1 Introduction

In many quantum mechanical applications one considers particles moving in an external potential field which is localized near a set \(\Sigma \) of measure zero. Such strongly localized fields can be modeled by singular potentials that are supported on \(\Sigma \) only; of particular importance in this regard are \(\delta \) and \(\delta '\)-interactions. To be more precise, assume that \(\Sigma \) splits \({\mathbb R}^2\) into a bounded domain \(\Omega _{+}\) and an unbounded domain \(\Omega _{-} = {\mathbb R}^2 {\setminus } \overline{\Omega _{+}}\), and consider the formal Schrödinger differential expressions

$$\begin{aligned} \mathcal {H}_{\delta , \alpha } = - \Delta + \alpha \delta _{\Sigma } \quad \text {and} \quad \mathcal {H}_{\delta ', \alpha } = - \Delta + \alpha \delta '_{\Sigma },\quad \alpha \in {\mathbb R}. \end{aligned}$$
(1.1)

These singular perturbations of the free Schrödinger operator \(-\Delta \) are characterized by certain transmission conditions along the interface \(\Sigma \) for the functions in the operator domain. For \(\delta \)-interactions one considers functions \(f: {\mathbb R}^2 \rightarrow {\mathbb C}\) such that the restrictions \(f_{\pm } = f \upharpoonright \Omega _{\pm }\) satisfy the transmission conditions

$$\begin{aligned} f_{+} = f_{-} \quad \text {and} \quad -\frac{\alpha }{2} \big ( f_{+} + f_{-} \big ) = \big ( \partial _{\nu } f_{+} - \partial _{\nu } f_{-} \big ) \quad \text {on } \Sigma , \end{aligned}$$
(1.2)

while \(\delta '\)-interactions are modeled by the transmission conditions

$$\begin{aligned} f_{+} - f_{-} = -\frac{\alpha }{2 } \big ( \partial _{\nu } f_{+} + \partial _{\nu } f_{-} \big ) \quad \text {and} \quad \partial _{\nu } f_{+} = \partial _{\nu } f_{-} \quad \text {on } \Sigma ; \end{aligned}$$
(1.3)

here \(\partial _{\nu } f_{\pm }\) is the normal derivative and \(\nu = (\nu _1, \nu _2)\) the unit normal vector field on \(\Sigma \) pointing outwards of \(\Omega _{+}\). The spectra and resonances of the self-adjoint realizations associated with the formal expressions (1.1) in \(L^2({\mathbb R}^2)\) are well understood, see, e.g., [8, 9, 12, 13, 15,16,17,18, 24]. In particular, the essential spectrum is given by \([0, \infty )\) and the discrete spectrum consists of at most finitely many points for every interaction strength \(\alpha < 0\), while there is no negative spectrum if \(\alpha \ge 0\).

In contrast to the transmission conditions (1.2) and (1.3) we are interested in a new type of transmission conditions of the form

$$\begin{aligned} (\nu _1 + i \nu _2) \big ( f_{+} - f_{-} \big ) = - \alpha \big ( \partial _{\overline{z}} f_{+} + \partial _{\overline{z}} f_{-} \big ) \quad \text {and} \quad \partial _{\overline{z}} f_{+} = \partial _{\overline{z}} f_{-} \quad \text {on } \Sigma , \end{aligned}$$
(1.4)

where \(\alpha \in {\mathbb R}\) and \(\partial _{\overline{z}} = \frac{1}{2} ( \partial _1 + i \partial _2)\) is the Wirtinger derivative. In the sequel such jump conditions will be referred to as oblique transmission conditions. Note that the conditions (1.4) can be rewritten as

$$\begin{aligned} f_{+} - f_{-} = - \frac{\alpha }{2} \big ( \partial _{\nu } f_{+} + \partial _{\nu } f_{-} + i \partial _{t} f_{+} + i \partial _{t} f_{-} \big ) \quad \text {and} \quad \partial _{\overline{z}} f_{+} = \partial _{\overline{z}} f_{-} \quad \text {on } \Sigma , \end{aligned}$$
(1.5)

where \(\partial _t\) denotes the tangential derivative. Thus, on a formal level there is some analogy to the \(\delta '\)-transmission conditions in (1.3), but it will turn out that the properties of the corresponding self-adjoint realization in \(L^2({\mathbb R}^2)\) differ significantly from those of Schrödinger operators with \(\delta '\)-interactions.

To make matters mathematically rigorous, assume that the curve \(\Sigma \) is the boundary of a bounded and simply connected \(C^\infty \)-domain \(\Omega _+\) with open complement \(\Omega _{-} = {\mathbb R}^2 {\setminus } \overline{\Omega _{+}}\), denote the \(L^2\)-based Sobolev space of first order by \(H^1\), let \(\gamma _D^\pm :H^1(\Omega _{\pm })\rightarrow L^2(\Sigma )\) be the Dirichlet trace operators, and define for \(\alpha \in {\mathbb R}\) the Schrödinger operator with oblique transmission conditions by

$$\begin{aligned} \begin{aligned} T_{\alpha } f&= \left( - \Delta f_{+} \right) \oplus \left( - \Delta f_{-} \right) ,\\ \mathrm{dom\,}T_{\alpha }&= \Big \{ f \in H^1(\Omega _{+}) \oplus H^1(\Omega _{-}) \, \big | \, \partial _{\overline{z}}f_{+} \oplus \partial _{\overline{z}} f_{-} \in H^1({\mathbb R}^2), \\& (\nu _1 + i \nu _2) \bigl (\gamma _D^+ f_{+} - \gamma _D^- f_{-} \bigr ) = - \alpha \bigl ( \gamma _D^+ ( \partial _{\overline{z}} f_{+}) + \gamma _D^- ( \partial _{\overline{z}} f_{-} )\bigr ) \Big \}. \end{aligned} \end{aligned}$$
(1.6)

The next theorem is the main result in this paper. We discuss the spectral properties of the Schrödinger operators \(T_\alpha \) and, in particular, we show in item (ii) that for every \(\alpha <0\) the operator \(T_\alpha \) is necessarily unbounded from below and the discrete spectrum in \((-\infty ,0)\) is infinite and accumulates to \(-\infty \). In items (iii) and (iv) we shall make use of the potential operator \(\Psi _{\lambda }:L^2(\Sigma ) \rightarrow L^2({\mathbb R}^2)\) and the single layer boundary integral operator \(S(\lambda ):L^2(\Sigma ) \rightarrow L^2(\Sigma )\) defined in (2.2) and (2.4), respectively.

Theorem 1.1

For any \(\alpha \in {\mathbb R}\) the operator \(T_{\alpha }\) is self-adjoint in \(L^2({\mathbb R}^2)\) and the essential spectrum is given by

$$\begin{aligned} \begin{aligned} \sigma _{\textrm{ess}}(T_{\alpha }) = [0, \infty ). \end{aligned} \end{aligned}$$

Furthermore, the following statements hold:

  1. (i)

    If \(\alpha \ge 0\), then \(\sigma _{\textrm{disc}}(T_{\alpha }) = \emptyset \) and \(T_\alpha \) is a nonnegative operator in \(L^2({\mathbb R}^2)\).

  2. (ii)

    If \(\alpha < 0\), then \(\sigma _{\textrm{disc}}(T_{\alpha }) \) is infinite, unbounded from below, and does not accumulate to 0. Moreover, for every fixed \(n \in \mathbb {N}\) the n-th discrete eigenvalue \(\lambda _n \in \sigma _{\textrm{disc}}(T_{\alpha })\) (ordered non-increasingly) admits the asymptotic expansion

    $$\begin{aligned} \lambda _n = - \frac{4}{\alpha ^2} + \mathcal {O}(1) \quad \text {for } \alpha \rightarrow 0^{-}, \end{aligned}$$

    where the dependence on n appears in the \(\mathcal {O}(1)\)-term.

