1 Introduction and Results

Since the pioneering discovery of E. Wigner on the universality of eigenvalue statistics of large random matrices [38], random matrix theory has become one of the most successful phenomenological theories to study disordered quantum systems, see [2] for a broad overview. Among many other applications, it has been used for open quantum systems and quantum transport, in particular to predict the distribution of transmission eigenvalues of scattering in quantum dots and wires. The theory has been developed over many excellent works starting with the ground-breaking papers by Mello et al. [30] and by Verbaarschot et al. [37]; for a complete overview with extensive references, see reviews by Beenakker [7, 8], Fyodorov and Savin [21] and Schomerus [35].

We will focus on quantum dots, i.e., systems without internal structure, coupled to an environment (electron reservoir) via scattering channels. Quantum wires, with a typically quasi one-dimensional internal structure, will be left for further works. In the simplest setup the quantum dot is described by a self-adjoint Hamiltonian (complex Hermitian matrixFootnote 1) \(H\in {\mathbb {C}}^{M\times M}\) acting on an M-dimensional state space \({\mathbb {C}}^M\). It is coupled to an environment with \(N_0\) effective degrees of freedom via an \(M\times N_0\) complex coupling matrix W. Following Wigner’s paradigm, both the Hamiltonian H and the coupling matrix W are drawn from random matrix ensembles respecting the basic symmetries of the model. Typically, the entries of W are independent, identically distributed (i.i.d.), while H is a Wigner matrix, i.e., it has i.i.d. entries on and above the diagonal. We allow for general distributions in contrast to most existing works in the literature that assume H has Gaussian or Lorentzian distribution.

The Hamiltonian of the total system at Fermi energy \(E\in {\mathbb {R}}\) is given by (see [7, Eq. (80)])

$$ {{\mathcal {H}}} = \sum _{a=1}^{N_0} | a \rangle E \langle a | + \sum _{\mu ,\nu =1}^M |\mu \rangle H_{\mu \nu } \langle \nu | + \sum _{\mu =1}^M\sum _{a=1}^{N_0} \Big [ |\mu \rangle W_{\mu a} \langle a| + |a\rangle W_{\mu a}^* \langle \mu | \Big ]. $$

One common assumption is that the interaction W is independent of the Fermi energy E.

At any fixed energy \(E\in {\mathbb {R}}\), we define the scattering matrix (see [7, Eq. (81)])

$$\begin{aligned} S(E) := I - 2\pi \mathrm {i}\,W^* (E\cdot I - H + \mathrm {i}\,\pi W W^*)^{-1} W \in {\mathbb {C}}^{N_0\times N_0}. \end{aligned}$$
(1.1)

This is the finite-dimensional analogue of the Mahaux–Weidenmüller formula in nuclear physics [28] that can be derived from \({{\mathcal {H}}}\) in the \(N_0\rightarrow \infty \) limit. The definition (1.1) will be the starting point of our mathematical analysis. Since \(H=H^*\), one can easily check that S(E) is unitary.

To distinguish between transmission and reflection, we assume that the scattering channels are split into two groups, left and right channels, with dimensions \(N_1+N_2=N_0\) and the interaction Hamiltonian is also split accordingly; \(W = (W_1 , W_2) \in {\mathbb {C}}^{M\times (N_1+N_2)}\). Therefore, S(E) has a natural \(2\times 2\) block structure and we can write it as (see [7, Eq. (23)])

$$\begin{aligned} S(E) = \left( \begin{array}{cc} R &{}\quad T' \\ T &{}\quad R' \end{array} \right) , \end{aligned}$$
(1.2)

where \(R\in {\mathbb {C}}^{N_1 \times N_1}\), \(R'\in {\mathbb {C}}^{N_2 \times N_2}\) are the reflection matrices and \(T\in {\mathbb {C}}^{N_2 \times N_1}\), \(T' \in {\mathbb {C}}^{N_1 \times N_2}\) are the transmission matrices.

As a consequence of unitarity, one finds that \(TT^*\), \(T'(T')^*\), \(I-RR^*\) and \(I -R' (R')^*\) have the same set of nonzero eigenvalues. For simplicity, we assume that \(N_1=N_2=N\), i.e., generically these four matrices have no zero eigenvalues, the general \(N_1 \ne N_2\) case has been considered in [11]. We denote these transmission eigenvalues by \(\lambda _1, \lambda _2, \ldots , \lambda _N\). They express the rate of the transmission through each channel. By unitarity of S, clearly \(\lambda _i\in [0,1]\) for all i; \(\lambda _i=0\) means the channel is closed, while \(\lambda _i=1\) corresponds to a fully open channel. The transmission eigenvalues carry important physical properties of the system. For example, \({\mathrm {Tr}}TT^* =\sum _i \lambda _i\) gives the zero temperature conductance (Landauer formula [7, Eq. (33)]), while

$$\begin{aligned} \sum _i \lambda _i (1-\lambda _i) = {\mathrm {Tr}}TT^* - {\mathrm {Tr}}(TT^*)^2 \end{aligned}$$
(1.3)

is the shot noise power, giving the zero temperature fluctuation of the current (Büttiker’s formula, [12, 7, Eq. (35)]). The dimensionless ratio of the shot noise power and the conductance is called the Fano factor (see [8, Eq. (2.15)])

$$\begin{aligned} F: =\frac{ \sum _i \lambda _i (1-\lambda _i)}{\sum _i \lambda _i}. \end{aligned}$$
(1.4)

The current fluctuation is therefore given by a certain linear statistics of the transmission eigenvalues, and thus, it can be computed from the density \(\rho \) of these eigenvalues. Therefore, determining \(\rho \) is a main task in the theory of quantum dots.

In many physical situations, it is found that \(\rho \) has a bimodal structure with a peak at zero and a peak at unit transmission rates. Furthermore, \(\rho \) exhibits a power law singularity at the edges of its support [0, 1]. One main result of the theory in [7] is that in the \(M\gg N\gg 1\) regime at energy \(E=0\) the density of transmission eigenvalues for a quantum dot is given by

$$\begin{aligned} \rho _{ \text{ Bee }}(\lambda ) = \frac{1}{\pi \sqrt{\lambda (1-\lambda )}}, \end{aligned}$$
(1.5)

(the answer is different for quantum wires), i.e., it has an inverse square root singularity at both edges, see [8, Eq. (3.12)]. In this case, the Fano factor is \(F=1/4\) which fits well the experimental data.

The goal of this paper is to revisit and substantially generalize the problem of transmission eigenvalues with very different methods than Beenakker and collaborators used. While those works used invariant matrix ensembles for H and relied on explicit computations for the circular ensemble, we consider very general distribution for the matrix elements of H. In particular, we show that in the regime \(\phi := N/M \in {(}0,1{]}\), \(M\rightarrow \infty \), the empirical density of transmission eigenvalues has a deterministic limit \(\rho =\rho _{\phi }\) and we give a simple algebraic equation to compute it. The solution can be continuously extended as \(\phi \rightarrow 0\), the equation simplifies for \(\phi =0\), the density becomes explicitly computable and it coincides with (1.5); \(\lim _{\phi \rightarrow 0} \rho _{\phi } =\rho _{\phi =0}= \rho _{ \text{ Bee }}\) for \(E=0\) and our formula holds for any E in the bulk spectrum of H. While no short explicit formula is available for the general case \(\phi \in (0, 1]\), we can analyze the singularities of \( \rho _{\phi }\) for any fixed \(\phi \in (0, 1]\) in detail.

More precisely, we rigorously prove that for any fixed \(\phi \in (0,1)\) the density \(\rho _{\phi }\) has an inverse square root singularity both at 0 and at 1,

$$\begin{aligned} \begin{aligned}&\rho _{\phi } (\lambda ) \sim \lambda ^{-1/2},&\;\quad \text { for } 0<\lambda \ll 1, \;\;\quad&\qquad \text { and } \\&\rho _{\phi } (\lambda ) \sim (1-\lambda )^{-1/2},&\text { for } 0<1-\lambda \ll 1,&\quad \end{aligned} \end{aligned}$$
(1.6)

qualitatively in line with \(\rho _{ \text{ Bee }}\) from (1.5). However, \(\rho _{\phi }\) is not symmetric around \(\lambda =\frac{1}{2}\) and the Fano factor at \(E=0\) slightly differs from \(\frac{1}{4}\) when \(\phi \ne 0\). Figure 1 shows the Fano factor for different values of \(\phi \) numerically computed from our theory. We mention that the deviation from 1/4 is within \(2\%\) for the entire range of \(\phi \in [0,1)\) which is well below the error bar of the experimental results presented in Fig. 6 of [7] and adapted from [31].

Fig. 1
figure 1

Fano factor for \(E=0\) and \(\phi \in (0,1)\)

The value \(\phi =1\) is special, since the singularity of \(\rho \) at \(\lambda \approx 0\) changes to

$$\begin{aligned} \rho _{\phi =1}(\lambda ) \sim \lambda ^{-2/3}, \;\quad \text{ for } 0<\lambda \ll 1 \end{aligned}$$
(1.7)

from \(\lambda ^{-1/2}\) in (1.6), while the inverse square root singularity at 1 persists. This enhancement of singularity signals the emergence of a \(\delta \)-function component at 0 in \(\rho _{\phi }\) as \(\phi \) becomes larger than 1, which is a direct consequence of \(TT^*\) not having full rank when \(N>M\). We remark that this regime is quite unphysical since it corresponds to more scattering channels than the total number of internal states of the quantum dot. In realistic quantum dots, N is smaller than M since not every mode of the dot may participate in scattering. We therefore do not pursue the detailed analysis of \(\rho _{\phi }\) for \(\phi >1\), although our method can easily be extended to these \(\phi \)’s as well.

There are two main differences between our model and that of Beenakker et al. First, the distribution imposed on the random matrix H is different and we consider random W. Second, our current method works in the entire regime \(\phi = N/M\in {(}0,1{]}\), while Beenakker assumes \(M\gg N\), i.e., \(\phi \ll 1\). We now explain both differences.

Following Wigner’s original vision, any relevant distribution must respect the basic symmetry of the model; in our case, this demands that H be complex Hermitian, while no symmetry constraint is imposed on W. Respecting basic symmetries, one may define ensembles essentially in two ways. Invariant ensembles are defined by imposing that the entire distribution be invariant; it is typically achieved via a global Gibbs factor times the natural flat measure on the space of matrices satisfying the basic symmetry. Wigner ensembles and their generalizations impose distributions directly on the matrix elements and often demand independence (up to the basic symmetry constraint). These two procedures typically yield different ensembles.

While in the simplest case of random Hermitian matrices both types of ensembles have been actively investigated, for ensembles with more complicated structure, like our S that is a rational function of the basic ingredients H and W (1.1), up to recently only the invariant approach was available. Sophisticated explicit formulas have been developed to find the joint distribution of eigenvalues for more and more complicated structured ensembles (see [18]), which could then be combined with orthogonal polynomial methods to obtain local correlation functions. The heavy reliance on explicit formulas, however, imposes a serious limitation on how complicated functions of random matrices, as well as how general distributions on these matrices can be considered. For example, the Gibbs factor is often restricted to Gaussian or closely related ensembles to remain in the realm of explicitly computable orthogonal polynomials.

There have been considerable developments in the other type of ensembles in the recent years. Departing from the invariant world, about 10 years ago the Wigner–Dyson–Mehta universality of local eigenvalue statistics has been proven for Wigner matrices with i.i.d. entries, see [16] and references therein. Later, the i.i.d. condition was relaxed and even matrices with quite general correlation structures among their entries can be handled [14]. One of the key ingredients was to better understand the Matrix Dyson equation (MDE), the basic equation governing the density of states [1]. Together with the linearization trick, this allows us to handle arbitrary polynomials in i.i.d. random matrices [13] and in the current work we extend our method to a large class of rational functions. Note that even if the building block matrices have independent entries, the linearization of their rational expressions will have dependence, but the general MDE can handle it [see (2.13)]. In our work, we deal only with bounded rational functions and relatively straightforward regularizations of unbounded rational functions, the general theory of unbounded rational functions is still in development, see [29] and references therein. We stress that the distribution of the matrix elements of H and W can be practically arbitrary. In particular, our result is not restricted to the Gaussian world.

In comparison, the result of Beenakker et al. [8, Sect. III.B], see also [10, Sect. IV], postulates that H is GUE, while W may even be deterministic and only its singular values are relevant. For example, if all nonzero singular values are equal one and \(N_0=2N\le M\), i.e., \(\phi \le \frac{1}{2}\), then S can be written as

$$ S = \frac{I + i\pi {{\widetilde{H}}}}{I- i\pi {{\widetilde{H}}}}, \qquad {{\widetilde{H}}}: = Q^T(H-E \cdot I )^{-1} Q, $$

where \(Q\in {\mathbb {C}}^{M\times 2N}\) with \(Q_{ij}=\delta _{ij}\). In fact, for the sake of explicit calculations, it is necessary to replace the GUE by a Lorentzian distribution (irrelevant constants ignored),

$$\begin{aligned} P(H) \sim \text{ det }[ I + H^2]^{-M}, \end{aligned}$$
(1.8)

since in this way \({{\widetilde{H}}}\) is also Lorentzian, and argue separately that in the sense of correlation functions (1.8) is close to a GUE when M is very large [10, Section III]. Under these conditions S becomes Haar distributed on U(2N) as \(M\rightarrow \infty \) and N is fixed; this is the step where \(M\gg N\) is necessary. Furthermore, one can verify [7, Section II.A.1] that at energy \(E=0\) the transmission eigenvalues of a Haar distributed scattering matrix follow the circular ensemble on [0, 1]. Therefore, \(\lambda _i\)’s have a well-known joint distribution

$$ P\big (\{\lambda _i\}\big )\sim \prod _{i<j} (\lambda _j-\lambda _i)^2 \quad \text{ on } \;\;\; [0,1]^N, $$

and their density can be easily computed, yielding (1.5) in the \(N\rightarrow \infty \) limit.

While Beenakker’s result relies on an impressive identity, it allows little flexibility in the inputs: H needs to be Lorentzian with very large dimension; moreover, \(M\gg N\) and \(E=0\) are required. In contrast, our setup allows for a large freedom in choosing the distribution of H, it covers the entire range \(M\ge N\) and also applies to any E in the bulk; however, for computational simplicity we model the contacts W also by a random matrix. While the most relevant regime for scattering on quantum dots is still \(M\gg N\), as scattering involves surface states only, a very recent work [20] introduces absorbing channels well inside the quantum dot that leads to physical models with \(M\sim N\).

The flexibility in our result stems from the fact that our method directly aims at the density of states via an extension of the MDE theory to linearizations of rational functions of random matrices. It seems unnecessarily ambitious, hence requiring too restrictive conditions, to attempt to find the joint distribution of all eigenvalues. Even for Wigner matrices, this is a hopeless task beyond the Gaussian regime. We remark that the present analysis of the density of transmission eigenvalues for the quantum dot is only one convenient application of our approach. This method is powerful enough to answer many related questions concerning the density of states such as the analysis of the scattering poles [19] as well as extensions from quantum dots to quantum wires that we will address in future work.

1.1 Model and Main Theorem

In order to accommodate all parameters that will appear in our analysis, we introduce a generalized version of the scattering matrix (1.1)

$$\begin{aligned} S(w):= I - 2 \,\mathrm {i}\,\gamma \, W^*(w\cdot I -H + \mathrm {i}\,\gamma \, WW^*)^{-1} W \, \end{aligned}$$
(1.9)

with \(\gamma >0\) and \(w\in {\mathbb {C}}_{+}\) for \(\phi \in {(}0,1/2{]}\) and \(w\in {\mathbb {C}}_+ \cup {\mathbb {R}}\) for \(\phi \in {(}1/2,1{]}\), where \({\mathbb {C}}_+ : =\{z \in {\mathbb {C}}: \mathrm {Im}z >0\}\) denotes the complex upper half plane. The constant \(\gamma >0\) quantifies the coupling between the internal quantum dot Hamiltonian H and the external leads \(W_1,W_2\), i.e., the effective Hamiltonian of the open quantum systems becomes \(H-\mathrm {i}\,\gamma W W^{*}\). The spectral parameter \(w\) is used to regularize the potentially unstable inverse in (1.9). Note that for \(\phi \in (0,1/2)\) the matrix \(WW^*\) has a nontrivial kernel (and even for \(\phi =1/2\) it has a very small eigenvalue), hence initially a regularization with a small positive imaginary part is necessary that will later be carefully removed. For the technically easier \(\phi \in {(}1/2,1{]}\) regime, we may directly choose \(w\) to be real since the eigenvalues of \(WW^*\) are bounded away from zero with very high probability.

The general formula (1.1) recovers the scattering matrix (1.1) by setting \(\gamma = \pi \), \(\mathrm {Re}w= E\) and \(\mathrm {Im}w= 0\). The regularized scattering matrix still has the \(2\times 2\)-block structure from (1.2) with \(T \in {\mathbb {C}}^{N \times N}\). We consider the following random matrix model [see (1.1) and (1.2)]

$$\begin{aligned} \varvec{T}_{w,\phi , \gamma } := T T^* = 4 \gamma ^2 W_{2}^*\frac{1}{w-H + \mathrm {i}\,\gamma W W^{*}} W_{1} W_{1}^* \frac{1}{{\overline{w}}-H - \mathrm {i}\,\gamma W W^{*}} W_{2} , \end{aligned}$$
(1.10)

where the triple of parameters \((w, \phi , \gamma )\) belongs to the same set as for (1.9). Furthermore, for \(M,N\in {\mathbb {N}}\), \(\phi := N/M\) the matrices \(H\in {\mathbb {C}}^{M \times M}\) and \(W_{1},W_{2} \in {\mathbb {C}}^{M\times N}\) are three independent random matrix ensembles satisfying the following assumptions

  • (H1qd) H is a Hermitian random matrix having independent (up to symmetry constraints) centered entries of variance 1/M;

  • (H2qd) \(W_{1}\) and \(W_{2}\) are (non-Hermitian) random matrices having independent centered entries of variance 1/M;

  • (H3qd) entries of H, \(W_{1}\) and \(W_{2}\), denoted by H(ij), \(W_{1}(i,j)\) and \(W_{2}(i,j)\) correspondingly, have finite moments of all orders, i.e., there exist \(\varphi _{n} >0\), \(n\in {\mathbb {N}}\), such that

    $$\begin{aligned} \begin{aligned}&\max _{\begin{array}{c} 1\le i,j \le M \end{array}} {\mathbb {E}}\big [\,|\sqrt{M}H(i,j)|^{n}\big ] \\&\quad + \max _{\begin{array}{c} 1\le i \le M \\ 1\le j \le N \end{array}}\Big ({\mathbb {E}}\big [\,|\sqrt{M}W_{1}(i,j)|^{n} \big ]+{\mathbb {E}}\big [\,|\sqrt{M}W_{2}(i,j)|^{n}\big ]\Big ) \le \varphi _{n} . \end{aligned} \end{aligned}$$
    (1.11)

Remark 1.1

(Constant matrices) In (1.10) and later in the paper, for \(B \in {\mathbb {C}}\), \(n\in {\mathbb {N}}\) and \(I_{n}\in {\mathbb {C}}^{n\times n}\) the identity matrix of size n, we use the shorthand notation \(B \cdot I_{n} = B\). This notation is used only when the dimension of \(I_{n}\) can be unambiguously determined from the context.

