭn, M.G.: R-functions – Analytic functions mapping the upper half-plane into itself. Am. Math. Soc. Transl. (2) 103, 1–18 (1974)zbMATHGoogle Scholar
- 20.Kac, I.S., Pivovarchik, V.N.: On the density of the mass distribution of a string at the origin. Integral Equ. Oper. Theory 81(4), 581–599 (2015)MathSciNetzbMATHGoogle Scholar
- 21.Kato, T.: Perturbation Theory for Linear Operators, Classics in Mathematics. Springer, Berlin (1995). Reprint of the 1980 editionGoogle Scholar
- 22.Kočubeĭ, A.N.: Extensions of symmetric operators and of symmetric binary relations. Math. Zametki 17, 41–48, (1975). Engl. transl. Math. Notes 17(1), 2528 (1975)Google Scholar
- 23.Kramar Fijavž, M., Mugnolo, D., Sikolya, E.: Variational and semigroup methods for waves and diffusion in networks. Appl. Math. Optim. 55(2), 219–240 (2007)MathSciNetzbMATHGoogle Scholar
- 24.Kreĭn, S.G.: Linear Differential Equations in Banach Space, Translations of Mathematical Monographs, vol. 29. American Mathematical Society, Providence (1971)Google Scholar
- 25.Kurasov, P.: Schrödinger operators on graphs and geometry. I. Essentially bounded potentials. J. Funct. Anal. 254(4), 934–953 (2008)MathSciNetzbMATHGoogle Scholar
- 26.Kurasov, P.: Quantum Graphs: Spectral Theory and Inverse Problems. Oper. Theory Adv. Appl. (2018) (to appear, Birkhäuser Basel)Google Scholar
- 27.Langer, H., Schenk, W.: Generalized second-order differential operators, corresponding gap diffusions and superharmonic transformations. Math. Nachr. 148, 7–45 (1990)MathSciNetzbMATHGoogle Scholar
- 28.Langer, H., Textorius, B.: On generalized resolvents and \(Q\)-functions of symmetric linear relations (subspaces) in Hilbert space. Pac. J. Math. 72(1), 135–165 (1977)MathSciNetzbMATHGoogle Scholar
- 29.Lenz, D., Pankrashkin, K.: New relations between discrete and continuous transition operators on (metric) graphs. Integral Equ. Oper. Theory 84(2), 151–181 (2016)MathSciNetzbMATHGoogle Scholar
- 30.Pivovarchik, V., Rozhenko, N., Tretter, C.: Dirichlet-Neumann inverse spectral problem for a star graph of Stieltjes strings. Linear Algebra Appl. 439(8), 2263–2292 (2013)MathSciNetzbMATHGoogle Scholar
- 31.Pivovarchik, V., Tretter, C.: Location and multiplicities of eigenvalues for a star graph of Stieltjes strings. J. Differ. Equ. Appl. 21(5), 383–402 (2015)MathSciNetzbMATHGoogle Scholar
- 32.Pivovarchik, V., Woracek, H.: Sums of Nevanlinna functions and differential equations on star-shaped graphs. Oper. Matrices 3(4), 451–501 (2009)MathSciNetzbMATHGoogle Scholar
- 33.Post, O.: Spectral Analysis on Graph-Like Spaces. Lecture Notes in Mathematics vol. 2039. Springer, Berlin (2012)Google Scholar
- 34.Schmüdgen, K.: Unbounded Self-Adjoint Operators on Hilbert Space, Graduate Texts in Mathematics, vol. 265. Springer, Dordrecht (2012)zbMATHGoogle Scholar
- 35.Simonov, S., Woracek, H.: Spectral multiplicity of selfadjoint Schrödinger operators on star-graphs with standard interface conditions. Integral Equ. Oper. Theory 78(4), 523–575 (2014)zbMATHGoogle Scholar
- 36.Vasil’ev, V.V., Piskarev, S.I.: Differential equations in Banach spaces. II. Theory of cosine operator functions. J. Math. Sci. (N.Y.) 122(2), 3055–3174 (2004)MathSciNetzbMATHGoogle Scholar
- 37.von Below, J.: Sturm–Liouville eigenvalue problems on networks. Math. Methods Appl. Sci. 10(4), 383–395 (1988)MathSciNetzbMATHGoogle Scholar
- 38.Winkler, H.: The inverse spectral problem for canonical systems. Integral Equ. Oper. Theory 22(3), 360–374 (1995)MathSciNetzbMATHGoogle Scholar
Copyright information
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.