  3. (iii)

    For \(\lambda \in {\mathbb C}{\setminus } [0, \infty )\) the Birman-Schwinger principle is valid:

    $$\begin{aligned} \lambda \in \sigma _\textrm{p}(T_{\alpha }) \quad \Longleftrightarrow \quad 1 \in \sigma _\textrm{p} \big (\alpha \lambda S(\lambda ) \big ). \end{aligned}$$
  4. (iv)

    For \(\lambda \in \rho (T_{\alpha })={\mathbb C}{\setminus } ( [0, \infty ) \cup \sigma _\textrm{p}(T_{\alpha }))\) the operator \(I - \alpha \lambda S(\lambda )\) is boundedly invertible in \(L^2(\Sigma )\) and the resolvent formula

    $$\begin{aligned} (T_{\alpha } - \lambda )^{-1} = (- \Delta - \lambda )^{-1} +\alpha \Psi _{\lambda } \bigl ( I - \alpha \lambda S(\lambda ) \bigr )^{-1} \Psi _{\overline{\lambda }}^{*} \end{aligned}$$

    holds, where \(- \Delta \) is the free Schrödinger operator defined on \(H^2({\mathbb R}^2)\).

To illustrate the significance of Theorem 1.1 we show that Schrödinger operators with oblique transmission conditions arise naturally as non-relativistic limits of Dirac operators with electrostatic and Lorentz scalar \(\delta \)-interactions. To motivate this, consider one-dimensional Dirac operators with \(\delta '\)-interactions of strength \(\alpha \in \mathbb {R}\) supported in the point \(\Sigma =\{0\}\). These are first order differential operators in \(L^2(\mathbb {R})^2\) and the singular interaction is modeled by transmission conditions for functions in the operator domain, which for sufficiently smooth \(f = (f_1,f_2) \in L^2(\mathbb {R})^2\) are given by

$$\begin{aligned} f_1(0+)-f_1(0-) = i \frac{\alpha c}{2} \big ( f_2(0+)+f_2(0-) \big ) \quad \text {and} \quad f_2(0+) = f_2(0-), \end{aligned}$$
(1.7)

where \(c>0\) is the speed of light. It is known that the associated self-adjoint Dirac operators converge in the non-relativistic limit to a Schrödinger operator with a \(\delta '\)-interaction of strength \(\alpha \); cf. [2, 19] and also [10, 11] for generalizations. It is not difficult to see that (1.7) can be rewritten as the transmission conditions associated with a Dirac operator with a combination of an electrostatic and a Lorentz scalar \(\delta \)-interaction of strengths \(\eta =- \frac{\alpha c^2}{2}\) and \(\tau = \frac{\alpha c^2}{2}\), respectively, as they were studied in dimension one recently in [7] and in higher space dimensions in, e.g., [3, 5,6,7].

To find a counterpart of the above result in dimension two, consider a Dirac operator with electrostatic and Lorentz scalar \(\delta \)-shell interactions of strength \(\eta \) and \(\tau \), respectively, supported on \(\Sigma \), which is formally given by

$$\begin{aligned} \mathcal {A}_{\eta , \tau } = A_0 + \left( \eta I_2 + \tau \sigma _3 \right) \delta _{\Sigma }; \end{aligned}$$
(1.8)

here \(A_0\) is the unperturbed Dirac operator, \(I_2\) is the \(2 \times 2\)-identity matrix and \(\sigma _3 \in \mathbb {C}^{2 \times 2}\) is given in (3.1). The differential expression \(\mathcal {A}_{\eta , \tau }\) gives rise to a self-adjoint operator \(A_{\eta , \tau }\) in \(L^2(\mathbb {R}^2)^2\), see (3.3). If one chooses, as above, \(\eta = -\frac{\alpha c^2}{2}\) and \(\tau = \frac{\alpha c^2}{2}\) and computes the non-relativistic limit, then instead of a Schrödinger operator with a \(\delta '\)-interaction one gets the somewhat unexpected limit \(T_\alpha \). Of course, this is compatible with the one-dimensional result described above, as the one-dimensional counterparts of (1.3) and (1.5) coincide, since there are no tangential derivatives in \(\mathbb {R}\). However, in higher dimensions Schrödinger operators with oblique transmission conditions should be viewed as the non-relativistic counterparts of Dirac operators with transmission conditions generalizing (1.7). Related results on non-relativistic limits of three-dimensional Dirac operators with singular interactions can be found in [4, 5, 21]. The precise result about the non-relativistic limit described above is stated in the following theorem and shown in Sect. 3.

Theorem 1.2

Let \(\alpha \in {\mathbb R}\). Then for all \(\lambda \in \mathbb {C} {\setminus } \mathbb {R}\) one has

$$\begin{aligned} \lim _{c \rightarrow \infty } \left( A_{-\alpha c^2/2, \alpha c^2/2} - (\lambda + c^2 /2) \right) ^{-1} = \left( \begin{array}{cc} \left( T_{\alpha } - \lambda \right) ^{-1} &{} 0 \\ 0 &{} 0 \end{array} \right) , \end{aligned}$$

where the convergence is in the operator norm and the convergence rate is \(\mathcal {O}\left( \frac{1}{c} \right) \).

Notations Throughout this paper \(\Omega _{+} \subseteq {\mathbb R}^2\) is a bounded and simply connected \(C^{\infty }\)-domain and \(\Omega _{-} = {\mathbb R}^2 {\setminus } \overline{\Omega _{+}}\) is the corresponding exterior domain with boundary \(\Sigma = \partial \Omega _{-} = \partial \Omega _{+}\). The unit normal vector field on \(\Sigma \) pointing outwards of \(\Omega _+\) is denoted by \(\nu \). Moreover, for \(z \in {\mathbb C}{\setminus }[0, \infty )\) we choose the square root \(\sqrt{z}\) such that \(\text {Im} \sqrt{z} > 0\) holds. The modified Bessel function of order \(j \in {\mathbb N}_0\) is denoted by \(K_j\).

For \(s \ge 0\) the spaces \(H^s({\mathbb R}^2)^n\), \(H^s(\Omega _{\pm })^n\), and \(H^s(\Sigma )^n\) are the standard \(L^2\)-based Sobolev spaces of \({\mathbb C}^n\)-valued functions defined on \({\mathbb R}^2\), \(\Omega _{\pm }\), and \(\Sigma \), respectively. If \(n=1\) we simply write \(H^s({\mathbb R}^2)\), \(H^s(\Omega _{\pm })\), and \(H^s(\Sigma )\). For negative \(s < 0\) we define the spaces \(H^s({\mathbb R}^2)^n\) and \(H^s(\Sigma )^n\) as the anti-dual spaces of \(H^{-s}({\mathbb R}^2)^n\) and \(H^{-s}(\Sigma )^n\), respectively. We denote the restrictions of functions \(f: \mathbb {R}^2 \rightarrow \mathbb {C}^n\) onto \(\Omega _\pm \) by \(f_\pm \); in this sense we write \(H^1(\mathbb {R}^2 {{\setminus }} \Sigma )^n = H^1(\Omega _+)^n \oplus H^1(\Omega _-)^n\) and identify \(f \in H^1(\mathbb {R}^2 {{\setminus }} \Sigma )^n\) with \(f_+ \oplus f_-\), where \(f_\pm \in H^1(\Omega _\pm )^n\). In the following \(\gamma _D^\pm :H^1(\Omega _{\pm })\rightarrow L^2(\Sigma )\) denote the Dirichlet trace operators and we shall write \(\gamma _D:H^1(\mathbb R^2)\rightarrow L^2(\Sigma )\) for the Dirichlet trace on \(H^1(\mathbb R^2)\); sometimes these trace operators are also viewed as bounded mappings to \(H^{1/2}(\Sigma )\).