Remark 1.2

In the sequel, we consider the coupling constant \(\gamma \) to be a fixed positive number; therefore, we will omit the dependence on \(\gamma \) in our notation.

Denote by \(\mu _{\varvec{T}_{w,\phi }}(d\lambda ):=\frac{1}{N}\sum _{i=1}^{N} \delta _{\lambda _{i}}\) the empirical spectral measure of \(\varvec{T}_{w,\phi }\), where \(\lambda _{i}\) are the eigenvalues of the Hermitian matrix \(\varvec{T}_{w,\phi }\). To simplify the presentation, we will assume in this paper that the dimensions of the matrices H, \(W_1\) and \(W_2\) grow over a subsequence \((N, M)= (kn, ln)\), \(n\in {\mathbb {N}}\), i.e., \(N/M=\phi \) is kept fixed. One could easily extend our argument to include the general situation when one considers two sequences \(N=N_n\) and \(M=M_n\) tending to infinity such that \(\phi _n = N_n/M_n \rightarrow \phi \).

We now formulate our two main results. The first one, Theorem 1.3, is the global law for the empirical eigenvalue density; it shows that \(\mu _{\varvec{T}_{w,\phi }}(d\lambda )\) has a deterministic limit denoted by \(\rho _{w,\phi }(d\lambda )\). The second result, Theorem 1.4, contains key properties of \(\rho _{w,\phi }(d\lambda )\). It shows that the regularization in the spectral parameter \(w\) can be removed and that \(\rho _{w,\phi }\) can be continuously extended to \(\phi \rightarrow 0\); moreover, it explicitly identifies its singularities at zero and one.

Theorem 1.3

(Global law). Fix a positive real number \(\gamma >0\) and a rational number \(\phi \in {(}0,1{]}\cap {\mathbb {Q}}\). Let \(w\) be a fixed spectral parameter: for \(\phi \in {(}0,1/2{]}\) we assume that \(w\in {\mathbb {C}}_+ \), while for \(\phi \in {(}1/2,1{]}\) we assume \(w\in {\mathbb {C}}_+ \cup {\mathbb {R}}\). Then, there exists a deterministic probability measure \(\rho _{w,\phi }(d\lambda )\) with \(\mathrm {supp}\, \rho _{w,\phi } \subset [0,1]\) such that \(\mu _{\varvec{T}_{w,\phi }}(d\lambda )\) converges weakly to \(\rho _{w,\phi }(d\lambda )\) in probability (and almost surely) as \(M,N \rightarrow \infty \) with \(N/M = \phi \).

Theorem 1.4

(Properties of the transmission eigenvalue density). Let the numbers \(\gamma , w,\phi \) and the measure \(\rho _{w,\phi }(d\lambda )\) be as in Theorem 1.3.

  1. (i)

    The function

    $$\begin{aligned} \Big ( \big ({(}0,1/2{]}\cap {\mathbb {Q}}\big ) \times {\mathbb {C}}_+ \Big )\cup \Big ( \big ( {(}1/2,1{]}\cap {\mathbb {Q}}\big )\times \big ( {\mathbb {C}}_+ \cup {\mathbb {R}}\big )\Big ) \ni (\phi ,w) \mapsto \rho _{w,\phi }(d\lambda ) \end{aligned}$$

    from Theorem 1.3 with values in probability measures on [0, 1] can be extended to a function on \({(}0,1{]} \times ( {\mathbb {C}}_+ \cup {\mathbb {R}})\) that is continuous in the weak topology. In particular, the weak limit

    $$\begin{aligned} \rho _{E,\phi }(d\lambda )=\lim _{\, {\mathbb {C}}_+ \ni w\rightarrow E}\rho _{w,\phi }(d\lambda ) \end{aligned}$$
    (1.12)

    exists for all \(E \in {\mathbb {R}}\). For every \(E \in {\mathbb {R}}\) and \(\phi \in {(}0,1{]}\) the measure \(\rho _{E,\phi }(d\lambda )\) is absolutely continuous, i.e., \(\rho _{E,\phi }(d\lambda )=\rho _{E,\phi }(\lambda )d \lambda \). The function \(\rho _{E,\phi }(\lambda )\) is bounded when \(\lambda \) is away from 0 and 1. Furthermore, the weak limit \(\rho _{E,0}(d\lambda ):=\lim _{\phi \downarrow 0}\rho _{E,\phi }(d\lambda )\) exists for every \(E \in {\mathbb {R}}\).

  2. (ii)

    If \(\phi = 1\) and \(\gamma >0\), then \(\rho _{E,\phi }(\lambda )\) has the following asymptotics near 0 and 1:

    1. (a)

      for \(E\in {\mathbb {R}}\)

      $$\begin{aligned} \rho _{E, 1}(\lambda ) = \frac{1}{\pi }\root 3 \of {\frac{1+E^2}{4 \gamma ^{2} }} \, \lambda ^{-2/3} + O\left( \lambda ^{-1/3}\right) \text{ as } \lambda \rightarrow 0_{+}, \end{aligned}$$
      (1.13)
    2. (b)

      for \(|E| < \frac{1}{ \gamma } (2 \sqrt{1 + 6 \gamma ^2 + \gamma ^4 }-2 \gamma ^2 - 2 )^{1/2}\)

      $$\begin{aligned} \rho _{E, 1}(\lambda ) = \frac{1}{\pi } \, \frac{ -4 \xi _{0}}{\xi _{0}^{2}+ \gamma ^2 E^2 + 4 }\, (1-\lambda )^{-1/2} + O\left( 1\right) \text{ as } \lambda \rightarrow 1_{-} , \end{aligned}$$
      (1.14)

      where \(\xi _{0} = -\sqrt{2 \sqrt{1 + 6 \gamma ^2 + \gamma ^4 }-2 \gamma ^2 -2- \gamma ^2 E^2}\).

  3. (iii)

    If \(\phi \in (0,1)\) and \(\gamma >0\), then

    1. (a)

      for \(E\in {\mathbb {R}}\)

      $$\begin{aligned} \rho _{E,\phi }(\lambda ) = \frac{1}{\pi } \, \frac{ \gamma ^2 E^2 \xi _0^4+ \big (\xi _0^2+2\phi \big )^2\big (4+ \gamma ^2 \xi _0^2\big )}{4 \gamma \xi _0\big (\xi _0^2+2\phi \big )^2}\, \lambda ^{-1/2} + O\left( 1\right) \text{ as } \lambda \rightarrow 0_{+}, \end{aligned}$$
      (1.15)

      where \(\xi _{0}>0\) is the unique positive solution of

      $$\begin{aligned} \xi _0^6 + \big (E^2+8\phi -4\big ) \xi _0^4+ 4 \phi \big (E^2+ 5\phi -4\big ) \xi _0^2+16(\phi -1)\phi ^2=0. \end{aligned}$$
      (1.16)
    2. (b)

      for \(|E| < E_*:=\frac{1}{ \gamma }\Big (2 \sqrt{ 1 + 2 (1+2\phi ) \gamma ^2 + (2 \phi -1)^2 \gamma ^4 } - 2 \gamma ^2 (2\phi -1) -2\Big )^{1/2}\) the density satisfies

      $$\begin{aligned} \rho _{E, \phi }(\lambda ) = \frac{1}{\pi } \, \frac{-4 \xi _1}{\xi _1^2 + \gamma ^2 E^2+4 }\,(1-\lambda )^{-1/2} + O\left( 1\right) \text{ as } \lambda \rightarrow 1_{-} , \end{aligned}$$
      (1.17)

      where

      $$\begin{aligned} \xi _1 = - \gamma \sqrt{E_*^2- E^2}. \end{aligned}$$
      (1.18)
  4. (iv)

    If \(\phi =0\), \(\gamma >0\) and \(|E| < 2\) the density is given globally by the explicit formula

    $$\begin{aligned} \rho _{E, 0 }(\lambda ) = \frac{1}{ \pi }\frac{\gamma \big (1+ \gamma ^2 \big ) }{\big (1+ \gamma ^2 \big )^2\lambda + \gamma ^2 \big (4-E^2\big )(1-\lambda )} \,\sqrt{\frac{4-E^2}{\lambda (1-\lambda )}}, \qquad \lambda \in (0,1). \end{aligned}$$
    (1.19)

Proofs of Theorems 1.3 and 1.4

We now explain the structure of the paper that contains the proof of both theorems. The density function \(\rho _{E, \phi }(\lambda )\) will be derived from a deterministic equation, the Dyson equation of the linearizaton of (1.10) [see Eqs. (2.13) and (2.27) for \(\phi =1\), as well as (3.7) and (3.18) for \(\phi \in (0,1)\) ]. This equation plays a central role in our analysis. Sections 2 and 3 contain the proofs of the model specific parts of Theorem 1.4 which use some key conclusions of the general theory on noncommutative rational functions developed in Appendix A. The \(\phi =1\) case is treated in Sect. 2 with full details, while in Sect. 3 we explain the modifications for the general rational \(\phi \in (0,1)\) case. Theorem 1.3 is proven in Lemma 2.5 for \(\phi = 1\) and Lemma 3.4 for general rational \(\phi \in (0,1)\) using the general global law from Appendix A. Part (ii) is proven in Sect. 2.5 after having established key properties of the solution for the Dyson equation. Section 3 follows the same structure as Sect. 2 and proves parts (i), (iii) and (iv) of Theorem 1.4 for the \(\phi \in [0,1)\) case. \(\square \)

Remark 1.5

The restriction \(|E|<E_*\) in Theorem 1.4 is used to identify the regime in which the density \(\rho _{E,\phi }\) has two singularities. For \(|E|>E_*\), the singularity at \(\lambda = 1\) disappears and the support of the density becomes bounded away from \(\lambda =1\). Physically, this indicates that there are no fully open channels in this regime. This effect is most severe in the case \(\phi =0\), where the system becomes completely reflective as |E| increases above the threshold \(E_*=2\). Our approach is also applicable for \(|E|>E_*\), but for the simplicity of the presentation, we restrict our analysis to the physically more relevant situation when two singularities exist simultaneously.

We formulated Theorem 1.3 about the specific matrix (1.10). However, our method works for very general noncommutative (NC) rational expressions in large matrices with i.i.d. entries (with or without Hermitian symmetry) generalizing our previous work [13] on polynomials. For convenience of the readers interested only in the concrete scattering problem, the main part of our paper focuses on this model and we defer the general theory to Appendix A. The details of this appendix are not necessary in order to follow the main arguments in Sects. 2 and 3 provided some basic facts from the appendix are accepted. These facts will be re-stated for the specialization to our concrete operator \(\varvec{T}_{ w,\phi }\) from (1.10). On the other hand, Appendix A is written in a self-contained form, so readers interested in the general theory can directly go to it, skipping the concrete application.

In Appendix A, we first give the precise definition of the set of admissible rational expressions that requires some technical preparation, see Sect. A.1 for details. Roughly speaking, we can consider any rational expression whose denominators are stably invertible with overwhelming probability. We remark that this property holds for (1.10) since the imaginary part of \(w-H + \mathrm {i}\,\gamma W W^{*}\) has a positive lower bound as long as \(\phi >1/2\) or \(w\in {\mathbb {C}}_+ \). With this definition at hand, we develop the theory of global and local laws as well as the identification of the pseudospectrum for such rational expressions in Sects. A.6 and A.5, respectively.

2 Proof of Theorem 1.4 for the Special Case \(\phi = 1\)

In this section, we study the model (1.10) for fixed \(\gamma >0\), \(\phi = 1\) and \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\). This special choice of parameter \(\phi \) ensures that the linearization of \(\varvec{T}_{ w,\phi }\) has a fairly simple structure, which makes the proof of Theorems 1.3 and 1.4 more transparent and streamlined. Generalization to \(\phi \in (0,1) \) is postponed to Sect. 3. Since the parameter \(\phi \) is fixed to be equal to 1, we will omit the dependence on \(\phi \) in the notation throughout the current section. The information about linearizations of general rational functions is collected in the Appendix A.1. Here, we often refer to specialization of these results to \(\varvec{T}_{ w}\).

2.1 Linearization Trick and the Dyson Equation for Linearization

We consider the random matrix model \(\varvec{T}_{ w}\) defined in (1.10) as a self-adjoint rational function of random matrices H, \(W_1\) and \(W_2\). In order to study its eigenvalues, we introduce the self-adjoint linearization matrix \(\varvec{H}_{ w}\in {\mathbb {C}}^{8N\times 8N}\)

$$\begin{aligned} \varvec{H}_{ w} := \left( \begin{array}{ccc|ccc|cc} 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad W_{2}^* \\ 0 &{}\quad 0 &{}\quad \frac{\mathrm {i}\,}{ \gamma } &{}\quad 0 &{}\quad 0 &{} \quad 0 &{} \quad 0 &{}\quad W_{2}^* \\ 0 &{}\quad -\frac{\mathrm {i}\,}{ \gamma } &{} \quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad W_{2}^* &{}\quad 0 \\ \hline 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad \frac{\mathrm {i}\,}{ \gamma } &{}\quad 0 &{}\quad W_{1}^* \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad -\frac{1}{4 \gamma ^2} &{}\quad 0 &{}\quad W_{1}^* &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad -\frac{\mathrm {i}\,}{ \gamma }&{}\quad 0 &{}\quad 0 &{}\quad W_{1}^* &{}\quad 0 \\ \hline 0 &{}\quad 0 &{}\quad W_{2} &{}\quad 0 &{}\quad W_{1} &{}\quad W_{1} &{}\quad 0 &{} \quad w-H \\ W_{2} &{}\quad W_{2} &{}\quad 0 &{}\quad W_{1} &{}\quad 0 &{}\quad 0 &{}\quad {\overline{w}} -H &{}\quad 0 \\ \end{array} \right) \end{aligned}$$
(2.1)

with \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\). Denote by \(J_{m} \in {\mathbb {C}}^{m\times m}\), \(m\in {\mathbb {N}}\), a matrix whose (1, 1) entry is equal to 1 and all other entries are equal to 0. For any \(n\in {\mathbb {N}}\) and \(R\in {\mathbb {C}}^{n \times n}\), we define \(\Vert R\Vert \) to be the operator norm of R. The following definition is a specialization of the notion of generalized resolvent from Appendix A.3.

Definition 2.1

(Generalized resolvent). We call the matrix-valued function \(z \mapsto (\varvec{H}_{w} - z J_{8}\otimes I_N)^{-1}\) defined for \(z\in {\mathbb {C}}_+\) the generalized resolvent of \(\varvec{H}_{w}\).

The results below are formulated using the notion of asymptotically overwhelming probability.

Definition 2.2

We say that a sequence of events \(\{\Omega _{N}\}_{N\in {\mathbb {N}}}\) holds asymptotically with overwhelming probability (a.w.o.p. for short) if for any \(D>0\) there exists \(C_{D}>0\) such that

$$\begin{aligned} {\mathbb {P}}[\,\Omega _{N}] \ge 1-\frac{C_{D}}{N^{D}} . \end{aligned}$$
(2.2)

Consider the set

$$\begin{aligned} \Theta _N := \Big \{\Vert H\Vert \le 3,\, \Vert W_{1}\Vert \le 3, \, \Vert W_2\Vert \le 3, \,\Vert (W W^*)^{-1}\Vert \le 12 \Big \}\; . \end{aligned}$$
(2.3)

Note that \( W W^*\) is a sample covariance matrix with concentration ratio 1/2, hence its spectrum is asymptotically supported on an arbitrarily small neighborhood of \([(1-\frac{1}{\sqrt{2}})^{2}, (1+\frac{1}{\sqrt{2}})^{2}]\) with very high probability. The limiting support follows from the classical Bai–Yin theorem [6], the corresponding large deviation result under somewhat different conditions follows, e.g., from Corollary V.2.1 of [17]. In fact, the boundedness of \(\Vert (WW^*)^{-1}\Vert \) and \(\Vert W_i\Vert \) also follows from [33], at least for subgaussian distributions. Alternatively, under our conditions (1.11) this result also follows from Lemma B.1 in the Appendix by choosing \(n=M\), \(l=1\) and \(m=2\). Similarly, from the properties of classical Wigner and iid ensembles (see, e.g., [5, Section 5]), we obtain that the event \(\{\,\Vert H\Vert \le 3,\, \Vert W_{1}\Vert \le 3, \, \Vert W_2\Vert \le 3 \, \}\) also holds a.w.o.p., and we conclude that for any \(D>0\) there exists \(C_D>\) such that

$$\begin{aligned} {\mathbb {P}}[\Theta _{N}] \ge 1- \frac{C_{D}}{N^{D}} . \end{aligned}$$
(2.4)

In the rest of this section, we consider the random matrix models (1.9) and (1.10) restricted to the set \(\Theta _N\). Since on this set the smallest eigenvalue of \(WW^*\) cannot be smaller that 1/12, we have that \(\Vert (w-H + \mathrm {i}\,\gamma W W^{*})^{-1} \Vert \le \frac{1}{12 \gamma }\) for all \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\), and thus, the model (1.10) is well defined on \(\Theta _N\). In the next lemma, we establish an a priori bound for the generalized resolvent \((\varvec{H}_{w} - z J_{8}\otimes I_N)^{-1}\).

Lemma 2.3

(Basic properties of the generalized resolvent). 

  1. (i)

    For any \( \gamma >0\) there exists \(C_{\gamma }>0\) such that a.w.o.p.

    $$\begin{aligned} \big \Vert (\varvec{H}_{ w} - z J_{8} \otimes I_N)^{-1}\big \Vert \le C_{\gamma } \bigg (1 + \frac{1}{\mathrm {Im}z}\bigg ) \end{aligned}$$
    (2.5)

    uniformly for all \(z\in {\mathbb {C}}_{+}\) and \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\).