For a Hilbert space \(\mathcal {H}\) we write \(\mathcal {L}(\mathcal {H})\) for the space of all everywhere defined, linear, and bounded operators on \(\mathcal {H}\). Furthermore, the domain, kernel, and range of a linear operator T from a Hilbert space \(\mathcal {G}\) to \(\mathcal {H}\) are denoted by \(\text {dom}\, T\), \(\text {ker}\, T\), and \(\text {ran}\, T\), respectively. The resolvent set, the spectrum, the essential spectrum, the discrete spectrum, and the point spectrum of a self-adjoint operator T are denoted by \(\rho (T)\), \(\sigma (T)\), \(\sigma _{\text {ess}}(T)\), \(\sigma _{\text {disc}}(T)\), and \(\sigma _p (T)\). The eigenvalues of compact self-adjoint operators \(K \in \mathcal {L}(\mathcal {H})\) are denoted by \(\mu _n(K)\) and are ordered by their absolute values.

2 Proof of Theorem 1.1

In this section the main result of this paper will be proved. For this, some families of integral operators are used. Define for \(\lambda \in {\mathbb C}{\setminus } [0, \infty )\) the function \(L_{\lambda }\) by

$$\begin{aligned} L_{\lambda }(x) = \frac{\sqrt{ \lambda }}{2 \pi } K_1 \bigl ( - i \sqrt{ \lambda } |x| \bigr ) \frac{ x_1 - i x_2}{|x|}, \quad x = (x_1,x_2) \in \mathbb {R}^2 {\setminus } \{0\}, \end{aligned}$$
(2.1)

and the operator \(\Psi _{\lambda }: L^2(\Sigma ) \rightarrow L^2({\mathbb R}^2)\) by

$$\begin{aligned} \Psi _{\lambda } \varphi (x) = \int _{\Sigma } L_{\lambda }(x-y) \varphi (y) d \sigma (y), \quad \varphi \in L^2(\Sigma ), ~x \in {\mathbb R}^2{\setminus }\Sigma . \end{aligned}$$
(2.2)

Moreover, for \(\lambda \in {\mathbb C}{\setminus } [0, \infty )\) we make use of the single layer potential \(SL(\lambda ): L^2(\Sigma ) \rightarrow H^1(\mathbb {R}^2)\) and the single layer boundary integral operator \(S(\lambda ): L^2(\Sigma ) \rightarrow L^2(\Sigma )\) associated with \(-\Delta - \lambda \) that are defined by

$$\begin{aligned} SL(\lambda ) \varphi (x) = \int _\Sigma \frac{1}{2 \pi } K_0 \bigl ( - i \sqrt{\lambda } |x- y| \bigr ) \varphi (y) d \sigma (y), \quad \varphi \in L^2(\Sigma ),\quad x \in \mathbb {R}^2 {\setminus } \Sigma , \end{aligned}$$
(2.3)

and

$$\begin{aligned} S(\lambda ) \varphi (x) = \int _{\Sigma } \frac{1}{2 \pi } K_0 \bigl ( - i \sqrt{\lambda } |x- y| \bigr ) \varphi (y) d \sigma (y), \quad \varphi \in L^2(\Sigma ), ~x \in \Sigma . \end{aligned}$$
(2.4)

It is known that \(SL(\lambda )\) and \(S(\lambda )\) are bounded and \(\text {ran}\, S(\lambda ) \subseteq H^{1}(\Sigma )\); cf. [25, Theorem 6.12 and Theorem 7.2]. In particular, \(S(\lambda )\) gives rise to a compact operator in \(H^s(\Sigma )\) for every \(s \in [0,1]\). Furthermore, \(S(\lambda )\) is self-adjoint and positive for \(\lambda < 0\) (see Step 1 in the proof of Proposition 2.2). Some properties of \(\Psi _{\lambda }\) and \(S(\lambda )\) that are important in the proof of Theorem 1.1 are summarized in the following two propositions; cf. Appendix A for the proof of Propositions 2.1 and 2.2.

Proposition 2.1

Let \(\lambda \in {\mathbb C}{\setminus } [0, \infty )\) and let \(\Psi _{\lambda }\) be given by (2.2). Then

$$\begin{aligned} \Psi _{\lambda } = -2 i \partial _z SL(\lambda ): L^2(\Sigma ) \rightarrow L^2({\mathbb R}^2) \end{aligned}$$
(2.5)

is bounded and the following is true:

  1. (i)

    \(\Psi _{\lambda }\) gives rise to a bijective mapping \(\Psi _{\lambda }: H^{1/2}(\Sigma ) \rightarrow \mathcal {H}_{\lambda }\), where

    $$\begin{aligned} \begin{aligned} \mathcal {H}_{\lambda }:= \big \{ f \in H^1(\mathbb {R}^2 {\setminus } \Sigma ) \, | \, \partial _{\overline{z}}f_{+} \oplus \partial _{\overline{z}} f_{-} \in H^1({\mathbb R}^2), \left( - \Delta - \lambda \right) f_{\pm } = 0\, \textrm{on }\, \Omega _{\pm } \big \}. \end{aligned} \end{aligned}$$
  2. (ii)

    \(\Psi _{\lambda }^{*}: L^2({\mathbb R}^2) \rightarrow L^2(\Sigma )\) is a compact operator, \(\Psi _\lambda ^* = -2i \gamma _D \partial _{\overline{z}} (-\Delta -\overline{\lambda })^{-1}\), and \(\mathrm{ran\,}\Psi _{\lambda }^{*} \subseteq H^{1/2}(\Sigma )\).

  3. (iii)

    For all \(\varphi \in H^{1/2}(\Sigma )\) the jump relations

    $$\begin{aligned} \begin{aligned} i ( \nu _1 + i \nu _2) \big ( \gamma _D^{+} ( \Psi _{\lambda } \varphi )_{+} - \gamma _D^{-} ( \Psi _{\lambda } \varphi )_{-} \big )&= \varphi ,\\ -i \big ( \gamma _D^{+} \partial _{\overline{z}} ( \Psi _{\lambda } \varphi )_{+} + \gamma _D^{-} \partial _{\overline{z}} ( \Psi _{\lambda } \varphi )_{-} \big )&= \lambda S(\lambda ) \varphi , \end{aligned} \end{aligned}$$

    hold.

For \(\lambda < 0\) denote by \(\mu _n(S(\lambda ))\) the discrete eigenvalues of the positive self-adjoint operator \(S(\lambda )\) ordered non-increasingly and with multiplicities taken into account.

Proposition 2.2

Let \(S(\lambda )\) be defined by (2.4) and let \(n \in \mathbb {N}\) be fixed. Then the following holds:

  1. (i)

    The function \((- \infty , 0) \ni \lambda \mapsto \lambda \mu _n( S(\lambda ) )\) is continuous, strictly monotonically increasing and

    $$\begin{aligned} \lim _{\lambda \rightarrow 0^{-} }\lambda \mu _n( S(\lambda ) ) = 0 \quad \text {and} \quad \lim _{\lambda \rightarrow -\infty } \lambda \mu _n( S(\lambda ) ) = - \infty . \end{aligned}$$
  2. (ii)

    For \(a<0\) the unique solution \(\lambda _n(a) \in (- \infty , 0)\) of \(\lambda \mu _n(S(\lambda )) = a\) (see (i)) admits the asymptotic expansion \(\lambda _n(a) = - 4 a^2 + \mathcal {O}(1)\) for \(a \rightarrow - \infty \), where the dependence on n appears in the \(\mathcal {O}(1)\)-term.

Proof of Theorem 1.1

Step 1. We verify that \(T_{\alpha }\) is symmetric in \(L^2({\mathbb R}^2)\). Observe first that for \(f \in \mathrm{dom\,}T_\alpha \) we have \(\partial _{\overline{z}} f_{\pm } \in H^1(\Omega _{\pm })\) and \(\Delta f_\pm = 4 \partial _{z} \partial _{\overline{z}} f_{\pm } \in L^2(\Omega _{\pm })\), and hence \(T_\alpha \) is well-defined. Moreover, as \(C_0^{\infty }(\mathbb {R}^2 {{\setminus }} \Sigma ) \subseteq \mathrm{dom\,}T_{\alpha }\) it is also clear that \(\text {dom}\,T_\alpha \) is dense. In order to show that \(T_{\alpha }\) is symmetric, we note that integration by parts in \(\Omega _{\pm }\) yields for \(f,g \in \mathrm{dom\,}T_{\alpha }\)

$$\begin{aligned} \begin{aligned}&( - \Delta f_{\pm } , g_{\pm } )_{L^2(\Omega _{\pm })} = ( -4 \partial _{z} \partial _{\overline{z}} f_{\pm } , g_{\pm } )_{L^2(\Omega _{\pm })} \\ {}&\qquad = 4 ( \partial _{\overline{z}} f_{\pm } , \partial _{\overline{z}} g_{\pm } )_{L^2(\Omega _{\pm })} \mp 2 \big ( (\nu _1 - i \nu _2) \gamma _D^{\pm } ( \partial _{\overline{z}} f_{\pm } ), \gamma _D^{\pm } g_{\pm } \big )_{L^2(\Sigma )} \\ {}&\qquad = 4 ( \partial _{\overline{z}} f_{\pm } , \partial _{\overline{z}} g_{\pm } )_{L^2(\Omega _{\pm })} \mp 2 \big ( \gamma _D^{\pm } ( \partial _{\overline{z}} f_{\pm }), (\nu _1 + i \nu _2) \gamma _D^{\pm } g_{\pm }) \big )_{L^2(\Sigma )}. \end{aligned} \end{aligned}$$
(2.6)

Now, consider (2.6) for \(f=g\) and add the equations for \(\Omega _{+}\) and \(\Omega _{-}\). Then, using \(\gamma _D^{+} ( \partial _{\overline{z}} f_{+})=\gamma _D^{-} ( \partial _{\overline{z}} f_{-})\) and the transmission condition for \(f \in \mathrm{dom\,}T_{\alpha }\), one finds that

$$\begin{aligned} \begin{aligned} \big ( T_{\alpha } f , f \big )_{L^2({\mathbb R}^2)}&= 4 \big ( \Vert \partial _{\overline{z}} f_{+} \Vert ^2_{L^2(\Omega _{+})} + \Vert \partial _{\overline{z}} f_{-} \Vert ^2_{L^2(\Omega _{-})} \big ) \\ {}&\qquad - \big ( \gamma _D^{+} ( \partial _{\overline{z}} f_{+}) + \gamma _D^{-} ( \partial _{\overline{z}} f_{-} ) , (\nu _1 + i \nu _2) ( \gamma _D^{+} f_{+} - \gamma _D^{-} f_{-} ) \big )_{L^2(\Sigma )} \\ {}&= 4 \Vert \partial _{\overline{z}} f_{+} \oplus \partial _{\overline{z}} f_{-} \Vert ^2_{L^2(\mathbb {R}^2)} + \alpha \Vert \gamma _D^{+} ( \partial _{\overline{z}} f_{+}) + \gamma _D^{-} ( \partial _{\overline{z}} f_{-}) \Vert ^2_{L^2(\Sigma )} \in \mathbb {R}. \end{aligned} \end{aligned}$$
(2.7)

Since this holds for all \(f \in \mathrm{dom\,}T_{\alpha }\), we conclude that \(T_{\alpha }\) is symmetric.

Step 2. Proof of the Birman-Schwinger principle in (iii): To show the first implication, assume that \(\lambda \in {\mathbb C}{\setminus } [0, \infty )\) with \(1 \in \sigma _p ( \alpha \lambda S(\lambda ) )\) and choose \(\varphi \in {\mathrm{ker\,}\,}( I - \alpha \lambda S(\lambda )) {\setminus } \{0\}\). Then it follows from the mapping properties of \(S(\lambda )\) that \(\varphi = \alpha \lambda S(\lambda ) \varphi \in H^{1/2}(\Sigma )\) holds. Therefore, Proposition 2.1 (i) implies that \(f:= \Psi _{\lambda } \varphi \in \mathcal {H}_\lambda \) fulfils \(f \ne 0\), \(f \in H^1(\mathbb {R}^2 {\setminus } \Sigma )\), \(\partial _{\overline{z}}f_{+} \oplus \partial _{\overline{z}} f_{-} \in H^1({\mathbb R}^2)\) and, as \(\varphi \in {\mathrm{ker\,}\,}( 1 - \alpha \lambda S(\lambda )) {\setminus } \{0\}\), Proposition 2.1 (iii) implies

$$\begin{aligned} i (\nu _1 + i \nu _2) \big ( \gamma _D^{+} f_{+} - \gamma _D^{-} f_{-} \big ) = \varphi = \alpha \lambda S(\lambda ) \varphi = -i \alpha \big ( \gamma _D^{+} ( \partial _{\overline{z}} f_{+}) + \gamma _D^{-} ( \partial _{\overline{z}} f_{-} ) \big ). \end{aligned}$$

Hence, \(f \in \mathrm{dom\,}T_\alpha \). Moreover, as \(f \in \mathcal {H}_\lambda \), we conclude \(f \in {\mathrm{ker\,}\,}( T_{\alpha } - \lambda ) {\setminus } \{0\}\) and hence \(\lambda \in \sigma _p (T_{\alpha })\).

To show the second implication, we assume that \(\lambda \in \sigma _p (T_{\alpha })\) is given and we choose \(f \in {\mathrm{ker\,}\,}(T_{\alpha } - \lambda ) {\setminus } \{0\}\). Then, by Proposition 2.1 (i) there exists a unique \(\varphi \in H^{1/2}(\Sigma )\) such that \(f = \Psi _{\lambda } \varphi \). Moreover, using \(f \in \mathrm{dom\,}T_\alpha \) and Proposition 2.1 (iii) one finds that

$$\begin{aligned} 0 = i (\nu _1 + i \nu _2) \big ( \gamma _D^{+} f_{+} - \gamma _D^{-} f_{-} \big ) + i \alpha \big ( \gamma _D^{+} ( \partial _{\overline{z}} f_{+}) + \gamma _D^{-} ( \partial _{\overline{z}} f_{-} ) \big ) = ( I - \alpha \lambda S(\lambda ))\varphi . \end{aligned}$$

Since \(\varphi \ne 0\), we conclude \(1 \in \sigma _p ( \alpha \lambda S(\lambda ))\).

Step 3. Next, we prove that \(T_{\alpha }\) is a self-adjoint operator and the resolvent formula in (iv). Let \(\lambda \in {\mathbb C}{\setminus } ( [0, \infty ) \cup \sigma _p (T_{\alpha }) )\) be fixed. First, we show that \(I - \alpha \lambda S(\lambda )\) gives rise to a bijective map in \(H^s(\Sigma )\) for every \(s \in [0,1]\). Recall that \(S(\lambda )\) is compact in \(H^s(\Sigma )\). Since \(I - \alpha \lambda S(\lambda )\) is injective for our choice of \(\lambda \) by the Birman-Schwinger principle in (iii), Fredholm’s alternative shows that \(I - \alpha \lambda S(\lambda )\) is indeed bijective.