  2. (ii)

    For all \( w\in {\mathbb {C}}_+ \cup {\mathbb {R}}\), \(z\in {\mathbb {C}}_{+}\) and \(1\le i,j\le N\)

    $$\begin{aligned} \Big [\big (\varvec{H}_{ w}-zJ_{8}\otimes I_N\big )^{-1}\Big ]_{ij} = \Big [\big (\varvec{T}_{ w}- zI_N\big )^{-1}\Big ]_{ij} ,\quad 1\le i,j \le N . \end{aligned}$$
    (2.6)

Proof

Let \(\varvec{H}_{ w}^{\mathrm {(init)}}\) be the linearization matrix obtained via the linearization algorithm described in Appendix A.2

$$\begin{aligned} \varvec{H}_{ w}^{\mathrm {(init)}} = \left( \begin{array}{cccccccc} 0&{}0&{}0&{}0&{}0&{}2 \gamma W_{2}^{*}&{}0&{}0 \\ 0&{}0&{}0&{}0&{}W_{1}&{} -(w-H) &{} -\sqrt{ \gamma } W_{1} &{} -\sqrt{ \gamma } W_{2} \\ 0&{}0&{}0&{}0&{}0&{} -\sqrt{ \gamma } W_{1}^{*} &{} -\mathrm {i}\,&{} 0 \\ 0&{}0&{}0&{}0&{}0&{} -\sqrt{ \gamma } W_{2}^{*} &{} 0 &{} -\mathrm {i}\,\\ 0&{} W_{1}^{*} &{}0&{}0&{}-1&{}0&{}0&{}0 \\ 2 \gamma W_{2} &{} -( {\overline{w}} -H) &{} -\sqrt{ \gamma } W_{1} &{} - \sqrt{ \gamma } W_{2} &{}0&{}0&{}0&{}0 \\ 0&{}-\sqrt{ \gamma } W_{1}^{*} &{} \mathrm {i}\,&{}0&{}0&{}0&{}0&{}0 \\ 0&{}-\sqrt{ \gamma } W_{2}^{*} &{} 0&{}\mathrm {i}\,&{}0&{}0&{}0&{}0 \end{array} \right) . \end{aligned}$$
(2.7)

Denote by \(\{E_{ij}, 1\le i,j \le 8 \,\}\) the standard basis of \({\mathbb {C}}^{\,8\times 8}\)

$$\begin{aligned} E_{ij} = \big (\delta _{ki}\delta _{jl}\big )_{k,l=1}^{8} , \end{aligned}$$

where \(\delta _{\alpha \beta }\) is the Kronecker delta. Then, one can easily check that \(\varvec{H}_{ w}\) in (2.1) can be obtained from \(\varvec{H}_{ w}^{\mathrm {(init)}}\) by applying the following transformation

$$\begin{aligned} \varvec{H}_{ w} = {\widetilde{T}}{\widetilde{P}}_{78}{\widetilde{P}}_{23}{\widetilde{P}}_{34}{\widetilde{P}}_{67}{\widetilde{P}}_{28} \varvec{H}_{ w}^{\mathrm {(init)}} {\widetilde{P}}_{28}{\widetilde{P}}_{67}{\widetilde{P}}_{34}{\widetilde{P}}_{23}{\widetilde{P}}_{78} {\widetilde{T}} \end{aligned}$$
(2.8)

with \({\widetilde{T}} = \mathrm {diag}\big (1,-2\sqrt{ \gamma },\frac{1}{2 \gamma ^{3/2}} , -2\sqrt{ \gamma },-\frac{1}{2 \gamma } , \frac{1}{2 \gamma ^{3/2}}, -2 \gamma , \frac{1}{2 \gamma }\big ) \otimes I_{N}\), \({\widetilde{P}}_{ij}= \big ( E_{ij} + E_{ji} + \sum _{l \notin \{i,j\}} E_{ll} \big ) \otimes I_N\). Note that all transposition matrices \({\widetilde{P}}_{ij}\) in (2.8) leave the upper-left \(N\times N\) block intact. Thus, (2.6) follows from the Definition A.5 of linearization and the Schur complement formula [see, e.g., (A.20)] by taking \({\mathcal {A}}= {\mathbb {C}}^{N\times N}\).

In order to prove the bound (2.5), consider the set \(\Theta _N\) defined in (2.3). One can see that \(\varvec{H}_{ w}^{\mathrm {(init)}}\) satisfies the bound (2.5) by specializing Lemma A.10 for \({\mathcal {A}}= {\mathbb {C}}^{N\times N}\), \(x_1 = H\), \(y_{1} = W_{1}\), \(y_{2} = W_{2}\), \(C = 12/\gamma \) and the rational expression q being (1.10) on the set \(\Theta _N\), as well as using the standard relation between the operator and max norms, similarly as in, e.g., (A.47). On the other hand, \(\varvec{H}_{ w}^{\mathrm {(init)}}\) and \(\varvec{H}_{ w}\) are related by (2.8). For any \(R\in {\mathbb {C}}^{8\times 8}\), applying the transformation \(R \mapsto {\widetilde{P}}_{ij} R {\widetilde{P}}_{ij}\) does not change the norm of R, while applying the map \(R \mapsto {\widetilde{T}} R \, {\widetilde{T}}\) might change the norm by an irrelevant nonzero constant factor only. We thus conclude that \(\varvec{H}_{ w}\) also satisfies the bound (2.5) with a constant \(C_{\gamma }\) being possibly different than the one for \(\varvec{H}_{ w}^{\mathrm {(init)}}\). Since the set \(\Theta _N\) satisfies (2.4), the bound (2.5) holds a.w.o.p. \(\square \)

Define

$$\begin{aligned} \kappa _{1} = \left( \begin{array}{ccc} 0&{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \frac{\mathrm {i}\,}{ \gamma } \\ 0 &{}\quad -\frac{\mathrm {i}\,}{ \gamma } &{}\quad 0 \end{array} \right) ,\quad \kappa _{2} = \left( \begin{array}{ccc} 0&{}\quad 0&{}\quad \frac{\mathrm {i}\,}{ \gamma } \\ 0 &{}\quad -\frac{1}{4 \gamma ^2} &{}\quad 0 \\ -\frac{\mathrm {i}\,}{ \gamma } &{}\quad 0 &{}\quad 0 \end{array} \right) \end{aligned}$$
(2.9)

and

$$\begin{aligned} \kappa _{3} = \left( \begin{array}{cc} 0&{}\quad w\\ {\overline{w}} &{}\quad 0 \end{array} \right) ,\quad \kappa _{4} = \left( \begin{array}{ccc} 0&{}\quad 1 &{}\quad 1 \\ 1 &{}\quad 0 &{}\quad 0 \end{array} \right) ,\quad \kappa _{5} = \left( \begin{array}{ccc} 0 &{}\quad 0 &{}\quad 1 \\ 1 &{}\quad 1 &{}\quad 0 \end{array} \right) . \end{aligned}$$
(2.10)

With this notation \(\varvec{H}_{ w}\) can be rewritten as

$$\begin{aligned} \varvec{H}_{ w} = K_0( w)\otimes I_N + K_1 \otimes H + L_1\otimes W_1 + L_1^{*}\otimes W_1^* + L_2\otimes W_2 + L_2^{*}\otimes W_2^* , \end{aligned}$$
(2.11)

where \(K_0=K_0( w),K_1,L_1,L_2 \in {\mathbb {C}}^{8\times 8}\) are given by their block structures as

$$\begin{aligned}&K_0 = \left( \begin{array}{ccc} \kappa _{1}&{}\quad 0 &{}\quad 0 \\ 0 &{}\quad \kappa _{2} &{}\quad 0 \\ 0 &{}\quad 0 &{} \quad \kappa _{3} \end{array} \right) ,\quad K_1 = \left( \begin{array}{ccc} 0&{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad - \sigma _{1} \end{array} \right) ,\quad L_1 = \left( \begin{array}{ccc} 0&{}\quad 0&{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ 0 &{} \kappa _4 &{} 0 \end{array} \right) ,\nonumber \\&L_2 = \left( \begin{array}{ccc} 0&{}\quad 0&{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 \\ \kappa _5 &{}\quad 0 &{}\quad 0 \end{array} \right) , \end{aligned}$$
(2.12)

and \(\sigma _{1} = \left( \begin{array}{cc} 0 &{} \quad 1 \\ 1 &{} \quad 0 \end{array} \right) \) is the usual Pauli matrix. Consider the Dyson equation for linearization (DEL)

$$\begin{aligned} -\frac{1}{M} = z J_{8} - K_0 (w) + \Gamma \big [ M\big ] \end{aligned}$$
(2.13)

with a linear map \(\Gamma : {\mathbb {C}}^{\,8\times 8} \rightarrow {\mathbb {C}}^{\,8\times 8}\) given by

$$\begin{aligned} \Gamma \big [R\big ] := K_{1} R K_{1} + L_1 R L_1^* + L_1^* R L_1+ L_2 R L_2^* + L_2^* R L_2 ,\qquad R\in {\mathbb {C}}^{\,8\times 8}. \end{aligned}$$
(2.14)

Lemma 2.4

(Existence and basic properties of the solution to the DEL (2.13)) For any \(\gamma >0\), \(w\in {\mathbb {C}}_{+} \cup {\mathbb {R}}\) and \(z \in {\mathbb {C}}_{+}\) define \(M_{z, w} \in {\mathbb {C}}^{\,8\times 8}\) as

$$\begin{aligned} M_{z, w}:= & {} ({\mathrm {id}}_8 \otimes \, \tau _{{\mathcal {S}}}) \Big ( \big (K_0( w) - z J_{8}\big )\otimes \mathbb {1}_{{\mathcal {S}}}+ K_{1}\otimes s\nonumber \\&+ L_1\otimes c_1 +L_1^{*}\otimes c_1^* + L_2 \otimes c_2 + L_2^{*} \otimes c_2^*\Big )^{-1} , \end{aligned}$$
(2.15)

where \(s,c_1,c_2\) are freely independent semicircular and circular elements and \(\tau _{{\mathcal {S}}}\) is a tracial state on a \(C^*\)-probability space \(({\mathcal {S}}, \tau _{{\mathcal {S}}})\) with the unit element \(\mathbb {1}_{{\mathcal {S}}}\). Then,

  1. (i)

    For any \(\gamma >0\) there exists \( C_{\gamma }>0\) such that \(M_{z, w}\) satisfies the a priori bound

    $$\begin{aligned} \Vert M_{z, w}\Vert \le C_{\gamma } \Big (1 + \frac{1}{\mathrm {Im}z}\Big ) \end{aligned}$$
    (2.16)

    uniformly for all \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\).

  2. (ii)

    For any \(\gamma >0\), \(w\in {\mathbb {C}}_{+} \cup {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\), matrix \(M_{z, w}\) satisfies the DEL (2.13) and has positive semidefinite imaginary part, \(\mathrm {Im}M_{z, w} \ge 0\). Moreover, for all \(\gamma >0\) and \(w\in {\mathbb {C}}_{+} \cup {\mathbb {R}}\), the matrix-valued function \(z\mapsto M_{z, w}\) is analytic on \({\mathbb {C}}_{+}\).

  3. (iii)

    For any \(\gamma >0\), \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\) function \(z\mapsto M_{z, w}\) admits the representation

    $$\begin{aligned} M_{z, w} = M^{\infty }_{ w} + \int _{{\mathbb {R}}} \frac{V_{ w}(d\lambda )}{\lambda - z} , \end{aligned}$$
    (2.17)

    where \(M_{w}^{\infty } \in {\mathbb {C}}^{\,8\times 8}\), and \(V_{ w}(d\lambda )\) is a positive semidefinite matrix-valued measure on \({\mathbb {R}}\) with compact support. In particular, \(\lim _{z\rightarrow \infty } M_{z, w} = M^{\infty }_{ w}\).

Proof

Fix \(\gamma >0\) and denote the noncommutative rational (in fact, polynomial) expression

$$\begin{aligned} q_{1,w}(x,y_1,y_2,y_1^*,y_2^*) := w-x + \mathrm {i}\,\gamma (y_1y_1^* + y_2 y_2^*) . \end{aligned}$$
(2.18)

Using the fact that \(c_1 c_1^* + c_2 c_2^*\) has free Poisson distribution of rate 2, which in particular implies that \(\Vert (c_1 c_1^* + c_2 c_2^*)^{-1}\Vert \le (1-\frac{1}{\sqrt{2}})^{-1}\), we conclude that the triple \((s,c_1,c_2)\) belongs to the effective domain \({\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) with \(C = \gamma ^{-1} (1-\frac{1}{\sqrt{2}})^{-1}\) for all \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\) (see Sect. A.1 for the corresponding definitions).

Following the proofs of Lemmas A.7 and A.10 and specializing them to our concrete case, we obtain that for any fixed \(\gamma >0\) there exists \(C_{\gamma }>0\) such that

$$\begin{aligned}&\Big \Vert \big ( (K_0(w) {-} z J_{8})\otimes \mathbb {1}_{{\mathcal {S}}}{+} K_{1}\otimes s{+} L_1\otimes c_1 {+}L_1^{*}\otimes c_1^* \nonumber \\&\quad + L_2 \otimes c_2 {+} L_2^{*} \otimes c_2^*\big )^{-1} \Big \Vert _{{\mathbb {C}}^{\,8\times 8} \,\otimes \,{\mathcal {S}}} \le C_{\gamma } \Big (1 + \frac{1}{\mathrm {Im}z}\Big ) \end{aligned}$$
(2.19)

uniformly for all \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\), which yields the a priori bound (2.16). Part (ii) of Lemma 2.4 now follows directly from parts (ii)–(iv) of Lemma A.12. Finally, (2.17) follows from the representation (A.27) in Lemma A.12. \(\square \)

With DEL (2.13), we associate the corresponding stability operator \({\mathscr {L}}_{z, w}:{\mathbb {C}}^{\,8\times 8} \rightarrow {\mathbb {C}}^{\,8\times 8}\) given by

$$\begin{aligned} {\mathscr {L}}_{z, w}\big [R \,\big ] := R - M_{z, w} \, \Gamma \big [R\,\big ] \,M_{z, w} ,\quad R\in {\mathbb {C}}^{\,8\times 8} . \end{aligned}$$
(2.20)

The following lemma is directly obtained from Proposition A.19 and establishes Theorem 1.3 and the weak limits in the part (i) of Theorem 1.4 for \(\phi =1\).

Lemma 2.5

(Global law for \(\mu _{\varvec{T}_{ w}}\)). For \( w\in {\mathbb {C}}_{+} \cup {\mathbb {R}}\) the empirical spectral measure \(\mu _{\varvec{T}_{ w}}(d\lambda )\) converges weakly in probability (and almost surely) to \(\rho _{ w}(d\lambda )\), where \(\rho _{ w}(d\lambda ):=\langle e_1, V_{ w}(d\lambda ) \, e_1 \rangle \) is the (1, 1) component of the matrix-valued measure \(V_{ w}(d\lambda )\) from (2.17). Moreover, \(\mathrm {supp} \, \rho _{ w}(d\lambda ) \subset [0,1]\) for all \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\) and \(\rho _{ w}(d\lambda )\) converges weakly to \(\rho _{ E }(d\lambda )\) as \(w\in {\mathbb {C}}_+\) tends to \(E\in {\mathbb {R}}\).

Proof

We apply Proposition A.19 to the rational expression in random matrices (1.10). By (A.78), for any \(w\in {\mathbb {C}}_{+}\) and fixed \(\theta >0\) the generalized resolvent of the linearization \((\varvec{H}_{ w} - z J_{8} \otimes I_N)^{-1}\), corresponding to \(\varvec{G}_{z}\) in (A.78), converges uniformly on \(\{\,z : \, \mathrm {Im}z \ge \theta ^{-1}, \, |z| \le \theta \,\}\) to \(M_{z,w} \otimes I_N\), corresponding to \(M_{z}^{(\mathrm {sc})} \otimes I_N\) in (A.78). In particular, the identity (2.6), similarly to (A.79) (see also Definition A.17), implies the pointwise convergence of the Stieltjes transform of the empirical spectral measure of \(\varvec{T}_{w}\): for any \(\theta , \varepsilon , D > 0\) there exists a constant \(C_{\theta , \varepsilon ,D}>0\) such that

$$\begin{aligned} {\mathbb {P}}\bigg [\,\Big |\frac{1}{N}{\mathrm {Tr}}\, \Big (\varvec{T}_{w} - z I_N \Big )^{-1} - M_{z, w}(1,1)\Big | \ge \frac{N^{\varepsilon }}{N} \bigg ] \le \frac{C_{\theta , \varepsilon ,D}}{N^{D}} \end{aligned}$$
(2.21)

uniformly on \(\{\,z : \, \mathrm {Im}z \ge \theta ^{-1}, \, |z| \le \theta \,\}\) and \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\). The convergence in (2.21) together with (2.17) imply that the weak convergence

$$\begin{aligned} \lim _{N\rightarrow \infty } \mu _{\varvec{T}_{w}}(d\lambda ) = \rho _{ w}(d\lambda ) \end{aligned}$$
(2.22)

holds in probability (see, e.g., [4, Theorem 2.4.4]).

Now we prove the almost sure convergence. Take any \(z_0\in {\mathbb {C}}_{+}\), a sequence \(\{z_i\}\subset {\mathbb {C}}_{+}\) of different complex numbers with \( z_{i} \rightarrow z_{0}\) such that \(z_0\) is the accumulation point of \(\{z_i\}\). Define the sequence of events

$$\begin{aligned} A_N:= \bigg \{\max _{1\le i \le N}\Big \{\Big |\frac{1}{N}{\mathrm {Tr}}\, \Big (\varvec{T}_{w} - z_i I_N \Big )^{-1} - M_{z_{i}, w}(1,1)\Big |\Big \} \ge \frac{1}{\sqrt{N}}\bigg \} . \end{aligned}$$
(2.23)

Then it follows from (2.21) and the Borel–Cantelli lemma applied to \(\{A_N\}\) that with probability 1

$$\begin{aligned} \lim _{N\rightarrow \infty }\frac{1}{N}{\mathrm {Tr}}\, \Big (\varvec{T}_{w} - z_i I_N \Big )^{-1} = M_{z_i, w}(1,1) \end{aligned}$$
(2.24)

for all \(i\in {\mathbb {N}}\). Finally, applying the Vitali–Porter theorem we conclude that the weak convergence (2.22) holds almost surely.

To prove the bound on the support of \(\rho _{ w}(d\lambda )\) note that the scattering matrix \(S( w)\), related to \(\varvec{T}_{ w}= T T^{*}\) via (1.2), is unitary. This implies that all singular values of T are located in the interval [0, 1], thus \(\mathrm {supp}\, \mu _{\varvec{T}_{ w}} \subset [0,1]\). But from (2.22), we know that the empirical spectral measure of \(T T^{*}\) converges weakly to \(\rho _{w}(d\lambda )\), which yields \(\mathrm {supp} \rho _{w} \subset [0,1]\).

It follows from (2.19) and the definition of \(M_{z,w}\) in (2.15) that for any \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\), \(E\in {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\)

$$\begin{aligned} \Vert M_{z,w} - M_{z,E}\Vert \le 2\, C_{\gamma }^{2} \,\big |w- E\big |\, \Big (1 + \frac{1}{\mathrm {Im}z}\Big )^2 . \end{aligned}$$
(2.25)

This implies the pointwise convergence

$$\begin{aligned} \lim _{w\rightarrow E} M_{z,w} = M_{z,E} \end{aligned}$$
(2.26)

for all \(z\in {\mathbb {C}}_{+}\). By (2.17) and \(\rho _{ w}(d\lambda )=\langle e_1, V_{ w}(d\lambda ) \, e_1 \rangle \), the (1, 1)-components \(M_{z,w}(1,1)\) and \(M_{z,E}(1,1)\) define the Stieltjes transforms of the measures \(\rho _{w}(d\lambda )\) and \(\rho _{E}(d\lambda )\) correspondingly, from which we conclude that (2.26) yields the weak convergence of \(\rho _{w}(d\lambda )\) to \(\rho _{E}(d\lambda )\) as \(w\rightarrow E\). \(\square \)

Note that \(M_{z, w}\) is a matrix-valued Herglotz function. Therefore, from the properties of the (matrix-valued) Herglotz functions (see, e.g., [22, Theorems 2.2 and 5.4]), the absolutely continuous part of \(\rho _{ w}(d\lambda )\) is given by the inverse Stieltjes transform of \(M_{z, w}(1,1)\) (see Lemma A.12)

$$\begin{aligned} \rho _{ w}(\lambda ) := \lim _{\eta \downarrow 0} \frac{1}{\pi } \mathrm {Im}M_{\lambda +\mathrm {i}\,\eta , w}(1,1) . \end{aligned}$$
(2.27)

We call the function \(\rho _{ w}(\lambda )\), defined in (2.27), the self-consistent density of states of the solution to the DEL (2.13). It will be shown in Sect. 2.5 that \(\rho _{ w}(d\lambda )\) is in fact purely absolutely continuous, i.e., \(\rho _{ w}(d\lambda ) = \rho _{ w}(\lambda ) d\lambda \). The statements (a) and (b) of part (ii) of Theorem 1.4 will be derived from the study of \(M_{z, w}\) for the spectral parameter z being close to the real line.

Note that our particular choice of linearization (2.1) allows rewriting the original DEL (2.13) in a slightly simpler form. More precisely, if

$$\begin{aligned} R = \left( \begin{array}{ccc} R_{11}&{}\quad R_{12}&{} \quad R_{13} \\ R_{21}&{}\quad R_{22}&{}\quad R_{23} \\ R_{31}&{}\quad R_{32}&{}\quad R_{33} \end{array} \right) \in {\mathbb {C}}^{8\times 8} \end{aligned}$$
(2.28)

with \(R_{11}, R_{22} \in {\mathbb {C}}^{3\times 3}\) and \(R_{33} \in {\mathbb {C}}^{2\times 2}\), then (2.12) yields

$$\begin{aligned} \Gamma [R] = \left( \begin{array}{ccc} \kappa _{5}^{t} R_{33} \kappa _{5}&{}\quad 0 &{}\quad 0 \\ 0 &{}\quad \kappa _{4}^{t} R_{33} \kappa _{4}&{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \sigma _{1} R_{33} \sigma _{1} + \kappa _{4} R_{22} \kappa _{4}^{t} + \kappa _{5} R_{11} \kappa _{5}^{t} \end{array} \right) , \end{aligned}$$
(2.29)

so that the image \(\Gamma [R]\) is a block-diagonal matrix. Together with the definition of \(K_0\) in (2.12), this implies that the right-hand side in (2.13) is a block-diagonal matrix with blocks of sizes 3, 3, and 2 correspondingly. We conclude that any solution to the DEL (2.13) has a block-diagonal form, which, in particular, allows us to write

$$\begin{aligned} M_{z, w} = \left( \begin{array}{ccc} M_{1} &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad M_{2} &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad M_{3} \end{array} \right) \end{aligned}$$
(2.30)

with \(M_{1},M_{2} \in {\mathbb {C}}^{3\times 3}\) and \(M_{3} \in {\mathbb {C}}^{2\times 2}\), where we omit the dependence of the blocks on z and \( w\). Now DEL (2.13) can be decomposed into a system of three matrix equations of smaller dimensions

$$\begin{aligned} -\frac{1}{M_{1}} = zJ_{3}-\kappa _{1}+ \kappa _{5}^t M_{3} \kappa _{5} ,\quad -\frac{1}{M_{2}} = -\kappa _{2}+ \kappa _{4}^t M_{3} \kappa _{4} \end{aligned}$$
(2.31)

and

$$\begin{aligned} -\frac{1}{M_{3}}= & {} - \kappa _{3} (w) - \frac{ 1 }{ -\frac{1}{2 \gamma ^2z}(I_2 + \sigma _3)-\frac{1}{ \gamma }\sigma _2 + M_{3}} \nonumber \\&- \frac{1}{-2(I_2 - \sigma _3)-\frac{1}{ \gamma }\sigma _2+ M_{3} } + \sigma _{1} M_{3} \sigma _{1} , \end{aligned}$$
(2.32)

where \(\sigma _{1} = \left( \begin{array}{cc} 0 &{}\quad 1 \\ 1 &{}\quad 0 \end{array} \right) \), \(\sigma _{2}= \left( \begin{array}{cc} 0 &{}\quad -\mathrm {i}\,\\ \mathrm {i}\,&{}\quad 0 \end{array} \right) \) and \(\sigma _{3}= \left( \begin{array}{cc} 1 &{}\quad 0 \\ 0 &{}\quad -1 \end{array} \right) \) are the standard Pauli matrices. The proof of Theorem 1.4 is based on the study of matrices \(M_{1}\), \(M_{2}\) and \(M_{3}\).

2.2 Useful Identities

From now on and until the end of Sect. 2, we study the matrix-valued function \(M_{z,w}\) with \(w= E \in {\mathbb {R}}\). We start by collecting several important relations between the components of \(M_{z,E}\).

Lemma 2.6

For all \(E\in {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\)

$$\begin{aligned} M_{z,-E} = ( Q^{-}\,M_{z,E} \, Q^{-})^{t} , \end{aligned}$$
(2.33)

where \(Q^{-} = \mathrm {diag} (-1,-1,1,-1,1,1,-1,1) \in {\mathbb {C}}^{8\times 8} \). In particular, for all \(E\in {\mathbb {R}}\), \(z\in {\mathbb {C}}_{+}\) and \(1\le k\le 8\)

$$\begin{aligned} M_{z,E}(k,k) = M_{z,-E}(k,k) . \end{aligned}$$
(2.34)

Proof

Let \(\varvec{H}_{E}^{(\mathrm {sc})} \in {\mathcal {S}}^{8\times 8}\) be given by

$$\begin{aligned} \varvec{H}_{E}^{(\mathrm {sc})} := \left( \begin{array}{ccc|ccc|cc} 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad c_{2}^* \\ 0 &{}\quad 0 &{}\quad \frac{\mathrm {i}\,}{ \gamma } \mathbb {1}_{{\mathcal {S}}}&{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad c_{2}^* \\ 0 &{} \quad -\frac{\mathrm {i}\,}{ \gamma } \mathbb {1}_{{\mathcal {S}}}&{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad c_{2}^* &{}\quad 0 \\ \hline 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad \frac{\mathrm {i}\,}{ \gamma }\mathbb {1}_{{\mathcal {S}}}&{}\quad 0 &{}\quad c_{1}^* \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad -\frac{1}{4 \gamma ^2} \mathbb {1}_{{\mathcal {S}}}&{}\quad 0 &{}\quad c_{1}^* &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 &{} \quad -\frac{\mathrm {i}\,}{ \gamma } \mathbb {1}_{{\mathcal {S}}}&{}\quad 0 &{}\quad 0 &{}\quad c_{1}^* &{} \quad 0 \\ \hline 0 &{}\quad 0 &{} \quad c_{2} &{}\quad 0 &{}\quad c_{1} &{}\quad c_{1} &{} \quad 0 &{}\quad E \mathbb {1}_{{\mathcal {S}}}-s\\ c_{2} &{}\quad c_{2} &{} \quad 0 &{}\quad c_{1} &{}\quad 0 &{}\quad 0 &{}\quad E \mathbb {1}_{{\mathcal {S}}}-s&{}\quad 0 \\ \end{array} \right) , \end{aligned}$$
(2.35)

where \(s\) is a semicircular element, \(c_1,c_2\) are circular elements, all freely independent in a \(C^*\)-probability space \(({\mathcal {S}},\tau _{{\mathcal {S}}})\), so that

$$\begin{aligned} M_{z,E} := ({\mathrm {id}}_{8}\otimes \,\tau _{{\mathcal {S}}}) \big (\varvec{H}_{E}^{(\mathrm {sc})} - zJ_{8}\otimes \mathbb {1}_{{\mathcal {S}}}\big )^{-1} . \end{aligned}$$
(2.36)

Using the fact that \(-s\), \(-c_{1}^*\) and \(-c_{2}^*\) form again a freely independent family of one semicircular and two circular elements, we can easily check that (here \(\times \) denotes multiplication in \({\mathcal {S}}^{8\times 8}\))

$$\begin{aligned} ({\mathrm {id}}_{8}\otimes \,\tau _{{\mathcal {S}}})\Big (\big ((Q^{-}\otimes \mathbb {1}_{{\mathcal {S}}}) \times \big (\varvec{H}_{-E}^{(\mathrm {sc})} - zJ_{8}\otimes \mathbb {1}_{{\mathcal {S}}}\big ) \times (Q^{-}\otimes \mathbb {1}_{{\mathcal {S}}})\big )^{t}\Big )^{-1} = M_{z,E} , \end{aligned}$$
(2.37)

from which (2.33) follows after factorizing \(Q^{-}\). \(\square \)

Lemma 2.7

For all \(E\in {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\)

$$\begin{aligned} M_{z,E}(8,8) = 4 \gamma ^2 z\, M_{z,E}(7,7) . \end{aligned}$$
(2.38)

Proof

Using the Schur complement formula, the lower-right \(2\times 2\) block of the inverse of \(\varvec{H}^{\mathrm {(sc)}}_{E} - zJ_{8} \otimes \mathbb {1}_{{\mathcal {S}}}\) can be written as

$$\begin{aligned} \left( \begin{array}{cc} 4 \gamma ^2 c_{1} c_{1}^* &{}\quad E \mathbb {1}_{{\mathcal {S}}}-s+ \mathrm {i}\,\gamma (c_{1}c_{1}^* + c_{2}c_{2}^*) \\ E \mathbb {1}_{{\mathcal {S}}}-s- \mathrm {i}\,\gamma (c_{1}c_{1}^* + c_{2}c_{2}^*) &{}\quad \frac{1}{z} c_{2}c_{2}^* \end{array} \right) ^{-1} . \end{aligned}$$
(2.39)

For convenience, change the rows in the above matrix, so that

$$\begin{aligned} M_{3} \left( \begin{array}{cc} 0&{}\quad 1 \\ 1&{}\quad 0 \end{array} \right) = ({\mathrm {id}}_2\otimes \,\tau _{{\mathcal {S}}}) \left[ \left( \begin{array}{cc} E \mathbb {1}_{{\mathcal {S}}}-a &{}\quad \frac{1}{z} c_{2}c_{2}^* \\ 4 \gamma ^2 c_{1} c_{1}^* &{}\quad E \mathbb {1}_{{\mathcal {S}}}-a^* \end{array} \right) ^{-1} \right] , \end{aligned}$$
(2.40)

where we introduced

$$\begin{aligned} a := s+\mathrm {i}\,\gamma \big (c_{1}c_{1}^* + c_{2}c_{2}^*\big ) . \end{aligned}$$
(2.41)

Notice, that since \(c_{1}c_{1}^* + c_{2}c_{2}^*\) has a free Poisson distribution of rate 2, \(c_{1}c_{1}^* + c_{2}c_{2}^* \ge (1-\frac{1}{\sqrt{2}})^2 \mathbb {1}_{{\mathcal {S}}}\) and thus both diagonal elements of the matrix on the right-hand side of (2.40) are invertible. Rewrite the matrix in the square brackets in the following way: for the entries of the first row apply the Schur complement formula with respect to the (1, 1)-component, and for the second row apply the Schur complement formula with respect to the (2, 2)-component. This leads to the following expressions for \(M_{z,E}(7,7)\) and \(M_{z,E}(8,8)\)

$$\begin{aligned}&M_{z,E}(7,7) = \frac{1}{z}\tau _{{\mathcal {S}}} \Bigg (-\frac{1}{E \mathbb {1}_{{\mathcal {S}}}-a} c_{2}c_{2}^*\frac{1}{E \mathbb {1}_{{\mathcal {S}}}-a^*-\frac{4 \gamma ^2}{z} c_{1}c_{1}^*\frac{1}{E \mathbb {1}_{{\mathcal {S}}}-a}c_{2}c_{2}^*} \,\Bigg ) , \end{aligned}$$
(2.42)
$$\begin{aligned}&M_{z,E}(8,8) = 4 \gamma ^2 \tau _{{\mathcal {S}}}\Bigg (-\frac{1}{E \mathbb {1}_{{\mathcal {S}}}-a^*} c_{1}c_{1}^*\frac{1}{E \mathbb {1}_{{\mathcal {S}}}-a-\frac{4 \gamma ^2}{z} c_{2}c_{2}^*\frac{1}{E \mathbb {1}_{{\mathcal {S}}}-a^*}c_{1}c_{1}^*} \, \Bigg ) . \end{aligned}$$
(2.43)

Under \(\tau _{{\mathcal {S}}}\), we can swap the labels of \(c_1\) and \(c_2\) and replace a with \(-\,a^*\) without changing the value in (2.43). After completing these operations, we obtain

$$\begin{aligned} M_{z,E}(8,8) = 4 \gamma ^2 \tau _{{\mathcal {S}}}\Bigg (-\frac{1}{E \mathbb {1}_{{\mathcal {S}}}+a} c_{2}c_{2}^*\frac{1}{E \mathbb {1}_{{\mathcal {S}}}+a^*-\frac{4 \gamma ^2}{z} c_{1}c_{1}^*\frac{1}{E \mathbb {1}_{{\mathcal {S}}}+a}c_{2}c_{2}^*} \,\Bigg ) . \end{aligned}$$
(2.44)

Multiplying both fractions under \(\tau _{{\mathcal {S}}}\) in (2.44) by \(-\,1\), and swapping E to \(-\,E\) by (2.34), a comparison with (2.42) yields (2.38). \(\square \)

Lemma 2.8

For all \(E\in {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\)

$$\begin{aligned} M_{z,E}(8,7)-M_{z,E}(7,8) = \frac{\mathrm {i}\,}{ \gamma } M_{z,E}(8,8) \end{aligned}$$
(2.45)

Proof

Denote

$$\begin{aligned} T_1 := \left( \begin{array}{cc} -\frac{1}{ \gamma ^2 z} &{}\quad \frac{\mathrm {i}\,}{ \gamma } \\ -\frac{\mathrm {i}\,}{ \gamma } &{}\quad 0 \end{array} \right) + M_{3} ,\quad T_{2} := \left( \begin{array}{cc} 0 &{}\quad \frac{\mathrm {i}\,}{ \gamma } \\ -\frac{\mathrm {i}\,}{ \gamma } &{}\quad -4 \end{array} \right) + M_{3} , \end{aligned}$$
(2.46)

so that (2.32) can be rewritten as

$$\begin{aligned} \frac{1}{M_{3}}- E \sigma _{1} - \frac{1}{T_{1}} - \frac{1}{T_{2}} + \sigma _{1} M_{3} \sigma _{1} = 0 . \end{aligned}$$
(2.47)

Then from (2.38), we have that

$$\begin{aligned} \det T_{1} - \det T_{2} = 4M_{z,E}(7,7)- \frac{1}{ \gamma ^2 z} M_{z,E}(8,8) = 0 \end{aligned}$$
(2.48)

Rewrite (2.47) componentwise using (2.48)

$$\begin{aligned}&\Big (\frac{1}{\det M_{3}} -\frac{2}{\det T_{1}} + 1 \Big ) M_{z,E}(8,8) = -\frac{4}{\det T_1} , \end{aligned}$$
(2.49)
$$\begin{aligned}&\Big ( -\frac{1}{\det M_{3}}+\frac{2}{\det T_{1}} \Big ) M_{z,E}(7,8) + M_{z,E}(8,7) = E - \frac{2 \mathrm {i}\,}{ \gamma } \frac{1}{\det T_1 } , \end{aligned}$$
(2.50)
$$\begin{aligned}&\Big ( -\frac{1}{\det M_{3}}+\frac{2}{\det T_{1}} \Big ) M_{z,E}(8,7) + M_{z,E}(7,8) = E + \frac{2 \mathrm {i}\,}{ \gamma } \frac{1}{\det T_1 }, \end{aligned}$$
(2.51)
$$\begin{aligned}&\Big (\frac{1}{\det M_{3}} -\frac{2}{\det T_{1}} + 1 \Big ) M_{z,E}(7,7) = -\frac{1}{ \gamma ^2 z\det T_1} . \end{aligned}$$
(2.52)

Subtracting (2.51) from (2.50) gives

$$\begin{aligned} \Big ( \frac{1}{\det M_{3}} - \frac{2}{\det T_{1}} + 1\Big ) (M_{z,E}(8,7)-M_{z,E}(7,8)) = - \frac{ \mathrm {i}\,}{ \gamma } \frac{4}{\det T_1 } , \end{aligned}$$
(2.53)

which together with (2.49) implies (2.45). \(\square \)

2.3 Boundedness of \(M_{z,E}\) Away from \(z=0\) and \(z=1\)

Lemma 2.9

(Boundedness of \( M_{z,E} \)) For any small \(\theta >0\), there exists \(C_{\theta }>0\) such that

$$\begin{aligned} \sup \bigg \{\,\Vert M_{z,E}\Vert \, : \, | z |\ge \theta , |1- z | \ge \theta , \, \mathrm {Im}z > 0,\, |E|\le \frac{1}{\theta } \,\bigg \} \le C_{\theta }. \end{aligned}$$
(2.54)