Recall that \(T_\alpha \) is symmetric; cf. Step 1. Hence, to show that \(T_\alpha \) is self-adjoint, it suffices to verify that \(\text {ran}(T_\alpha -\lambda ) = L^2(\mathbb {R}^2)\) holds for \(\lambda \in {\mathbb C}{\setminus } ( [0, \infty ) \cup \sigma _p (T_{\alpha }) )\). Fix such a \(\lambda \), let \(f \in L^2(\mathbb {R}^2)\), and define

$$\begin{aligned} g = (- \Delta - \lambda )^{-1}f + \alpha \Psi _{\lambda } (I - \alpha \lambda S(\lambda ))^{-1} \Psi _{\overline{\lambda }}^{*} f, \end{aligned}$$
(2.8)

which is well-defined by the considerations above. Since \(\Psi _{\overline{\lambda }}^{*} f \in H^{1/2}(\Sigma )\) by Proposition 2.1 (ii) and \((I - \alpha \lambda S(\lambda ))^{-1}\) is bijective in \(H^{1/2}(\Sigma )\), we conclude with Proposition 2.1 (i) that \(\Psi _{\lambda } (I - \alpha \lambda S(\lambda ))^{-1} \Psi _{\overline{\lambda }}^{*} f \in \mathcal {H}_\lambda \subseteq H^1(\mathbb {R}^2 {\setminus } \Sigma )\). In particular, with \((- \Delta - \lambda )^{-1}f \in H^2(\mathbb {R}^2)\) this implies that \(g \in H^1(\mathbb {R}^2 {\setminus } \Sigma )\) and \(\partial _{\overline{z}}g_{+} \oplus \partial _{\overline{z}} g_{-} \in H^1({\mathbb R}^2)\). Moreover, with Proposition 2.1(ii)–(iii) we obtain that

$$\begin{aligned} \begin{aligned}&i (\nu _1 + i \nu _2) \big ( \gamma _D^{+} g_{+} - \gamma _D^{-} g_{-} \big ) + i \alpha \big ( \gamma _D^{+} ( \partial _{\overline{z}}g_{+}) + \gamma _D^{-}(\partial _{\overline{z}} g_{-}) \big ) \\ {}&\qquad = \alpha (I - \alpha \lambda S(\lambda ) )^{-1} \Psi _{\overline{\lambda }}^{*} f - \alpha \Psi _{\overline{\lambda }}^{*} f - \alpha ^2 \lambda S(\lambda ) (I - \alpha \lambda S(\lambda ) )^{-1} \Psi _{\overline{\lambda }}^{*} f \\ {}&\qquad = \alpha (I - \alpha \lambda S(\lambda )) (I - \alpha \lambda S(\lambda ) )^{-1} \Psi _{\overline{\lambda }}^{*} f - \alpha \Psi _{\overline{\lambda }}^{*} f = 0 \end{aligned} \end{aligned}$$

and hence, \(g \in \mathrm{dom\,}T_{\alpha }\). As \(\Psi _{\lambda } (I - \alpha \lambda S(\lambda ))^{-1} \Psi _{\overline{\lambda }}^{*} f \in \mathcal {H}_\lambda \) by Proposition 2.1 (i), we conclude

$$\begin{aligned} \begin{aligned} \left( - \Delta - \lambda \right) g_{\pm }&= \left( -\Delta - \lambda \right) \big ( (- \Delta - \lambda )^{-1} f \big )_{\pm } + \alpha \left( - \Delta - \lambda \right) \big ( \Psi _{\lambda } (I - \alpha \lambda S(\lambda ) )^{-1} \Psi _{\overline{\lambda }}^{*} f \big )_{\pm }\\&= \left( - \Delta - \lambda \right) \big ( (- \Delta - \lambda )^{-1} f \big )_{\pm } = f_{\pm }, \end{aligned} \end{aligned}$$

i.e. \((T_{\alpha } - \lambda ) g = f\). Since \(f \in L^2({\mathbb R}^2)\) was arbitrary, we conclude that \(\mathrm{ran\,}(T_{\alpha } - \lambda ) = L^2({\mathbb R}^2)\) and that \(T_{\alpha }\) is self-adjoint. Moreover, the resolvent formula in item (iv) follows from (2.8).

Step 4. Next, we show \(\sigma _{\text {ess}}(T_{\alpha }) = [0, \infty )\). For this fix some \(\lambda \in {\mathbb C}{\setminus } \mathbb {R}\). Since \(\Psi _{\overline{\lambda }}^{*}: L^2({\mathbb R}^2) \rightarrow L^2(\Sigma )\) is compact by Proposition 2.1 (ii), the resolvent formula in (iv) implies that \((T_{\alpha } - \lambda )^{-1} - (-\Delta - \lambda )^{-1}\) is a compact operator in \(L^2({\mathbb R}^2)\). Consequently, Weyl’s Theorem [27, Theorem XIII.14] yields that \(\sigma _\text {ess}(T_\alpha ) = \sigma _\text {ess}(-\Delta )=[0,\infty )\).

Step 5. Proof of (i): Let \(\alpha \ge 0\). Then, (2.7) implies that \(T_\alpha \) is non-negative and hence, \(\sigma (T_\alpha ) \subset [0,\infty )\). Since the latter set coincides with \(\sigma _\text {ess}(T_\alpha )\), see Step 4, we conclude \(\sigma _\text {disc}(T_\alpha )=\emptyset \).

Step 6. Proof of (ii): Let \(\alpha <0\). Since \(\sigma _{\text {ess}}(T_{\alpha }) = [0, \infty )\), it follows from the Birman-Schwinger principle in (iii) that

$$\begin{aligned} \sigma _{\text {disc}}(T_{\alpha }) = \left\{ \lambda _n \, | \, n \in {\mathbb N}\right\} = \left\{ \lambda < 0 \, | \, \exists n \in {\mathbb N}\text { such that } \lambda \mu _n(S(\lambda )) = \alpha ^{-1} \right\} \end{aligned}$$

holds. Note that by Proposition 2.2 the equation \(\lambda \mu _n(S(\lambda )) = \alpha ^{-1}\) has a unique solution \(\lambda _n\) for all \(n \in \mathbb {N}\). Moreover, for any \(n \in {\mathbb N}\) there cannot be infinitely many \(k \ne n\) with \(\lambda _n = \lambda _k\), since otherwise \(\alpha ^{-1} < 0\) would be an eigenvalue with infinite multiplicity of the self-adjoint and compact operator \(\lambda _n S(\lambda _n)\). Thus \(\sigma _{\text {disc}}(T_{\alpha })\) is indeed an infinite set. Furthermore, as \(S(\lambda )\) is a positive self-adjoint operator in \(L^2(\Sigma )\); cf. Step 1 in the proof of Proposition 2.2, we have by definition \(\mu _n(S(\lambda )) \ge \mu _{n+1}(S(\lambda ))\) implying \(\lambda \mu _n ( S(\lambda )) \le \lambda \mu _{n+1}(S(\lambda ))\). Therefore, the monotonicity of the map \(\lambda \mapsto \lambda \mu _n (S(\lambda ))\) from Proposition 2.2 yields \(\lambda _{n+1} \le \lambda _{n}\) for all \(n \in {\mathbb N}\). This shows that 0 cannot be an accumulation point of the sequence \((\lambda _n)_{n \in {\mathbb N}}\) and as \(\sigma _{\text {ess}}(T_{\alpha })\cap (-\infty ,0)=\emptyset \) the sequence \((\lambda _n)_{n \in {\mathbb N}}\) has no finite accumulation points, that is, \(\sigma _{\text {disc}}(T_{\alpha })\) must be unbounded from below.