Proof

Introduce the following notation for the entries of \(M_{3}\)

$$\begin{aligned} \left( \begin{array}{cc} m_{11}&{}\quad m_{12} \\ m_{21}&{}\quad m_{22} \end{array} \right) := \left( \begin{array}{cc} M_{z,E}(7,7) &{}\quad M_{z,E}(7,8) \\ M_{z,E}(8,7) &{}\quad M_{z,E}(8,8) \end{array} \right) , \end{aligned}$$
(2.55)

so that, in particular, (2.38) and (2.45) can be rewritten as

$$\begin{aligned} m_{22}&= 4 \gamma ^2 z m_{11} , \end{aligned}$$
(2.56)
$$\begin{aligned} m_{21} - m_{12}&= \frac{\mathrm {i}\,}{ \gamma } m_{22} . \end{aligned}$$
(2.57)

Our goal is to show that \(M_{z,E}\) and in particular \(M_{z,E}(1,1)\) is bounded everywhere if z is away from 0 or 1. From (2.31), we have that

$$\begin{aligned}&M_{1} = -\left( \begin{array}{ccc} m_{22} + z &{}\; m_{22} &{}\; m_{21} \\ m_{22} &{}\; m_{22} &{}\; m_{21} - \frac{\mathrm {i}\,}{ \gamma } \\ m_{12} &{}\; m_{12} + \frac{\mathrm {i}\,}{ \gamma } &{}\; m_{11} \end{array} \right) ^{-1} ,\nonumber \\&M_{2} = -\left( \begin{array}{ccc} m_{22} &{}\; m_{21} &{} \; m_{21} - \frac{\mathrm {i}\,}{ \gamma } \\ m_{12} &{}\; m_{11} + \frac{1}{4 \gamma ^2} &{}\; m_{11} \\ m_{12} + \frac{\mathrm {i}\,}{ \gamma } &{}\; m_{11} &{}\; m_{11} \end{array} \right) ^{-1} , \end{aligned}$$
(2.58)

which after some elementary computations yields

$$\begin{aligned} \det M_{1} = \frac{-1}{z \det T_1} ,\quad \det M_{2} = \frac{- \gamma ^2}{\det T_1} \end{aligned}$$
(2.59)

and

$$\begin{aligned} M_{z,E}(1,1) = -\frac{1}{z} - \frac{4m_{11}}{z\det T_{1}} , \end{aligned}$$
(2.60)

where \(T_{1}\) was defined in (2.46). The functions \(\{m_{ij}, 1\le i,j \le 2\}\), \(\{M_{i}, 1\le i \le 3\}\) and \(\{T_{i}, 1\le i \le 2\}\) defined above all depend on the variables z and E, but in order to make the exposition lighter we drop the explicit dependence on these variables from the notation. Using (2.58)–(2.60) it is enough to show that for any fixed \((z_{\infty },E_{\infty })\) with \(z_{ \infty } \in \overline{{\mathbb {C}}_{+}}\), \(z_{\infty } \notin \{0,1\}\) and \(E_{\infty } \in {\mathbb {R}}\) we have that \(\lim _{(z,E)\rightarrow (z_{\infty }, E_{\infty })} |m_{ ij }| <\infty \) for \(i,j\in \{1,2\}\) and \(\lim _{(z,E)\rightarrow (z_{\infty }, E_{\infty })}|\det T_{1}|>0\).

We now prove some additional relations that can be obtained from (2.56), (2.57) and (2.49)–(2.52). Plugging (2.56) and (2.57) into (2.50) (recall that we are using notation (2.55)) gives

$$\begin{aligned} \Big (-\frac{1}{\det M_{3}}+ \frac{2}{\det T_{1} }+1\Big ) m_{12} + 4 \gamma \mathrm {i}\,z m_{11} + \frac{2\mathrm {i}\,}{ \gamma \det T_{1}} - E = 0 , \end{aligned}$$
(2.61)

which, after applying (2.52) to the terms in the parenthesis, can be rewritten as

$$\begin{aligned} \Big ( \frac{1}{ \gamma ^2 z m_{11} \det T_{1}} +2\Big ) m_{12} = E - \frac{2 \mathrm {i}\,}{ \gamma \det T_{1}} - 4 \gamma \mathrm {i}\,z m_{11} . \end{aligned}$$
(2.62)

From the definitions of \(T_{1}\) and \(M_{3}\) we have

$$\begin{aligned} \det T_{1} = \det M_{3} + 4 (z-1) m_{11} - \frac{1}{ \gamma ^2} , \end{aligned}$$
(2.63)

while (2.52) gives

$$\begin{aligned} \frac{1}{\det T_{1}} \Big ( 2- \frac{1}{ \gamma ^2 z m_{11}} \Big ) = \frac{1}{\det M_{3}} + 1 . \end{aligned}$$
(2.64)

Combining (2.63) and (2.64), we get the following quadratic equation for \(\det M_{3}\)

$$\begin{aligned}&\big (\det M_{3}\big )^2 + \det M_{3}\bigg ( 4(z-1) m_{11} + \frac{1}{ \gamma ^2 z m_{11}} -\bigg (1+\frac{1}{ \gamma ^2}\bigg )\bigg )\nonumber \\&\quad + \bigg ( 4(z-1)m_{11} - \frac{1}{ \gamma ^2}\bigg ) = 0 . \end{aligned}$$
(2.65)

Note that (2.62)–(2.65) hold for all \(E\in {\mathbb {R}}\) and all \(z\in {\mathbb {C}}_{+}\).

Using the above relations, we proceed with a proof by contradiction. Assume that there exists a sequence \((z_{n},E_{n})_{n=1}^{\infty }\subset {\mathbb {C}}_{+}\times [-\theta ^{-1},\theta ^{-1}]\), such that \(|m_{11}^{(n)}| \rightarrow \infty \) as \(n\rightarrow \infty \) (here and below we denote the evaluations at \((z_{n},E_{n})\) by adding the superscript (n)). Solving (2.65) for \(\det M_{3}\) allows us to express \(\det M_{3}\) in terms of \(m_{11}\)

$$\begin{aligned} \det M_{3}= & {} \frac{1}{2} \bigg \{-4(z-1)m_{11}+\Big (1+\frac{1}{ \gamma ^2}\Big ) \nonumber \\&{\pm } 4(z-1)m_{11} \bigg [1 -\frac{1}{2}\bigg (\frac{1}{2(z-1)}\Big (1+\frac{1}{ \gamma ^2}\Big )+\frac{1}{z-1}\bigg )\frac{1}{m_{11}}\nonumber \\&+O\left( \frac{1}{|m_{11}|^2}\right) \bigg ] \bigg \} . \end{aligned}$$
(2.66)

By passing to a subsequence, we may assume that the choice of the \({\pm }\) sign in (2.66) is constant for all n.

If we take the − sign in (2.66), then

$$\begin{aligned} \det M_{3}^{(n)} = -4(z_{n}-1)m_{11}^{(n)} + O\left( 1\right) \end{aligned}$$
(2.67)

and by (2.64)

$$\begin{aligned} \det T_{1}^{(n)} \rightarrow 2 ,\quad n\rightarrow \infty . \end{aligned}$$
(2.68)

From (2.62), (2.56), (2.57) and (2.68),

$$\begin{aligned} m_{12}^{(n)} = -2 \gamma \mathrm {i}\,z_{n} m_{11}^{(n)} + O\left( 1\right) ,\quad m_{21}^{(n)} = 2 \gamma \mathrm {i}\,z_{n} m_{11}^{(n)} + O\left( 1\right) , \end{aligned}$$
(2.69)

and therefore

$$\begin{aligned} \det M_{3}^{(n)} = 4 \gamma ^2 z_{n} \left( m_{11}^{(n)}\right) ^2-4 \gamma ^2z_{n}^2 \left( m_{11}^{(n)}\right) ^2 = 4 \gamma ^2 z_{n}(1-z_{n})\left( m_{11}^{(n)}\right) ^2 {+} O\left( m_{11}^{(n)}\right) , \end{aligned}$$
(2.70)

which contradicts to (2.67) since \(|z_{n}(1-z_{n})|\) is separated away from 0.

If we take \(+\) sign in (2.66), then

$$\begin{aligned} \det M_{3}^{(n)} = -1 +O\left( \frac{1}{\big |m_{11}^{(n)}\big |}\right) , \end{aligned}$$
(2.71)

and from (2.63)

$$\begin{aligned} \det T_{1}^{(n)} = 4(z_{n}-1)m_{11}^{(n)} + O\left( 1\right) . \end{aligned}$$
(2.72)

But then again, from (2.62), (2.56), (2.57) and (2.72),

$$\begin{aligned} m_{12}^{(n)} = -2 \gamma \mathrm {i}\,z_{n} m_{11}^{(n)} + O\left( 1\right) ,\quad m_{21}^{(n)} = 2 \gamma \mathrm {i}\,z_{n} m_{11}^{(n)} + O\left( 1\right) , \end{aligned}$$
(2.73)

so that

$$\begin{aligned} \det M_{3}^{(n)} = 4 \gamma ^2 z_{n}(1-z_{n}) (m_{11}^{(n)})^2 + O\left( \big |m_{11}^{(n)}\big |\right) , \end{aligned}$$
(2.74)

which contradicts to (2.71). Therefore, we have proven that \(m_{11}\) is bounded everywhere away from the points \(z\in \{0,1\}\). It is clear from (2.56) that the boundedness of \(m_{11}\) guarantees the boundedness of \(m_{22}\). On the other hand, assuming that \(m_{12}\) and \(m_{21}\) [see (2.57)] are unbounded implies that \(|\det M_{3}^{(n)}|\rightarrow \infty \) and \(|\det T_1^{(n)}|\rightarrow \infty \) on some sequence \((z_n,E_n)_{n=1}^{\infty }\), which contradicts to the boundedness of \(m_{11}\) in (2.61). We conclude that all entries of \(M_{3}\) are bounded everywhere away from the points \(z\in \{0,1\}\).

Assume now that there exists a sequence \((z_{n},E_{n})_{n=1}^{\infty }\subset {\mathbb {C}}_{+}\times [-\theta ^{-1},\theta ^{-1}]\) such that \(\det T_{1}^{(n)} \rightarrow 0\) as \(n\rightarrow \infty \). Then by (2.62)

$$\begin{aligned} \left( \frac{1}{ \gamma ^2 z_{n} m_{11}^{(n)} }+O\left( \det T_{1}^{(n)}\right) \right) m_{12}^{(n)} = -\frac{2\mathrm {i}\,}{ \gamma } + O\left( \det {T_{1}}^{(n)}\right) , \end{aligned}$$
(2.75)

which together with the fact that \(m_{11}^{(n)}\) is bounded implies that

$$\begin{aligned} m_{12}^{(n)} = -2\mathrm {i}\,\gamma z_{n} m_{11}^{(n)} + O\left( \det T_{1}^{(n)}\right) . \end{aligned}$$
(2.76)

Then by (2.56) and (2.57) we find

$$\begin{aligned} \det M_{3}^{(n)} = 4 \gamma ^2 z_{n}(1-z_{n}) \big (m_{11}^{(n)}\big )^2 + O\left( \det T_{1}^{(n)}\right) , \end{aligned}$$
(2.77)

and conclude from (2.77) and (2.63) that \(\det M_{3}^{(n)}\) does not vanish as \(n\rightarrow \infty \). But then (2.64) implies

$$\begin{aligned} m_{11}^{(n)} = \frac{1}{2 \gamma ^2 z_{n} } + O\left( \det T_{1}^{(n)}\right) , \end{aligned}$$
(2.78)

and thus by (2.63) and (2.78) also

$$\begin{aligned} \det M_{3}^{(n)} = \frac{1}{ \gamma ^2} - 2(z_{n}-1) \frac{1}{ \gamma ^2 z_{n} } + O\left( \det T_{1}^{(n)}\right) , \end{aligned}$$
(2.79)

as well as

$$\begin{aligned} \det M_{3}^{(n)} = 4 \gamma ^2 z_{n}(1-z_{n}) \frac{1}{4 \gamma ^4 z_{n}^2 } + O\left( \det T_{1}^{(n)}\right) = (1-z_{n}) \frac{1}{ \gamma ^2 z_{n} } + O\left( \det T_{1}^{(n)}\right) , \end{aligned}$$
(2.80)

by (2.77) and (2.78), which contradicts to (2.79) since \(z_{n}\) is away from 0. We conclude that \(|\det T_{1}|\) is bounded away from 0, which together with (2.60) establishes (a non-effective version of) (2.54).

In order to keep the presentation simple, above we showed that \(\Vert M_{z,E}\Vert \) is bounded for \(z\in \overline{{\mathbb {C}}_{+}} {\setminus } \{0,1\}\) and \(E\in {\mathbb {R}}\) without providing an explicit effective bound \(C_{\theta }\) as formulated in (2.54). Note that the constants hidden in the \(O\left( \cdot \right) \) terms (for example, in (2.67)–(2.70), (2.71)–(2.74) or (2.75)–(2.80)) depend only on \(|z_n|\), \(|1-z_n|\) and \(E_n\). Therefore, using the assumptions \(|m_{11}^{(n)}| \ge D_{\theta }\) and \(|\det T_{1}^{(n)}|\le 1/D_{\theta }\) for n large enough and carefully chosen \(D_{\theta }>0\) instead of \(|m_{11}^{(n)}| \rightarrow \infty \) and \(|\det T_{1}^{(n)}|\rightarrow 0\) as \(n\rightarrow \infty \), together with the a priori bound (2.16), eventually leads to the uniform bound (2.54). \(\square \)

2.4 Singularities of \(M_{z,E}(1,1)\)

Lemma 2.10

(Behavior of \(M_{z,E}(1,1)\) in the vicinities of \(z=0\) and \(z=1\)

  1. (a)

    For all \(E\in {\mathbb {R}}\)

    $$\begin{aligned} M_{z,E}(1,1) = \root 3 \of {\frac{1+E^2}{4 \gamma ^{2}}}\,z^{-2/3} + O\left( \frac{1}{|z|^{1/3}}\right) \text{ as } z\rightarrow 0 ; \end{aligned}$$
    (2.81)
  2. (b)

    for all \(|E| < \frac{1}{ \gamma } \sqrt{2 \sqrt{1 + 6 \gamma ^2 + \gamma ^4}-2-2 \gamma ^2}\)

    $$\begin{aligned} M_{z,E}(1,1) = - \frac{16 \gamma ^2 \xi _{0}}{ \gamma ^2 E^2 + 4+16 \gamma ^4 \xi _{0}^{2}}\, (z-1)^{-1/2} + O\left( 1\right) \text{ as } z\rightarrow 1 , \end{aligned}$$
    (2.82)

    where the constant \(\xi _{0}<0\) is as in part (ii) of Theorem 1.4 (see also (2.95)).

The choices of the branches for \((\cdot )^{1/3}\) and \((\cdot )^{1/2}\) are specified in the course of the proof below.

Proof

We multiply (2.32) from the left by \(M_{3}\) and from the right by \((Z_1 + M_{3})(Z_1-Z_2)^{-1}(Z_2+M_{3})\) with the short hand notation

$$\begin{aligned} Z_1 := -\frac{1}{2 \gamma ^2 z}(I_2+\sigma _3) -\frac{1}{ \gamma }\sigma _2 , \qquad Z_2 := -2(I_2-\sigma _3) -\frac{1}{ \gamma }\sigma _2 . \end{aligned}$$
(2.83)

Subsequently, the equation for \(M_{3}\) takes the form \(\Delta =0\) with

$$\begin{aligned} \Delta := \Big (Z_1-M_{3} + M_{3} \sigma _1(M_{3}\sigma _1-E)(Z_1+M_{3})\Big )\frac{1}{Z_1-Z_2}(Z_2+M_{3}) -M_{3} . \end{aligned}$$
(2.84)

Using that \((Z_1-Z_2)^{-1} = - \frac{ \gamma ^2 z}{2}(I_2+\sigma _3) +\frac{1}{8}(I_2-\sigma _3)\) and performing the matrix products we see that the entries of \(\Delta \in {\mathbb {C}}^{2 \times 2}\) are polynomials in the entries of \(M_{3}\), E and z.

Step 1: Expansion around \(z=0\). We will now first construct a solution to (2.32) in a vicinity of \(z=0\) by asymptotic expansion. Later, in Step 3, we will show that the constructed solution coincides with \(M_{3}\) defined in (2.15) and (2.30). For this purpose, we write \(t = z^{1/3}\) with an analytic cubic root on \({\mathbb {C}}{\setminus } (\mathrm {i}\,(-\infty ,0]))\) such that \((-1)^{1/3} =-1\). Then, we make the ansatz

$$\begin{aligned} {\widetilde{M}}_3 = \left( \begin{array}{cc} \xi _{11} t^{-1}&{}\quad 4 \gamma ^2 \xi _{12} t \\ 4 \gamma ^2 \xi _{21} t &{}\quad 4 \gamma ^2 \xi _{22} t^2 \end{array} \right) , \end{aligned}$$
(2.85)

where we will determine the unknown functions \(\Xi _t:=(\xi _{ij}(t))_{i,j=1}^2\). Plugging (2.85) into (2.84) reveals that

$$\begin{aligned} \Delta = \left( \begin{array}{cc} 4 \gamma ^2 ( q_{11} + t p_{11}) &{}\quad 4 \gamma ^2t(q_{12} + t p_{12}) \\ -4 \gamma ^2t(q_{21} + t p_{21}) &{}\quad -16 \gamma ^4 t^3 (q_{22} + t p_{22}) \end{array} \right) , \end{aligned}$$
(2.86)

where the entries of \(P:=(p_{ij})_{i,j=1}^2\) are polynomials in \(t,E,\xi _{ij}\) and where \(Q:=(q_{ij})_{i,j=1}^2\) is given by

$$\begin{aligned} Q = \left( \begin{array}{cc} \frac{1}{16 \gamma ^4} + \xi _{11}^2 \xi _{22} - E \xi _{11} \xi _{12} &{}\quad \xi _{12} +E \xi _{11} \xi _{22} \\ \xi _{21} + E \xi _{11} \xi _{22} &{}\quad -\frac{1}{16 \gamma ^4}- \xi _{11} \xi _{22}^2+E \xi _{21} \xi _{22} \end{array} \right) . \end{aligned}$$
(2.87)

Thus, the equation \(\Delta =0\) is equivalent to \(Q+t P =0\). For \(t=0\), the three possible solutions are

$$\begin{aligned} \begin{aligned} \Xi _0 = \left( \begin{array}{cc} \zeta &{}\quad -E \zeta ^2 \\ -E \zeta ^2&{}\quad \zeta \end{array} \right) , \quad \text {where} \quad (1+E^2) \zeta ^3 = -\frac{1}{16 \gamma ^4}. \end{aligned} \end{aligned}$$
(2.88)

Since \(\det \nabla _{\Xi } Q|_{t=0} = -3 \zeta ^4(1+E^2)^2\) the equation \(Q+tP = 0\) is linearly stable at \(t=0\) and can thus be solved for \(\Xi _t\) in a neighborhood of \(\Xi _0\) when t is sufficiently small. Since the equation is polynomial, the solution \(\Xi _t\) admits a power series expansion in t. Now we define the analytic function \({\widetilde{M}}_3(z)\) through (2.85) on a neighborhood of \(z=0\) in \({\mathbb {C}}{\setminus } (\mathrm {i}\,(-\infty ,0]))\) with \(\Xi \) the solution to \(Q+tP = 0\) and the choice \(\zeta <0\) in (2.88).

We will now check that for any phase \(e^{\mathrm {i}\,\psi } \ne -1\) in the complex upper half-plane the imaginary part of \({\widetilde{M}}_3(\theta e^{\mathrm {i}\,\psi })\) is positive definite in the sense that for any fixed \(\varepsilon >0\) we have

$$\begin{aligned} \inf _{\psi \in [-\pi +\varepsilon ,0]}\inf _{\Vert u \Vert _{2}=1}\mathrm {Im}\langle u, {\widetilde{M}}_3(\theta e^{\mathrm {i}\,\psi }) u \rangle > 0 , \end{aligned}$$
(2.89)

for sufficiently small \(\theta >0\). Indeed, for any vector \(u=(u_1,u_2) \in {\mathbb {C}}^2\) we find

$$\begin{aligned} \mathrm {Im}\langle u, {\widetilde{M}}_3 u \rangle\ge & {} \beta \mathrm {Im}(-t^{-1} |u_1|^2 - t^2 |u_2|^2 ) - C |u_1||u_2| |t| \nonumber \\\ge & {} \beta _\varepsilon \bigg (\frac{1}{|t|} |u_1|^2 + |t|^2 |u_2|^2\bigg ) - \frac{C}{R}\frac{1}{|t|}|u_1|^2 - C R|u_2| |t|^3\qquad \end{aligned}$$
(2.90)

for some constants \(\beta ,C>0\), small t, an \(\varepsilon \)-dependent constant \(\beta _\varepsilon \) and any \(R>0\). For \(R=2C/\beta _\varepsilon \) and sufficiently small t this is still positive.

Step 2: Expansion around \(z=1\). For the expansion around \(z=1\), we proceed similarly to the discussion at \(z=0\). We set \(t = \sqrt{z-1}\), where the square root has a branch cut at \(\mathrm {i}\,(-\infty ,0]\) and \(\sqrt{1}=1\). Here, we make an ansatz with a reduced number of unknown functions by exploiting the identities (2.38) and (2.45), namely

$$\begin{aligned} {\widetilde{M}}_3 = \left( \begin{array}{cc} \frac{\xi }{4 \gamma ^2 t} &{}\quad \frac{E}{2}- \frac{\mathrm {i}\,\xi }{2 \gamma } (t^{-1}+t) + \frac{\nu t}{4 \gamma } \\ \frac{E}{2} + \frac{\mathrm {i}\,\xi }{2 \gamma } (t^{-1}+t) + \frac{\nu t}{4 \gamma } &{}\quad \xi (t^{-1}+t) \end{array} \right) . \end{aligned}$$
(2.91)

We will determine the unknown functions \(\xi \) and \(\nu \). Plugging (2.91) into (2.84), multiplying out everything and simplifying afterwards reveals

$$\begin{aligned} \Delta = \left( \begin{array}{cc} -\frac{1}{64 \gamma ^4 }(q_1+t p_1) &{}\quad -\frac{1}{32 \gamma ^3} (q_2 + t p_2) \\ \frac{1}{32 \gamma ^3} (q_2 + t p_2) &{}\quad \frac{1+t^2}{16 \gamma ^2}(q_1+t p_1 -2 \mathrm {i}\,(q_2 + t p_2)) \end{array} \right) , \end{aligned}$$
(2.92)

where \(p_1,p_2\) are polynomials in \(t,E,\xi ,\nu \) and where

$$\begin{aligned} \begin{aligned} q_1 =\,&\xi ^4 + (2 E^2 \gamma ^2 + 4 + 4 \gamma ^2) \xi ^2 + \gamma ^2( \gamma ^2 E^4 + 4 E^2 (1 + \gamma ^2)- 16 ) \\&-\, 16 \mathrm {i}\,\gamma ^3 E + \mathrm {i}\,E^2 \gamma ^2 \xi \nu + 4 \mathrm {i}\,\xi \nu + \mathrm {i}\,\xi ^3 \nu , \\ q_2 = \,&\xi ^3 \nu + (4 + \gamma ^2 E^{2}) \xi \nu - 16 \gamma ^{3} E \, . \end{aligned} \end{aligned}$$
(2.93)

The solution to the system \((q_1,q_2)=0\) at \(t=0\) has the form

$$\begin{aligned} \nu _0= & {} \frac{16 \gamma ^3 E}{\xi (4+ \gamma ^2 E^2+ \xi ^2 )}, \quad r:= \gamma ^2E^4 + \frac{1}{ \gamma ^2} \xi ^4 \nonumber \\&+ \frac{4}{ \gamma ^2} (1 + \gamma ^2) \xi ^2 + 4 E^2 \left( 1 + \gamma ^2 + \frac{1}{2} \xi ^2 \right) -16 =0 \end{aligned}$$
(2.94)

In \(r=0\) from (2.94) we choose the unique negative solution (as long as the expression inside the square root is positive)

$$\begin{aligned} \xi _0 = -\sqrt{2 \sqrt{1 + 6 \gamma ^2 + \gamma ^4}-2-2 \gamma ^2 - \gamma ^2 E^2} . \end{aligned}$$
(2.95)

We compute the Jacobian

$$\begin{aligned} \begin{aligned} \det \nabla _{\xi ,\nu }(q_1,q_2) = 4 \xi ^2 \left( 8 + 2 \gamma ^2 (12 + E^2) + \gamma ^4\left( \frac{2}{ \gamma ^4} \xi ^2 - 2 E^2\right) - 2 \gamma ^2 \xi ^2 + \gamma ^2 r\right) , \end{aligned} \end{aligned}$$
(2.96)

and evaluate at \(t=0\) with \(\xi =\xi _0\) to find

$$\begin{aligned} \det \nabla _{\xi _0,\nu _0}(q_1,q_2) = 16 \xi _0^2 \left( 1 {+} \gamma ^4 {+} \sqrt{ 1 {+} 6 \gamma ^2 {+} \gamma ^4} - \gamma ^2 \left( -6 {+} \sqrt{ 1 {+} 6 \gamma ^2 {+} \gamma ^4}\right) \right) <0. \end{aligned}$$
(2.97)

In particular, \(\xi \) and \(\nu \) admit a power series expansion in t.

Step 3: Coincidence of \({\widetilde{M}}_3\) and \(M_{3}\), asymptotic behavior of \(M_{z,E}(1,1)\). Here, we show that the asymptotic expansion \({\widetilde{M}}_3\) around \(z=0,1\) coincides with \(M_{3}\) defined in (2.15) and (2.30), the solution to DEL (2.32) constructed from free probability. For this purpose, for any small \(\delta >0\), we construct a modification \(M_{z,E}^{\delta }\) of the explicit solution \(M_{z,E}\) (given by (2.36)) as follows

$$\begin{aligned} M^{\delta }_{z,E} := ({\mathrm {id}}_{8} \otimes \,\tau _{{\mathcal {S}}})\big (\varvec{H}_E^{(\mathrm {sc}),\delta }- z J_{8} \otimes \mathbb {1}_{{\mathcal {S}}}\big )^{-1} , \end{aligned}$$
(2.98)

with

$$\begin{aligned} \qquad \varvec{H}_E^{(\mathrm {sc}),\delta } := \left( \begin{array}{ccc} \kappa _1 \otimes \mathbb {1}_{{\mathcal {S}}}&{}\quad 0 &{}\quad \kappa _5^t \otimes c_2^* \\ 0&{}\quad \kappa _2 \otimes \mathbb {1}_{{\mathcal {S}}}&{}\quad \kappa _4^t \otimes c_1^* \\ \kappa _5 \otimes c_2 &{}\quad \kappa _4 \otimes c_1 &{}\quad \sigma _1 \otimes (E \cdot \mathbb {1}_{{\mathcal {S}}}-(1-\delta )\cdot s) -\delta \sigma _1 {\widetilde{M}}_3 \sigma _1\otimes \mathbb {1}_{{\mathcal {S}}}\end{array} \right) , \end{aligned}$$
(2.99)

where \(c_i\) are circular and \(s\) semicircular elements. Since the original generalized resolvent has the bound \(\Vert M_{z,E}\Vert \le C (1+\frac{1}{\mathrm {Im}z})\), we also get

$$\begin{aligned} \Vert M_{z,E}^{\delta }-M_{z,E}\Vert \le C_{E,z}\delta . \end{aligned}$$
(2.100)

Now, similarly as in (2.31)–(2.32), we derive the DEL corresponding to the generalized resolvent of \(\varvec{H}_E^{(\mathrm {sc}),\delta }\) and find

$$\begin{aligned}&-\frac{1}{M_{1}^{\delta }} = z J_{3} - \kappa _{1} + (1-\delta ) \kappa _{5}^{t} M_{3}^{\delta } \kappa _{5} ,\quad -\frac{1}{M_{2}^{\delta }} = - \kappa _{2} + (1-\delta ) \kappa _{4}^{t} M_{3}^{\delta } \kappa _{4} , \end{aligned}$$
(2.101)
$$\begin{aligned}&-\frac{1}{M_{3}^{\delta }} = -E \sigma _1 - \frac{1}{Z_1 +M_{3}^{\delta }} - \frac{1}{Z_2 +M_{3}^{\delta }} + (1-\delta ) \sigma _1 M_{3}^{\delta }\sigma _1 + \delta \sigma _1 {\widetilde{M}}_{3}\sigma _1 , \end{aligned}$$
(2.102)

where we exploited the block structure of \(M_{z,E}^{\delta }\) and denoted

$$\begin{aligned} M_{z,E}^{\delta } = \left( \begin{array}{ccc} M_{1}^{\delta }&{}\quad 0 &{}\quad 0 \\ 0 &{}\quad M_{2}^{\delta } &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad M_{3}^{\delta } \end{array} \right) . \end{aligned}$$
(2.103)

Notice that \(\mathrm {Im}\big (\delta \sigma _1 {\widetilde{M}}_{3}\sigma _1\big )>0\) and \(R \mapsto -E \sigma _1 - (Z_1 +R)^{-1} - (Z_2 +R)^{-1} + (1-\delta ) \sigma _1 R \sigma _1\) is a positivity preserving analytic mapping. The existence and uniqueness of the solution to (2.102) now follow from [25, Theorem 2.1]. Inverting both sides of (2.102), we obtain a fixed point equation \(M_{3}^{\delta }= \Phi _{\delta }(M_{3}^{\delta })\) for \(M_{3}^{\delta }\), where the map \(\Phi _{\delta }\) can be read off from the inverse of the right-hand side of (2.102). Clearly, \(\Phi _{\delta }\) with any \(\delta >0\) is a contraction with respect to the Carathéodory metric mapping the set of \(2 \times 2\) matrices with (strictly) positive definite imaginary parts strictly into itself (for more details, see [25]). In particular, there is a unique solution with positive semidefinite imaginary part and thus \(M_{3}^{\delta } = {\widetilde{M}}_{3}\) for \(\delta >0\). Together with (2.100) and taking the limit \(\delta \downarrow 0\), we conclude \(M_{3} = \widetilde{M_{3}}\). In particular, \(M_{3}\) has a power law expansion around the singularities at \(z=0,1\).

Finally, plugging the expansions (2.85) and (2.91) into (2.60) yields the asymptotics (2.81)–(2.82). \(\square \)

2.5 Proof of (i) and (ii) of Theorem 1.4 for \(\phi = 1\)

We will need the following classical result from the theory of Herglotz functions (see, e.g., [22, Theorem 2.3])

Lemma 2.11

Let \(m:{\mathbb {C}}_{+}\rightarrow {\mathbb {C}}_{+}\) be a Herglotz function with representation

$$\begin{aligned} m(z) = m^{\infty } + \int _{{\mathbb {R}}} \frac{1}{\lambda - z} v (d\lambda ) \end{aligned}$$
(2.104)

for \(m^{\infty }\in {\mathbb {R}}\) and a Borel measure \(v(d\lambda )\) on \({\mathbb {R}}\). If for some \(p>1\) and interval \(I\subset {\mathbb {R}}\)

$$\begin{aligned} \sup _{0< \eta< 1}\int _{I}|\mathrm {Im}m(\lambda + \mathrm {i}\,\eta )|^{p} d\lambda < \infty , \end{aligned}$$
(2.105)

then the measure \(v(d\lambda )\) is absolutely continuous on I.

Proof of (i) and (ii) of Theorem 1.4for \(\phi = 1\). The weak limits in the part (i) of Theorem 1.4 have been established in Lemma 2.5.

We know from (2.17) that \(M_{z,E}(1,1)\) is a Herglotz function admitting the representation (2.104) with \(m^{\infty } = M_{E}^{\infty }(1,1) \) and \(v(d\lambda ) = \langle e_1, V_{E}(d\lambda ) \, e_1\rangle = \rho _{E}(d\lambda )\). Therefore, the absolute continuity of \(\rho _{E}(d\lambda ) = \rho _{E}(\lambda ) d\lambda \) is a direct consequence of Lemma 2.11 and Lemmas 2.9 and 2.10 establishing together the integrability of \(\lambda \mapsto \mathrm {Im}M_{\lambda + \mathrm {i}\,\eta ,E}(1,1)\) in the form (2.105) on the whole real axis for any \(p<3/2\).

Finally, the asymptotic behavior of \(\rho _{E}(\lambda )\) (1.13)–(1.14) near its singularities at 0 and 1 follows from Lemma 2.10 and the inverse Stieltjes transform formula (2.27). This, together with the global law established in Lemma 2.5, finishes the proof of Theorem 1.4 for \(\phi = 1\) . \(\square \)

3 Proof of Theorem 1.4 for General Rational \(\phi \in (0,1)\)

In this section, we explain how the techniques described in Sect. 2 can be used to prove Theorems 1.3 and 1.4 in the case when \(\phi = k/l\in (0,1)\) for fixed \(k,l\in {\mathbb {N}}\), i.e., when \(M=ln\) and \(N=kn\) for some integer n tending to infinity. In addition to the steps used in Sect. 2, we need to tensorize the setup to accommodate for rectangular matrices. For example, the \(M\times N = ln\times kn\) matrices \(W_1\) and \(W_2\) will be viewed as \(l\times k\) rectangular block matrices with blocks of dimension \(n\times n\). As we will see later, this allows us to treat the linearization matrix as a Kronecker random matrix in independent Wigner and i.i.d. matrices, which in turn makes various probabilistic estimates of the error terms in the corresponding DEL readily accessible from [3]. The restriction \(\phi = k/l\) makes the presentation conceptually easier.

Note that different values of \(\phi = k/l \in (0,1)\) require slightly different approaches. For \(\phi \in (1/2,1)\), the matrix \(\varvec{T}_{E,\phi } \in {\mathbb {C}}^{kn \times kn}\) given by (1.10) is well defined and bounded with very high probability. Indeed, in this case the product \(W W^*\) is a sample covariance matrix with concentration ratio \(\frac{1}{2\phi } \in (0,1)\), therefore similarly as in Sect. 2.1, one can define a sequence of events \(\Theta _{\phi ,N}\) holding a.w.o.p. [see (2.4) and (3.2)] such that for any small \(\delta >0\) and big enough \(n\ge n_0( \delta )\), the spectrum of \(W W^{*}\) is contained inside the interval \([(1-\frac{1}{\sqrt{2 \phi }})^{2}(1- \delta ) ,(1+\frac{1}{\sqrt{2 \phi }})^{2}(1+ \delta ) ]\) when restricted to \(\Theta _{\phi , N}\) (see, e.g., [5, Section 5]). Thus, as n tends to infinity, the matrices in the denominator of (1.10) have positive imaginary parts and therefore bounded inverses with very high probability.

On the other hand, for \(\phi \in {(}0,1/2{]}\) we need to proceed via regularization by replacing \(\mathrm {i}\,\gamma W W^*\) with \(\mathrm {i}\,\gamma W W^{*} + \mathrm {i}\,\epsilon \phi I_{M}\) for some small \(\epsilon > 0\) to ensure the invertibility of the denominators in (1.10). This requires a more careful analysis of the \(\epsilon \)-dependence of various bounds and identities before we can take \(\epsilon \rightarrow 0\).

Note also that for \(\phi > 1\), \(\mathrm {Rank}(\varvec{T}_{E,\phi })\le M\) and the spectral measure of the matrix \(\varvec{T}_{E,\phi }\) has an atom of mass \(1-1/2\phi \) at zero, while, as we will show, for \(\phi <1\) the spectral measure \(\mu _{\varvec{T}_{E,\phi }}\) does not have the pure point component. The regime \(\phi =1\) studied in Sect. 2 is borderline: the limiting spectral measure of \(\varvec{T}_{E,\phi }\) does not have atom at 0, but its behavior near the origin is more singular than in the case \(\phi <1\).

3.1 Linearization Trick and the Dyson Equation for Linearization

In order to apply the linearization trick for \(\phi = k/l\in (0,1)\), we split H, \(W_{1}\) and \(W_{2}\) into blocks of size \(n\times n\), so that

$$\begin{aligned} H = ({\widehat{H}}_{ij})_{\begin{array}{c} i=1 \ldots l \\ j=1 \ldots l \end{array}} ,\quad W_{1} = ({\widehat{W}}_{1,ij})_{\begin{array}{c} i=1 \ldots l \\ j=1 \ldots k \end{array}} ,\quad W_{2} = ({\widehat{W}}_{2,ij})_{\begin{array}{c} i=1 \ldots l \\ j=1 \ldots k \end{array}} , \end{aligned}$$
(3.1)

with \({\widehat{H}}_{ij}, {\widehat{W}}_{1,ij}, {\widehat{W}}_{2,ij} \in {\mathbb {C}}^{n\times n}\). Note that for any \(i\in \{1,\ldots ,l\}\), \(\sqrt{l}\cdot {\widehat{H}}_{ii}\) is a (normalized) Wigner matrix of size n, and for any \(i\ne j\), \(\sqrt{l} \cdot {\widehat{H}}_{ij}\) is a (normalized) i.i.d. matrix of size n. Similarly, all the matrices \(\sqrt{l} \cdot {\widehat{W}}_{1,ij}\) and \(\sqrt{l} \cdot {\widehat{W}}_{2,ij}\) are also i.i.d. matrices of size n. All these matrices are independent apart from the natural constraint \({\widehat{H}}_{ij} = {\widehat{H}}_{ji}^{*}\).