It remains to prove the asymptotic expansion in item (ii). By the above considerations \(\lambda _n\) is determined as the unique solution of \(\lambda \mu _n(S(\lambda )) = \alpha ^{-1}\). Clearly, if \(\alpha \rightarrow 0^-\), then \(a:= \alpha ^{-1} \rightarrow -\infty \). Hence, it follows from Proposition 2.2 (ii) with \(a = \alpha ^{-1}\) that \(\lambda _n = - \frac{4}{\alpha ^2} + \mathcal {O}(1)\) for \(\alpha \rightarrow 0^-\) and that the dependence on n appears in the \(\mathcal {O}(1)\)-term. \(\square \)

3 Proof of Theorem 1.2

In this section we show that \(T_\alpha \) is the non-relativistic limit of a family of Dirac operators with electrostatic and Lorentz scalar \(\delta \)-shell potentials formally given by (1.8), whose interaction strengths are suitably scaled. First, we recall the rigorous definition of the operator \(A_{\eta , \tau }\) associated with (1.8), see [5,6,7] for details. Let

$$\begin{aligned} \begin{aligned} \sigma _1 = \left( \begin{array}{cc} 0 &{}{} 1\\ 1 &{}{} 0 \\ \end{array}\right) , \sigma _2 = \left( \begin{array}{cc} 0 &{}{} -i\\ i &{}{} 0 \\ \end{array}\right) , \text{ and } \sigma _3 = \left( \begin{array}{cc} 1 &{}{} 0\\ 0 &{}{} -1 \\ \end{array}\right) \end{aligned} \end{aligned}$$
(3.1)

be the Pauli spin matrices and denote the \(2 \times 2\) identity matrix by \(I_2\). Furthermore, for \(x = (x_1,x_2) \in {\mathbb R}^2\) we will use the abbreviations

$$\begin{aligned} \sigma \cdot x = \sigma _1 x_1 + \sigma _2 x_2 \quad \text {and} \quad \sigma \cdot \nabla = \sigma _1 \partial _1 + \sigma _2 \partial _2. \end{aligned}$$
(3.2)

We define Dirac operators with electrostatic and Lorentz scalar \(\delta \)-shell interactions of strengths \(\eta , \tau \in {\mathbb R}\) in \(L^2({\mathbb R}^2)^2\) by

$$\begin{aligned} \begin{aligned} A_{\eta , \tau } f&= \left( - i c (\sigma \cdot \nabla ) +\frac{c^2}{2} \sigma _3 \right) f_{+} \oplus \left( - i c (\sigma \cdot \nabla ) + \frac{c^2}{2} \sigma _3 \right) f_{-}, \\ \mathrm {dom\,}A_{\eta , \tau }&= \bigg \{ f \in H^1(\Omega _{+})^2 \oplus H^1(\Omega _{-})^2 \, \big | \\ {}&~~ i c (\sigma \cdot \nu ) \left( \gamma _D^{+}f_{+} - \gamma _D^{-}f_{-} \right) + \frac{1}{2} ( \eta I_2+ \tau \sigma _3 ) \left( \gamma _D^{+}f_{+} + \gamma _D^{-}f_{-} \right) = 0 \bigg \}. \end{aligned} \end{aligned}$$
(3.3)

It is shown in [6, 7] that \(A_{\eta , \tau }\) is self-adjoint in \(L^2({\mathbb R}^2)^2\), whenever \(\eta ^2-\tau ^2 \ne 4 c^2\), and as in [5] one sees that these operators are the self-adjoint realisations of the formal differential expression (1.8). In the above definition we are using units such that \(\hbar = 1\) and consider the mass \(m=\frac{1}{2}\), but we keep the speed of light c as a parameter for the discussion of the non-relativistic limit \(c \rightarrow \infty \).

Throughout this section we make use of the self-adjoint free Dirac operator \(A_0\), which coincides with the operator\(A_{0,0}\) given in (3.3) and which is defined on \(H^1(\mathbb {R}^2)^2\). For \(\lambda \in \rho (A_0)=\mathbb {C} {\setminus }\big ((-\infty , -\frac{c^2}{2}] \cup [\frac{c^2}{2}, \infty )\big )\) the integral kernel of the resolvent of \(A_0\) is given by \(G_\lambda (x-y)\), where \(G_\lambda (x)\) is defined for \(x \in {\mathbb R}^2 {\setminus } \{0\}\) by

$$\begin{aligned} G_{\lambda }(x)= & {} \frac{1}{2 \pi c} \sqrt{ \frac{\lambda ^2}{c^2} - \frac{c^2}{4}} K_1 \left( -i \sqrt{ \frac{\lambda ^2}{c^2} - \frac{c^2}{4}} |x| \right) \frac{1}{|x|} (\sigma \cdot x) \nonumber \\{} & {} + \frac{1}{2 \pi c} K_0 \left( -i \sqrt{ \frac{\lambda ^2}{c^2} - \frac{c^2}{4} } |x| \right) \left( \frac{\lambda }{c} I_2 + \frac{c}{2} \sigma _3 \right) ; \end{aligned}$$
(3.4)

cf. [6, equation (3.2)]. With this function we define the two families of integral operators

$$\begin{aligned} \begin{aligned} \Phi _{\lambda } \varphi (x)&= \int _{\Sigma } G_{\lambda }(x-y) \varphi (y) \textrm{d} \sigma (y), \quad \varphi \in L^2(\Sigma )^2, ~x \in {\mathbb R}^2{\setminus }\Sigma , \\ \mathcal {C}_{\lambda } \varphi (x)&= \lim _{\varepsilon \rightarrow 0^{+}} \int \limits _{\Sigma {\setminus } B(x,\varepsilon )} G_{\lambda }(x-y) \varphi (y) \textrm{d} \sigma (y), \quad \varphi \in L^2(\Sigma )^2, ~ x \in \Sigma , \end{aligned} \end{aligned}$$
(3.5)

where \(B(x,\varepsilon )\) is the ball of radius \(\varepsilon \) centered at x. Both operators \(\Phi _{\lambda }: L^2(\Sigma )^2 \rightarrow L^2({\mathbb R}^2)^2\) and \(\mathcal {C}_{\lambda }: L^2(\Sigma )^2 \rightarrow L^2(\Sigma )^2\) are well-defined and bounded; cf. [6, Proposition 3.3 and equation (3.7)].

In the following lemma, which is a preparation for the proof of Theorem 1.2, we will use the matrices

$$\begin{aligned} M_1 = \left( \begin{array}{cc} 1 &{} 0 \\ 0 &{} 0 \end{array} \right) , \quad M_2 = \left( \begin{array}{cc} 0 &{} 1 \\ 0 &{} 0 \end{array} \right) , \quad \text {and} \quad M_3 = \left( \begin{array}{cc} 0 &{} 0 \\ 0 &{} 1 \end{array} \right) ; \end{aligned}$$

products of scalar operators and matrices are understood componentwise, e.g.

$$\begin{aligned} (-\Delta - \lambda )^{-1} M_1 = \left( \begin{array}{cc} (-\Delta - \lambda )^{-1} &{} 0 \\ 0 &{} 0 \end{array} \right) : L^2({\mathbb R}^2)^2 \rightarrow L^2({\mathbb R}^2)^2. \end{aligned}$$

Lemma 3.1

Let \(\lambda \in {\mathbb C}{\setminus } {\mathbb R}\). Then there exists a constant \(K>0\), depending only on \(\lambda \) and \(\Sigma \), such that the estimates

$$\begin{aligned}&\displaystyle \begin{aligned}&\displaystyle \big \Vert \big ( A_0 - (\lambda + c^2 / 2 ) \big ) ^{-1} - \left( - \Delta - \lambda \right) ^{-1} M_1 \big \Vert \le \frac{K}{c},&\end{aligned} \end{aligned}$$
(3.6a)
$$\begin{aligned}&\displaystyle \big \Vert c \Phi _{\lambda + c^2 / 2} M_3 - \Psi _{\lambda } M_2 \big \Vert \le \frac{K}{c},&\end{aligned}$$
(3.6b)
$$\begin{aligned}&\displaystyle \big \Vert c M_3 \Phi _{\lambda +c^2 / 2}^{*} - M_2^{\top } \Psi _{\lambda }^{*} \big \Vert \le \frac{K}{c},&\end{aligned}$$
(3.6c)
$$\begin{aligned}&\displaystyle \big \Vert c^2 M_3 \mathcal {C}_{\lambda + c^2 / 2} M_3 - \lambda S(\lambda ) M_3 \big \Vert \le \frac{K}{c},&\end{aligned}$$
(3.6d)

are valid for all sufficiently large \(c>0\).