Define the events

$$\begin{aligned} \Theta _{\phi , N} := \left\{ \begin{array}{ll} \big \{ \Vert H\Vert \le 3,\, \Vert W_{1}\Vert \le 3, \, \Vert W_2\Vert \le 3, \,\Vert (W W^*)^{-1}\Vert \le \frac{1}{(1-\frac{1}{\sqrt{2 \phi }})^{2}(1-\delta )} \big \}, &{} \phi \in (1/2,1), \\ \{ \Vert H\Vert \le 3,\, \Vert W_{1}\Vert \le 3, \, \Vert W_2\Vert \le 3 \, \}, &{} \phi \in {(}0,1/2{]}, \end{array} \right. \end{aligned}$$
(3.2)

for some \(\delta >0\). Similarly as in Sect. 2.1, one can show that the events \(\Theta _{\phi ,N}\) hold a.w.o.p. and the random matrix models (1.9) and (1.10) for \(\phi \in (1/2,1)\) and \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\) restricted to \(\Theta _{\phi ,N}\) are well defined (see Remark A.21 and Lemma B.1). At the same time, in order to deal with \(\phi \in {(}0,1/2{]}\), we consider the regularized models (1.9) and (1.10) with \(w\in {\mathbb {C}}_{+} \), i.e., \(\mathrm {Im}w>0\) strictly positive, which guarantees the invertibility of \(w-H + \mathrm {i}\,\gamma W W^{*}\) without any additional restrictions on H and W.

The linearization matrix \(\varvec{H}_{w,\phi }\) for (1.10) is defined as in (2.1). This is a Kronecker random matrix consisting of \((6k+2l)\times (6k+2l)\) blocks of size n. We now introduce a tensorized version of the generalized resolvent that takes into account the additional structure coming from (3.1).

Definition 3.1

(Generalized resolvent). Let \((\phi , w) \in ({(}0,1/2{]}\times {\mathbb {C}}_{+})\cup ((1/2,1)\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\) with \(\phi = k/l \in {\mathbb {Q}}\). We call the matrix-valued function \({\mathbb {C}}_{+}\ni z \mapsto (\varvec{H}_{w,\phi }-z\varvec{J}_{k}\otimes I_{n})^{-1}\) the generalized resolvent of \(\varvec{H}_{w,\phi }\). Here, we denote \(\varvec{J}_{k}:= \sum _{i=1}^{k} E_{ii} \in {\mathbb {C}}^{(6k+2l)\times (6k+2l)}\) with \(\{E_{ij}\}\) being the standard basis of \({\mathbb {C}}^{(6k+2l)\times (6k+2l)}\).

Lemma 3.2

(Basic properties of the generalized resolvent). 

  1. (i)

    For any \(\gamma >0\) and \(\phi \in (1/2,1)\cap {\mathbb {Q}}\), there exists \(C_{\gamma ,\phi }>0\) such that a.w.o.p.

    $$\begin{aligned} \big \Vert (\varvec{H}_{w,\phi } - z\varvec{J}_{k}\otimes I_{n})^{-1}\big \Vert \le C_{\gamma ,\phi }\bigg (1 + \frac{1}{\mathrm {Im}z}\bigg ) \end{aligned}$$
    (3.3)

    uniformly for all \(z\in {\mathbb {C}}_{+}\) and \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\).

  2. (ii)

    For any \(\gamma >0\), \(\phi \in {(}0,1/2{]}\cap {\mathbb {Q}}\) and \(w\in {\mathbb {C}}_{+}\), there exists \(C_{\gamma ,\phi ,w}>0\) such that a.w.o.p.

    $$\begin{aligned} \big \Vert (\varvec{H}_{w,\phi } - z\varvec{J}_{k}\otimes I_{n})^{-1}\big \Vert \le C_{\gamma ,\phi ,w}\bigg (1 + \frac{1}{\mathrm {Im}z}\bigg ) \end{aligned}$$
    (3.4)

    for all \(z\in {\mathbb {C}}_{+}\).

  3. (iii)

    For all \((\phi , w) \in \big (({(}0,1/2{]}\cap {\mathbb {Q}})\times {\mathbb {C}}_{+}\big )\cup \big (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}})\big )\) and \(1\le i,j\le N\)

    $$\begin{aligned} \Big [(\varvec{H}_{w,\phi }-z\varvec{J}_{k}\otimes I_{n})^{-1}\Big ]_{ij} = \Big [(\varvec{T}_{w,\phi }- zI_N)^{-1}\Big ]_{ij} ,\quad 1\le i,j \le N . \end{aligned}$$
    (3.5)

Proof

Denote by \({\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) the effective domain related to the noncommutative rational function \(q_{1,w}(x,y_1,y_2,y_1^*,y_2^*)= w-x + \mathrm {i}\,\gamma (y_1y_1^* + y_2 y_2^*)\). More precisely [see (A.4)], the set \({\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) consists of all triples of elements \(({\hat{x}},{\hat{y}}_1,{\hat{y}}_2)\) such that

$$\begin{aligned} \left\| \frac{1}{q_{1,w}({\hat{x}},{\hat{y}}_1,{\hat{y}}_2,{\hat{y}}_1^*,{\hat{y}}_2^*)}\right\| \le C . \end{aligned}$$
(3.6)

As explained at the beginning of this section, for \(\phi = k/l \in (1/2,1)\) a.w.o.p. the random matrices H, \(W_1\) and \(W_2\) defined in (3.1) belong to the domain \({\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) with constant [see (3.2) with \(\delta = 1/2\)] \(C=\frac{1}{2 \gamma (1-\frac{1}{\sqrt{2\phi }})}\) depending only on \(\gamma \) and \(\phi \).

For \(\phi = k/l \in {(}0,1/2{]}\) on the other hand, the evaluation of the function \(q_{1,w}\) on random matrices H, \(W_1\) and \(W_2\) is invertible only for \(\mathrm {Im}w>0\), in which case the norm of the inverse of \(q_{1,w}\) evaluated on any triple \(({\hat{x}},{\hat{y}}_1,{\hat{y}}_2)\) is bounded by \(\frac{1}{\gamma \mathrm {Im}w}\).

With the above choice of the constant C in \({\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) depending on the value of \(\phi \), the proof of (i)–(iii) can be obtained from the same argument as in Lemma 2.3 by restricting \(\varvec{H}_{w,\phi }\) to the events \(\Theta _{\phi , N}\) defined in (3.2) and taking into account the dimensions of H, \(W_{1}\) and \(W_{2}\) and the relation \(N=kn\) (see Remark A.21). \(\square \)

From the structure of the linearization, we derive the DEL corresponding to \(\varvec{H}_{w,\phi }\)

$$\begin{aligned} -\frac{1}{M} = z \varvec{J}_{k} - K_0(w) + \Gamma _{\phi }[M] \end{aligned}$$
(3.7)

for an unknown matrix-valued function M depending on z, \(w\) and \(\phi \), having the following components:

  1. (i)

    the expectation matrix is given by

    $$\begin{aligned} K_0(w) := \left( \begin{array}{ccc|ccc|ccc} &{}\quad \kappa _{1}\otimes I_{k}&{} &{}&{}&{} &{}&{}&{} \\ &{}&{}&{}&{}&{}&{}&{}&{} \\ \hline &{}&{}&{}&{}&{}&{}&{}&{} \\ &{}&{} &{} &{}\quad \kappa _{2} \otimes I_{k}&{} &{}&{}&{} \\ &{}&{}&{}&{}&{}&{}&{}&{} \\ \hline &{}&{}&{}&{}&{}&{}&{}&{} \\ &{}&{}&{}&{}&{} &{} &{}\quad \kappa _{3}(w) \otimes I_{l}&{} \end{array} \right) \end{aligned}$$
    (3.8)

    with matrices \(\kappa _{1}, \kappa _{2},\kappa _{3}(w)\) defined as in (2.9)–(2.10);

  2. (ii)

    the operator \(\Gamma _{\phi } : {\mathbb {C}}^{(6k+2l)\times (6k+2l)} \rightarrow {\mathbb {C}}^{(6k+2l)\times (6k+2l)}\) maps an arbitrary matrix

    $$\begin{aligned} R = \left( \begin{array}{ccc} R_{11}&{}\quad R_{12}&{}\quad R_{13} \\ R_{21}&{}\quad R_{22}&{}\quad R_{23} \\ R_{31}&{}\quad R_{32}&{}\quad R_{33} \end{array} \right) \in {\mathbb {C}}^{(6k+2l)\times (6k+2l)} \end{aligned}$$
    (3.9)

    with \(R_{11}, R_{22} \in {\mathbb {C}}^{3k\times 3k}\), \(R_{33} \in {\mathbb {C}}^{2l\times 2l}\) into a block diagonal matrix with the first \(3k\times 3k\) block equal to

    $$\begin{aligned} \kappa _{5}^{t}\bigg (\Big ({\mathrm {id}}_{2}\otimes \frac{1}{l}{\mathrm {Tr}}_{l}\Big ) R_{33}\bigg )\kappa _{5}\otimes I_{k} , \end{aligned}$$
    (3.10)

    the second \(3k\times 3k\) diagonal block equal to

    $$\begin{aligned} \kappa _{4}^{t}\bigg (\Big ({\mathrm {id}}_{2}\otimes \frac{1}{l}{\mathrm {Tr}}_{l}\Big ) R_{33}\bigg )\kappa _{4}\otimes I_{k} , \end{aligned}$$
    (3.11)

    and the lower-right \(2l\times 2l\) block equal to

    $$\begin{aligned}&\kappa _{5}\bigg (\Big ({\mathrm {id}}_{3}\otimes \frac{1}{l}{\mathrm {Tr}}_{k}\Big ) R_{11}\bigg )\kappa _{5}^{t} \otimes I_{l} + \kappa _{4}\bigg (\Big ({\mathrm {id}}_{3}\otimes \frac{1}{l}{\mathrm {Tr}}_{k}\Big ) R_{22}\bigg )\kappa _{4}^{t} \otimes I_{l} \nonumber \\&\quad + \sigma _{1} \bigg ( \Big ({\mathrm {id}}_{2} \otimes \frac{1}{l}{\mathrm {Tr}}_{l}\Big ) R_{33}\bigg ) \sigma _{1} \otimes I_{l} . \end{aligned}$$
    (3.12)

    Here, for \(n\in {\mathbb {N}}\), we denote by \({\mathrm {Tr}}_{n}\) the trace of an \(n\times n\) matrix \(\kappa _{4},\kappa _{5}\) are defined as in (2.10) and \(\sigma _{1}\) is a standard Pauli matrix. The operator \(\Gamma _{\phi }\) is the tensorized analogue of \(\Gamma \) from (2.29).

Now we can proceed similarly as for \(\phi =1\) in Sect. 2 just the \((3+3+2)\) \(\times \) \((3+3+2)\) structure of the linearized matrices is replaced by larger block matrices structured as \((3k+3k+2l)\times (3k+3k+2l)\).

Lemma 3.3

(Existence and basic properties of the solution to the DEL (3.7)) For any \(\gamma >0\), \((\phi , w) \in (({(}0,1/2{]}\cap {\mathbb {Q}})\times {\mathbb {C}}_{+})\cup (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\) and \(z\in {\mathbb {C}}_{+}\) define \(M_{z, w} \in {\mathbb {C}}^{\,(6k+2l)\times (6k+2l)}\) as

$$\begin{aligned} M_{z,w} :=\,&({\mathrm {id}}_{6k+2l} \otimes \,\tau _{{\mathcal {S}}}) \bigg [ \Big ((K_0(w) - z \varvec{J}_{k}) \otimes \mathbb {1}_{{\mathcal {S}}}+ K_{1}\otimes H^{\mathrm {(sc)}} + L_1\otimes W_{1}^{\mathrm {(sc)}} \nonumber \\&+\,L_1^{*}\otimes \left( W_{1}^{\mathrm {(sc)}}\right) ^* + L_2 \otimes W_{2}^{\mathrm {(sc)}} + L_2^{*} \otimes \left( W_{2}^{\mathrm {(sc)}}\right) ^{*}\Big )^{-1}\bigg ] , \end{aligned}$$
(3.13)

where \(W_{1}^{(\mathrm {sc})}\) and \(W_{2}^{(\mathrm {sc})}\) are \(l\times k\) matrices consisting of freely independent circular elements multiplied by \(1/\sqrt{l}\), and \(H^{(\mathrm {sc})}\) is an \(l\times l\) self-adjoint matrix with freely independent semicircular elements multiplied by \(1/\sqrt{l}\) on the diagonal and freely independent circulars multiplied by \(1/\sqrt{l}\) above the diagonal.

Then,

  1. (i)

    For any \(\gamma >0\) and \(\phi \in (1/2,1)\cap {\mathbb {Q}}\), there exists \(C_{\gamma ,\phi }>0\) such that \(M_{z,w}\) satisfies the a priori bound

    $$\begin{aligned} \Vert M_{z,w} \Vert \le C_{\gamma ,\phi } \Big (1 + \frac{1}{\mathrm {Im}z}\Big ) \end{aligned}$$
    (3.14)

    uniformly for all \(w\in {\mathbb {C}}_{+}\cup {\mathbb {R}}\) and \(z\in {\mathbb {C}}_{+}\).

  2. (ii)

    For any \(\gamma >0\), \(\phi \in {(}0,1/2{]}\cap {\mathbb {Q}}\) and \(w\in {\mathbb {C}}_{+}\), there exists \(C_{\gamma ,\phi ,w}>0\) such that function \(M_{z,w}\) satisfies the a priori bound

    $$\begin{aligned} \Vert M_{z,w} \Vert \le C_{\gamma ,\phi ,w} \Big (1 + \frac{1}{\mathrm {Im}z}\Big ) . \end{aligned}$$
    (3.15)
  3. (iii)

    For any \(\gamma >0\), \((\phi , w) \in (({(}0,1/2{]}\cap {\mathbb {Q}})\times {\mathbb {C}}_{+})\cup (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\) and \(z\in {\mathbb {C}}_{+}\), matrix \(M_{z,w}\) satisfies the DEL (3.7) and has positive semidefinite imaginary part, \(\mathrm {Im}M_{z,w} \ge 0\). Moreover, for all \(\gamma >0\), \((\phi , w) \in (({(}0,1/2{]}\cap {\mathbb {Q}})\times {\mathbb {C}}_{+})\cup (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\), the matrix-valued function \(z\mapsto M_{z,w}\) is analytic on \({\mathbb {C}}_{+}\).

  4. (iv)

    For any \(\gamma >0\) and \((\phi , w) \in (({(}0,1/2{]}\cap {\mathbb {Q}})\times {\mathbb {C}}_{+})\cup (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\) function \(z\mapsto M_{z,w}\) admits the representation

    $$\begin{aligned} M_{z,w} = M^{\infty }_{w,\phi } + \int _{{\mathbb {R}}} \frac{V_{w,\phi }(d\lambda )}{\lambda - z} , \end{aligned}$$
    (3.16)

    where \(M_{w,\phi }^{\infty } \in {\mathbb {C}}^{(6k + 2l)\times (6k+2l)}\) is a self-adjoint matrix, and \(V_{w,\phi }(d\lambda )\) is a positive-semidefinite matrix-valued measure on \({\mathbb {R}}\) with compact support.

Proof

The proof follows from parts (i)–(v) of Lemma A.12 (see also Remark A.21). Similarly as in the proof of Lemma 3.2, notice that the noncommutative rational expression \(q_{1,w}= w-x + \mathrm {i}\,\gamma (y_1y_1^* + y_2 y_2^*)\) evaluated on matrices \( x= H^{\mathrm {(sc)}}\), \(y_1 = W_1^{\mathrm {(sc)}}\) and \(y_2 = W_2^{\mathrm {(sc)}}\) expresses different behavior in variable \(w\) for \(\phi \in (0,1/2) \) and \(\phi \in [1/2,1)\).

In the first case, \(\phi \in (1/2,1)\), the invertibility of \(q_{1,w}\) evaluated on \((s,c_1,c_2)\) does not depend on \(w\) and thus \((s,c_1,c_2) \in {\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) with \(C=C(\gamma ,\phi )\). In the case \(\phi \in {(}0,1/2{]}\), \(q_{1,w}\) is invertible if and only if \(\mathrm {Im}w>0\), and the norm of \((q_{1,w})^{-1}\) depends on \(\mathrm {Im}w\). This, in particular, means that \((s,c_1,c_2) \in {\mathcal {D}}_{\varvec{q}_{0},\{q_{1,w}\};C}\) with a \(w\)-dependent constant \(C(\gamma ,\phi ,w)\).

This leads to two different a priori estimates: a bound (3.14) uniform in \(w\) for \(\phi \in (1/2,1)\) and a \(w\)-dependent bound for \(\phi \in {(}0,1/2{]}\). The rest of the proof follows directly from Lemma A.12. \(\square \)

We omit the dependence of \(M_{z,w}\) on \(\phi \) for brevity. With these notations, we have the following global law establishing Theorem 1.3 and partially (i) of Theorem 1.4 for \(\phi \in (0,1)\). The proof of the weak limit (1.12) for \(\phi \in (0,1/2)\) is postponed to Sect. 3.6.

Lemma 3.4

(Global law for \(\varvec{T}_{w,\phi }\), \(\phi \in (0,1)\)). For \((\phi , w) \in ({(}0,1/2{]}\times {\mathbb {C}}_{+})\cup ((1/2,1)\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\), \(\phi = k/l\), the empirical spectral measure \(\mu _{\varvec{T}_{w,\phi }}(d\lambda )\) converges weakly in probability (and almost surely) to \(\rho _{w, \phi }(d\lambda )\), where

$$\begin{aligned} \rho _{w,\phi }(d\lambda ) := \frac{1}{k}{\mathrm {Tr}}( \varvec{J}_{k} \, V_{w,\phi }(d\lambda ) ) \end{aligned}$$
(3.17)

is the normalized trace of the upper-left \(k\times k\) submatrix of the matrix-valued measure \(V_{w,\phi }(d\lambda )\) from (3.16). The support of the measure \(\rho _{w, \phi }(d\lambda )\) is a subset of the interval [0, 1]. Moreover, for any \(\phi \in (1/2,1)\) and \(E\in {\mathbb {R}}\), the measure \(\rho _{w, \phi }(d\lambda )\) converges weakly to \(\rho _{ E , \phi }(d\lambda )\) as \(w\in {\mathbb {C}}_+\) tends to \(E\in {\mathbb {R}}\).

Proof

The proofs are similar to the case \(\phi =1\) (see Lemma 2.5 and Remark A.21) after taking into account the dimensions of the matrices H, \(W_1\) and \(W_2\) and the additional structure (3.1). \(\square \)

Definition 3.5

(Self-consistent density of states). We call the function

$$\begin{aligned} \rho _{w,\phi }(\lambda ) := \lim _{\eta \downarrow 0} \frac{1}{\pi k} \mathrm {Im}{\mathrm {Tr}}(\varvec{J}_{k} M_{\lambda +\mathrm {i}\,\eta ,w}) \end{aligned}$$
(3.18)

that gives the absolutely continuous part of \(\rho _{w,\phi }(d\lambda )\), the self-consistent density of states of the model (1.10).