Proof

We use a similar strategy as in the proof of [4, Proposition 5.2]. In the following let \(\lambda \in {\mathbb C}{\setminus } {\mathbb R}\) be fixed. Then \(\lambda + \frac{c^2}{2} \in {\mathbb C}{\setminus } {\mathbb R}\) and hence all operators in (3.6a)–(3.6d) are well-defined. One verifies by direct calculation that for sufficiently large \(c>0\) and all \(t \in [0,1]\)

$$\begin{aligned} 0 < \frac{1}{2} \left| \sqrt{\lambda } \right| \le \left| \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} \right| \le \frac{3}{2} \left| \sqrt{ \lambda } \right| \quad \text {and} \quad \frac{1}{2} \text {Im} \sqrt{\lambda } \le \text {Im} \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} \end{aligned}$$
(3.7)

hold. With the well-known asymptotic expansions of the modified Bessel functions and \(K_1'(z) = -K_0(z) - \frac{1}{z} K_1(z)\), (see [1]) one shows that there exist constants \(\widehat{K}, \kappa , R > 0\), depending only on \(\lambda \), such that

$$\begin{aligned} \left| K_j \left( - i \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} |x| \right) \right| \le \widehat{K} \left\{ \begin{array}{ll} |x|^{-1}, &{} \text {for } |x| < R, \\ e^{- \kappa |x|}, &{} \text {for } |x| \ge R, \end{array}\right. \end{aligned}$$
(3.8)

and

$$\begin{aligned} \left| K_1' \left( - i \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} |x| \right) \right| \le \widehat{K} \left\{ \begin{array}{ll} |x|^{-2}, &{} \text {for } |x| < R, \\ e^{- \kappa |x|}, &{} \text {for } |x| \ge R, \end{array}\right. \end{aligned}$$
(3.9)

hold for all \(x \in {\mathbb R}^2 {\setminus } \{0\}\), \(j \in \{0,1\}\), \(t \in [0,1]\), and sufficiently large \(c>0\).

Next, with \(G_{\lambda +c^2/2}\) defined by (3.4) we find

$$\begin{aligned} G_{\lambda + c^2 / 2 }(x)= & {} \frac{1}{2\pi c} \sqrt{\lambda +\frac{\lambda ^2}{c^2}} K_1 \left( - i \sqrt{\lambda +\frac{\lambda ^2}{c^2}} |x| \right) \frac{1}{|x|} ( \sigma \cdot x) \nonumber \\{} & {} + \frac{1}{2 \pi c} K_0 \left( -i \sqrt{\lambda +\frac{\lambda ^2}{c^2}} |x| \right) \left( \frac{\lambda }{c} I_2 + c M_1 \right) . \end{aligned}$$
(3.10)

Let

$$\begin{aligned} U_\lambda (x) = \frac{1}{2 \pi } K_0 \bigl ( - i \sqrt{ \lambda } |x| \bigr ), \quad x \in {\mathbb R}^2 {\setminus } \{0\}, \end{aligned}$$

be the integral kernel of the resolvent of the free Laplace operator; cf. [28, Chapter 7.5]. Then

$$\begin{aligned} G_{\lambda + c^2 / 2}(x) - U_\lambda (x) M_1 = t_1(x) + t_2(x) + t_3(x) \end{aligned}$$

holds, where the matrix-valued functions \(t_1, t_2\), and \(t_3\) are given by

$$\begin{aligned} \begin{aligned} t_1(x)&= \frac{1}{2 \pi c} \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} K_1 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) \frac{ \sigma \cdot x}{|x|}, \\ t_2(x)&= \frac{1}{2 \pi } \left( K_0 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) - K_0 \bigl ( - i \sqrt{ \lambda } |x| \bigr ) \right) M_1, \\ t_3(x)&= \frac{\lambda }{2 \pi c^2} K_0 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) I_2. \end{aligned} \end{aligned}$$

With (3.7) and (3.8) applied with \(t=1\) one finds that there exist constants \(k_1, \kappa , R > 0\), depending only on \(\lambda \), such that for \(j \in \{1,3\}\) and sufficiently large \(c>0\) one has

$$\begin{aligned} \left| t_j(x) \right| \le \frac{k_1}{c} \left\{ \begin{array}{ll} |x|^{-1}, &{} \text {for } |x| < R, \\ e^{- \kappa |x|}, &{} \text {for } |x| \ge R. \end{array}\right. \end{aligned}$$

To estimate \(t_2\), we use \(K_0' = -K_1\) and obtain with the fundamental theorem of calculus, (3.7), and (3.8)

$$\begin{aligned} \begin{aligned}&\left| K_0 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) - K_0 \left( - i \sqrt{ \lambda } |x| \right) \right| \le \int _0^1 \left| \frac{d}{d t} K_0 \left( -i \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} |x| \right) \right| d t \\ {}&\quad = \int _0^1 \frac{|\lambda |^2 |x|}{\left| \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} \right| } \frac{1}{2 c^2} \left| K_1 \left( -i \sqrt{ \lambda + t \frac{\lambda ^2}{c^2}} |x| \right) \right| d t \\ {}&\quad \le \frac{k_2}{c^2} \left\{ \begin{array}{ll} 1, &{}{} \text{ for } |x| < R, \\ e^{- \frac{\kappa }{2} |x|}, &{}{} \text{ for } |x| \ge R, \end{array}\right. \end{aligned} \end{aligned}$$
(3.11)

with a constant \(k_2\) which depends only on \(\lambda \). Thus, if we define \(k_3 = 2 k_1 + \frac{k_2 R}{2 \pi }\), then

$$\begin{aligned} \left| G_{\lambda + c^2 / 2}(x) - U_\lambda (x) M_1 \right| \le \frac{k_3}{c} \left\{ \begin{array}{ll} |x|^{-1}, &{} \text {for } |x| < R, \\ e^{- \frac{\kappa }{2} |x|}, &{} \text {for } |x| \ge R. \end{array}\right. \end{aligned}$$

This estimation for the integral kernel yields with the Schur test; cf. [4, Proposition A.3] for a similar argument,

$$\begin{aligned} \begin{aligned} \big \Vert \big ( A_0 - (\lambda + c^2 / 2) \big ) ^{-1} - \left( -\Delta - \lambda \right) ^{-1} M_1 \big \Vert \le \frac{K}{c} \end{aligned} \end{aligned}$$

for all sufficiently large \(c>0\), which is the first claimed estimate (3.6a).

Next, we prove (3.6b). Recall that the integral kernel \(L_{\lambda }\) of \(\Psi _\lambda \) is given by (2.1). Using that \(\sigma _1 M_3 = M_2\), \(\sigma _2 M_3 = -i M_2\), and \(M_1 M_3 = 0\), we obtain with (3.10) the decomposition

$$\begin{aligned} c G_{\lambda + c^2 / 2}(x) M_3 - L_{\lambda } (x) M_2 = \tau _1(x) + \tau _2(x) + \tau _3(x) \end{aligned}$$

with

$$\begin{aligned} \begin{aligned} \tau _1(x)&= \frac{1}{2 \pi } \left( \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} - \sqrt{\lambda } \right) K_1 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) \frac{x_1 - i x_2}{|x|} M_2, \\ \tau _2(x)&= \frac{\sqrt{ \lambda }}{2 \pi } \left( K_1 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) - K_1 \bigl ( -i \sqrt{ \lambda } |x| \bigr ) \right) \frac{x_1 - i x_2}{|x|} M_2, \\ \tau _3(x)&= \frac{\lambda }{2 \pi c} K_0 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) M_3. \end{aligned} \end{aligned}$$

Similar as above it can be shown that there exists a \(k_4>0\), depending only on \(\lambda \), such that for all \(j \in \{1,2,3\}\)

$$\begin{aligned} \left| \tau _j(x) \right| \le \frac{k_4}{c} \left\{ \begin{array}{ll} |x|^{-1}, &{} \text {for } |x| < R, \\ e^{- \frac{\kappa }{2} |x|}, &{} \text {for } |x| \ge R; \end{array}\right. \end{aligned}$$

to see the estimate for \(\tau _2\) one has to use (3.9). With the help of the Schur test the estimate (3.6b) follows (see also [4, Proposition A.4] for a similar argument); the constant \(k_4\) depends in this case on \(\lambda \) and \(\Sigma \). The estimate in (3.6c) follows by taking adjoints.

It remains to prove (3.6d). Taking \(M_3 (\sigma \cdot x) M_3 = 0 \), which holds for any \(x \in \mathbb {R}^2\), and (3.11) into account we obtain that

$$\begin{aligned} \begin{aligned}&\big | c^2 M_3 G_{\lambda + c^2 / 2}(x) M_3 - \lambda U_\lambda (x) M_3 \big | \\&\quad = \frac{|\lambda |}{2 \pi } \left| K_0 \left( -i \sqrt{ \lambda + \frac{\lambda ^2}{c^2}} |x| \right) - K_0 \bigl ( -i \sqrt{ \lambda } |x| \bigr ) \right| \le \frac{k_5}{c^2} \end{aligned} \end{aligned}$$

holds for all \(x \in {\mathbb R}^2 {\setminus } \{0\}\). Using the dominated convergence theorem, one sees that

$$\begin{aligned} \big ( c^2 M_3 \mathcal {C}_{\lambda + c^2 / 2} M_3 f \big )(x) = \int _{\Sigma } c^2 M_3 G_{\lambda + c^2 / 2}(x-y) M_3 f(y) d \sigma (y) \end{aligned}$$

holds for all \(f \in L^2(\Sigma )^2\) and \(x \in \Sigma \), i.e. the integral does not have to be understood as principal value. Thus we obtain with the Schur test [23, III. Example 2.4] that

$$\begin{aligned} \begin{aligned} \big \Vert c^2 M_3 \mathcal {C}_{\lambda + m c^2} M_3 - \lambda S(\lambda ) M_3 \big \Vert \le \frac{K}{c^2}. \end{aligned} \end{aligned}$$

In this case, the constant K depends on \(\lambda \) and \(\Sigma \). This yields (3.6d) and finishes the proof of this lemma. \(\square \)

Now we are prepared to prove Theorem 1.2 and show that \(A_{-\alpha c^2/2, \alpha c^2/2}\) converges in the non-relativistic limit to \(T_\alpha \) defined in (1.6).

Proof of Theorem 1.2

Let \(\lambda \in {\mathbb C}{\setminus } {\mathbb R}\) be fixed. Then it follows from [7, Lemma 5.4, Proposition 5.5, Theorem 5.6, and Lemma 5.9] (see also [6, Theorem 4.6]) that the operator \(I_2 - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2/2}: L^2(\Sigma )^2 \rightarrow L^2(\Sigma )^2\) is boundedly invertible and the resolvent of \(A_{- \alpha c^2/2, \alpha c^2/2} - c^2/2\) is given by

$$\begin{aligned} \begin{aligned}\big (A_{-\alpha c^2/2 , \alpha c^2/2} -&(\lambda + c^2/2) \big )^{-1} = \big ( A_{0}-(\lambda +c^2/2)\big ) ^{-1}\\ {}&+ \Phi _{\lambda + c^2/2} \left( I - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2/2} \right) ^{-1} \alpha c^2 M_3 \Phi _{\overline{\lambda } + c^2/2}^{*}. \end{aligned} \end{aligned}$$
(3.12)

Because of \(M_3 = M_3^2\) it follows from [26, Proposition 2.1.8] that

$$\begin{aligned} \sigma \big ( M_3 \mathcal {C}_{\lambda +c^2/2} \big ) \cup \{ 0 \} = \sigma \big ( M_3 \mathcal {C}_{\lambda + c^2 / 2} M_3 \big ) \cup \{ 0 \}. \end{aligned}$$

In particular, this yields that the operator \(I - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2/2} M_3\) is boundedly invertible in \(L^2(\Sigma )^2\) for all \(c>0\) and a direct calculation shows

$$\begin{aligned} (I - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2 / 2} )^{-1} M_3 = M_3 ( I - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2 / 2} M_3)^{-1}. \end{aligned}$$
(3.13)

Recall that for \(\lambda \in \mathbb {C} {\setminus } \mathbb {R}\) also \(I - \alpha \lambda S(\lambda )\) is boundedly invertible in \(L^2(\Sigma )\); cf. Theorem 1.1 (iv). Hence, we obtain from Lemma 3.1 and [23, IV. Theorem 1.16] that

$$\begin{aligned} \begin{aligned} \big \Vert ( I - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2 / 2} M_3)^{-1} - (I - \alpha \lambda S(\lambda ) M_3)^{-1} \big \Vert \le \frac{K}{c} \end{aligned} \end{aligned}$$
(3.14)

holds for all sufficiently large \(c>0\) with a constant \(K>0\) which depends only on \(\lambda \), \(\alpha \), and \(\Sigma \).

To conclude, note that (3.12) and (3.13) yield

$$\begin{aligned} \begin{aligned} \big (A_{-\alpha c^2/2 , \alpha c^2/2} -&(\lambda + c^2 / 2) \big )^{-1} = \big ( A_0 - (\lambda + c^2 / 2) \big )^{-1} \\&+ c \Phi _{\lambda + c^2 / 2} M_3 ( I - \alpha c^2 M_3 \mathcal {C}_{\lambda + c^2 / 2} M_3)^{-1} \alpha c M_3 \Phi _{\overline{\lambda }+ c^2/2} ^{*}, \end{aligned} \end{aligned}$$

while Theorem 1.1 (iv) and \(M_2 M_3 M_2^{\top } = M_1\) show

$$\begin{aligned} \begin{aligned} (T_{\alpha } - \lambda )^{-1} M_1&= ( - \Delta - \lambda )^{-1} M_1 + \Psi _{\lambda } ( I - \alpha \lambda S(\lambda ) )^{-1} \alpha \Psi _{\overline{\lambda }}^{*} M_1 \\&= ( - \Delta - \lambda )^{-1} M_1 + \Psi _{\lambda } M_2 (I - \alpha \lambda S(\lambda ) M_3)^{-1} \alpha M_2^{\top } \Psi _{\overline{\lambda }}^{*}. \end{aligned} \end{aligned}$$

Using Lemma 3.1 and (3.14) the last two displayed formulae finally lead to the claimed convergence result and it also follows that the order of convergence is \(\mathcal {O}(\frac{1}{c})\). \(\square \)