Since \(\mathrm {supp} (\rho _{w,\phi })\subset [0,1]\) by unitarity of \(S(w)\), part (iii) of Theorem 1.4 can be established by proving the boundedness of the upper-left \(k\times k\) minor of \(M_{z,w}\) for the spectral parameter z bounded away from 0 and 1 (Sect. 3.3), and analyzing the asymptotic behavior of this upper-left submatrix in the vicinity of the special points \(z=0\) and \(z=1\) (Sect. 3.4). The study of \(M_{z,w}\) is simplified by the particular form of \(K_{0}(w)\) and \(\Gamma _{\phi }\), which implies that

$$\begin{aligned} M_{z,w} = \left( \begin{array}{ccc} M_1\otimes I_{k}&{}&{} \\ &{}\quad M_{2} \otimes I_{k} &{} \\ &{}&{}\quad M_{3} \otimes I_{l} \end{array} \right) \end{aligned}$$
(3.19)

with \(M_1,M_2\in {\mathbb {C}}^{\,3\times 3}\) and \(M_{3}\in {\mathbb {C}}^{\,2\times 2}\) satisfying

$$\begin{aligned} -\frac{1}{M_1} = zJ_{3}-\kappa _{1} + \kappa _{5}^{t} M_{3} \kappa _{5} ,\quad -\frac{1}{M_{2}} = -\kappa _{2} + \kappa _{4}^{t} M_{3} \kappa _{4} \end{aligned}$$
(3.20)

and

$$\begin{aligned} -\frac{1}{M_{3}} = - \kappa _{3}(w) + \phi \kappa _{5} M_{1} \kappa _{5}^{t} + \phi \kappa _{4} M_{2} \kappa _{4}^{t} + \sigma _{1} M_{3} \sigma _{1} . \end{aligned}$$
(3.21)

Similarly as in the case \(\phi =1\), plugging (3.20) into (3.21) leads to the following self-consistent equation for \(M_{3}\)

$$\begin{aligned} -\frac{1}{M_{3}}= & {} - \kappa _{3}(w) - \frac{ \phi }{-\frac{1}{2\gamma ^2 z}(I_{2}+ \sigma _{3})-\frac{1}{\gamma }\sigma _{2}+M_{3}} \nonumber \\&- \frac{\phi }{-2(I_{2}- \sigma _{3})-\frac{1}{\gamma }\sigma _{2}+M_{3}} + \sigma _{1} M_{3} \sigma _{1} , \end{aligned}$$
(3.22)

which is the analogue of (2.32) for \(\phi \ne 1\).

3.2 Useful Identities

Below we prove that identities similar to (2.34), (2.38) and (2.45) hold for \(\phi \in (0,1)\).

Lemma 3.6

For all \((\phi , w) \in (({(}0,1/2{]} \cap {\mathbb {Q}})\times {\mathbb {C}}_{+})\cup (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\), \(\gamma >0\) and \(z\in {\mathbb {C}}_{+}\)

  1. (i)

    \(M_{z,w}(i,i) = M_{z,-{\overline{w}}}(i,i)\) for all \(1\le i \le 6k+2l\);

  2. (ii)

    \( M_{z,w}(6k+l+1,6k+l+1) = 4 \gamma ^2 z M_{z,w}(6k+1,6k+1)\);

  3. (iii)

    \( M_{z,w}(6k+l+1,6k+1) - M_{z,w}(6k+1,6k+l+1) = \frac{\mathrm {i}\,}{ \gamma }M_{z,w}(6k+l+1,6k+l+1)\Big (1-\frac{ \gamma \mathrm {Im}w}{2 \phi } \det T_1\Big )\), where, similarly as in (2.46), we denoted

    $$\begin{aligned} T_1 = \left( \begin{array}{cc} -\frac{1}{ \gamma ^2 z} &{}\quad \frac{\mathrm {i}\,}{ \gamma } \\ -\frac{\mathrm {i}\,}{ \gamma } &{}\quad 0 \end{array} \right) + M_{3} \end{aligned}$$
    (3.23)

Denote the entries of \(M_{3}\) in (3.19) by \(m_{ij}\), \(1\le i,j\le 2\). Then, the parts (ii) and (iii) of the above lemma can be rewritten as

$$\begin{aligned} m_{22}&= 4 \gamma ^2 z m_{11} , \end{aligned}$$
(3.24)
$$\begin{aligned} m_{21} - m_{12}&= \frac{\mathrm {i}\,}{ \gamma }m_{22} \Big (1-\frac{ \gamma \mathrm {Im}w}{2 \phi } \det T_1\Big ) . \end{aligned}$$
(3.25)

Proof

In order to establish Lemma 3.6, we can follow the proofs of Lemmas 2.62.8 and apply them to the matrix

$$\begin{aligned}&(K_0(w) - z \varvec{J}_{k}) \otimes \mathbb {1}_{{\mathcal {S}}}+ K_{1}\otimes H^{\mathrm {(sc)}} + L_1\otimes W_{1}^{\mathrm {(sc)}} +L_1^{*}\otimes \left( W_{1}^{\mathrm {(sc)}}\right) ^* \nonumber \\&\quad + L_2 \otimes W_{2}^{\mathrm {(sc)}} + L_2^{*} \otimes \left( W_{2}^{\mathrm {(sc)}}\right) ^{*} . \end{aligned}$$
(3.26)

Note that the above matrix (3.26) is obtained from (2.35) by substituting \(c_{1}\), \(c_{2}\) and \(s\) with matrices \(W_{1}^{(\mathrm {sc})}\), \(W_{2}^{(\mathrm {sc})}\) and \(H^{(\mathrm {sc})}\) correspondingly, and taking into account the dimensions of these matrices. For example, if we replace each diagonal entry of the matrix \(Q^{-}\) from the proof of Lemma 2.6 by the tensor product of this entry and a corresponding identity matrix (\(I_{k}\) or \(I_{l}\)), we obtain that the diagonal blocks of \(M_{z,w}\) and \(M_{z, -{\overline{w}}}\) coincide.

In order to prove (3.24), similarly as in the proof of Lemma 2.7, use the Schur complement formula with respect to the invertible upper-left \(6k\times 6k\) submatrix of (3.26) to write the \(2l\times 2l\) lower-right submatrix of its inverse as

$$\begin{aligned} \left( \begin{array}{cc} 4 \gamma ^2 W_{1}^{(\mathrm {sc})} \left( W_{1}^{(\mathrm {sc})}\right) ^* &{} w{-}H^{(\mathrm {sc})}{+} \mathrm {i}\,\gamma \Big (W_{1}^{(\mathrm {sc})} \big (W_{1}^{(\mathrm {sc})}\big )^* {+} W_{2}^{(\mathrm {sc})} \big (W_{2}^{(\mathrm {sc})}\big )^*\Big ) \\ {\overline{w}}{-}H^{(\mathrm {sc})}{-} \mathrm {i}\,\gamma \left( W_{1}^{(\mathrm {sc})} \big (W_{1}^{(\mathrm {sc})}\big )^* {+} W_{2}^{(\mathrm {sc})} \big (W_{2}^{(\mathrm {sc})}\big )^*\right) &{} \frac{1}{z} W_{2}^{(\mathrm {sc})} \big (W_{2}^{(\mathrm {sc})}\big )^* \end{array} \right) ^{-1} . \end{aligned}$$
(3.27)

Then, we can switch the blocks of (3.27) as in (2.40) and apply the Schur complement formula with respect to

$$\begin{aligned} w- H^{(\mathrm {sc})} + \mathrm {i}\,\gamma \Big (W_{1}^{(\mathrm {sc})} \big (W_{1}^{(\mathrm {sc})}\big )^* + W_{2}^{(\mathrm {sc})} \big (W_{2}^{(\mathrm {sc})}\big )^*\Big ) . \end{aligned}$$
(3.28)

The expression in (3.28) is invertible: for \((\phi , w) \in (((1/2,1)\cap {\mathbb {Q}})\times ({\mathbb {C}}_{+}\cup {\mathbb {R}}))\) the spectrum of \(W_{1}^{(\mathrm {sc})} (W_{1}^{(\mathrm {sc})})^* + W_{2}^{(\mathrm {sc})} (W_{2}^{(\mathrm {sc})})^*\) follows the free Poisson distribution with rate \(2\phi \in (1,2)\) and is therefore bounded away from zero, and for \((\phi , w) \in (({(}0,1/2{]}\cap {\mathbb {Q}})\times {\mathbb {C}}_{+})\) the invertibility is guaranteed by \(\mathrm {Im}w\) being strictly positive. Due to the properties of freely independent circular and semicircular elements, switching the labels of the pair \((W_{1}^{(\mathrm {sc})},W_{2}^{(\mathrm {sc})})\) or changing the sign of \(H^{(\mathrm {sc})}\) does not change the value of an expression involving these matrices after applying \({\mathrm {id}}_{6k+2l}\otimes \,\tau _{{\mathcal {S}}}\). Therefore, by proceeding as in (2.42)–(2.44) with \(M_{z,w}(7,7)\) and \(M_{z, w}(8,8)\) replaced by the corresponding \(l\times l\) blocks of \(M_{z, w}\), and using the diagonal structure of these blocks (3.19) and the part (i) of this lemma, we obtain (3.24).

Now it is straightforward to check by plugging (3.24) into (3.22) and following the proof of Lemma 2.8, that (3.25) holds. This proves Lemma 3.6.

\(\square \)

3.3 Boundedness of \(M_{z,w}\) away from \(z=0\) and \(z=1\) for \(\phi \in (0,1)\)

The goal of this section is to establish a uniform bound on \(\Vert M_{z,w}\Vert \) for parameter \(w\) close to the real line and parameter z bounded away from 0 and 1. This will be used later in the proof of Theorem 1.4, in particular to show the absolute continuity of the measure \(\rho _{E,\phi }(d \lambda )\). In the case \(\phi \in (1/2,1)\) we can set \(w= E \in {\mathbb {R}}\) and work directly with \(M_{z,E}\). For \(\phi \in {(}0,1/2{]}\), we prove the uniform bound for \(\Vert M_{z,w}\Vert \) with \(w= E + \mathrm {i}\,\phi \, \epsilon \) and small \(\epsilon >0\), which will allow taking the limit \(\epsilon \rightarrow 0\) in Sect. 3.4.

Lemma 3.7

(Boundedness of \(M_{z,w}\)

  1. (i)

    Case \(\phi \in {(}1/2,1{]}\): For any \(\gamma >0\) and small \(\theta >0\), there exists \(C_{\theta , \gamma }>0\) such that

    $$\begin{aligned} \sup \Big \{\,\Vert M_{z,E}\Vert \, : \, \phi \in (1/2,1)\cap {\mathbb {Q}},\, | z |\ge \theta , \, |1- z | \ge \theta , \, \mathrm {Im}z > 0 ,\, |E|\le \frac{1}{\theta }\, \Big \} \le C_{\theta , \gamma }; \end{aligned}$$
    (3.29)
  2. (ii)

    Case \(\phi \in (0, 1)\): Let \(w= E + \mathrm {i}\,\phi \epsilon \), \(\epsilon >0\). For any \(\gamma >0\), small \(\theta >0\), \(\phi _0\in (0,1/2)\) and \(\epsilon _0>0\) small enough there exists \(C_{\theta , \gamma ,\phi _0,\epsilon _0}>0\) such that

    $$\begin{aligned} \Vert M_{z,w}\Vert \le C_{\theta , \gamma ,\phi _0,\epsilon _0}\end{aligned}$$
    (3.30)

    uniformly on the set

    $$\begin{aligned} \Big \{ \phi \in [\phi _0,1-\phi _0]\cap {\mathbb {Q}}, |z|\ge \theta , |1-z| \ge \theta , \mathrm {Im}z > 0, |E|\le \frac{1}{\theta }, \epsilon \in (0, \epsilon _0]\, \Big \} . \end{aligned}$$
    (3.31)

Proof

Consider first (3.29) for which \(\phi \in (1/2,1)\). By setting \(\epsilon = 0\) in Lemma 3.6, we can proceed by establishing (3.29) in the same manner as in the proof of Lemma 2.9. To guarantee a uniform bound in parameters z, E and \(\phi \), instead of a sequence \((z_n, E_{n})_{n=1}^{\infty }\) as in the proof of Lemma 2.9, we assume the existence of a sequence \(((z_n, E_n, \phi _n))_{n=1}^{\infty }\), \(z_n \in {\mathbb {C}}_+\), \(|E|\le 1/\theta \), \(\phi _n \in (1/2,1)\), on which \(|m_{11}^{(n)}|\rightarrow \infty \) or \(|\det T_{1}^{(n)}|\rightarrow 0\). Note that for \(\phi _n\in (1/2,1)\) the leading terms in the analogues of the mutually contradicting pairs of statements (2.67)/(2.70), (2.71)/(2.74) and (2.79)/(2.80) do not depend on \(\phi _n\) (with the exception of (2.71) where the constant \(-1\) is replaced by \(\phi _n - 2\)).

For (3.30), i.e., \(\phi \in [\phi _0,1- \phi _0]\), the above argument has to be slightly adjusted to ensure a uniform bound for small \(\epsilon >0\). Instead of a sequence \(((z_n, E_n, \phi _n))_{n=1}^{\infty }\) as in the first part of the proof, we now assume the existence of a sequence \(((z_n, E_n ,\phi _n, \epsilon _n))_{n=1}^{\infty }\), \(z_n \in {\mathbb {C}}_+\), \(|E|\le 1/\theta \), \(\phi _n \in [\phi _0,1-\phi _0]\), \(\epsilon _n>0\), on which \(|m_{11}^{(n)}|\rightarrow \infty \) or \(|\det T_{1}^{(n)}|\rightarrow 0\).

The analogues of (2.67)/(2.70) in this case are

$$\begin{aligned} \det M_{3}^{(n)}&= -4(z_{n}-1)m_{11}^{(n)} + O\left( 1\right) , \end{aligned}$$
(3.32)
$$\begin{aligned} \det M_{3}^{(n)}&= 4 \gamma ^2 z_n (1-z_n(1-\epsilon _n\phi _n \gamma )^2)\left( m_{11}^{(n)}\right) ^2 + O\left( m_{11}^{(n)}\right) , \end{aligned}$$
(3.33)

which contradict each other if \(|m_{11}^{(n)}|\rightarrow \infty \), \(\epsilon _n\le \epsilon _0\) is small enough and \(z_n\) is bounded away from 0 and 1.

Instead of the pair (2.71)/(2.74), we take the analogues of (2.71) and (2.72) for \(\phi _n \in [\phi _0, 1-\phi _0]\) and \(\epsilon _n>0\)

$$\begin{aligned} \det M_{3}^{(n)}&= \phi _n -2 +O\left( \frac{1}{\big |m_{11}^{(n)}\big |}\right) , \end{aligned}$$
(3.34)
$$\begin{aligned} \det T_{1}^{(n)}&= 4(z_{n}-1)m_{11}^{(n)} + O\left( 1\right) , \end{aligned}$$
(3.35)

and observe that since \(1 + 1/(\phi _n-2)\ge \phi _0/(1+\phi _0)>0\) for \(\phi _n \in [\phi _0,1-\phi _0]\), the above equations contradict to

$$\begin{aligned} \left( \frac{1}{\det M_{3}} -\frac{2 \phi _n}{\det T_{1}} + 1 \right) m_{11}^{(n)} = -\frac{\phi _n}{ \gamma ^2 z\det T_1}, \end{aligned}$$
(3.36)

the analogue of (2.52), in the regime \(|m_{11}^{(n)}|\rightarrow \infty \).

For the last pair (2.79)/(2.80) note that in the analogue of (2.75)

$$\begin{aligned} \Big (\frac{\phi _n}{ \gamma ^2 z_{n} m_{11}^{(n)} }+O\left( \det T_{1}^{(n)}\right) \Big )m_{12}^{(n)} = -\frac{2 \phi _n \mathrm {i}\,}{ \gamma } + O\left( \det {T_{1}}^{(n)}\right) \end{aligned}$$
(3.37)

with \(|\det {T_{1}}^{(n)}|\rightarrow 0\), the parameter \(\phi _n\) disappears after dividing (3.37) by \(\phi _n\); therefore, we can proceed exactly as in the proof of Lemma 2.9. We conclude, similarly as in Lemma 2.9, that \(\Vert M_{z,w}\Vert \) is uniformly bounded provided that \(\min \{|z|, |1-z|\} \ge \theta \), \(|E|\le 1/\theta \), \(\phi _0 \le \phi \le 1-\phi _0\) and \(\epsilon \le \epsilon _0\) for some \(\phi _0 >0\) and \(\theta ,\epsilon _0>0\) small enough. \(\square \)

3.4 Singularities of \(M_{z,E}(1,1)\)

For \(\phi \in (1/2,1)\), the solution matrix \(M_{z,E}\) with \(E\in {\mathbb {R}}\) is given directly via (3.13). For \(\phi \in {(}0,1/2{]}\), this formula cannot be directly applied when \(w= E \in {\mathbb {R}}\) is real; we need an additional regularization argument. Nevertheless, in the next lemma we show the existence of the solution to the Dyson equation (3.20)–(3.22) for \(\phi \in {(}0,1/2{]}\) and \(w= E \in {\mathbb {R}}\) and we establish the asymptotic behavior of \(M_{z, E}(1,1)\) near \(z=0\) and \(z=1\) for \(\phi \in (0,1)\). We start by constructing an expansion of the solution \(M_{z, w}\) in the vicinity of \(z=0\) for \(\phi \in (0,1)\) and \(w= E + \mathrm {i}\,\phi \epsilon \) with \(\epsilon >0\) sufficiently small. We then use this expansion to extend \(M_{z,w}\) to \(w= E \in {\mathbb {R}}\) for \(\phi \in {(}0,1/2{]}\) by taking \(\epsilon \downarrow 0\) and to study the asymptotic behavior of \(M_{z,E}\) at special points \(z=0\) and \(z=1\).

Recall that the solution to the Dyson equation has the block structure (3.19), where \(M_{1},M_{2}\) are determined by \(M_{3}\) through (3.20) and \(M_{3}\) satisfies (3.22).

Lemma 3.8

(Existence of \(M_{z,E}\) for \(\phi \in {(}0,1/2{]}\) and singularities of \(M_{z, E}(1,1)\) for \(\phi \in (0,1)\)). 

  1. (a)

    For \(\gamma >0\), let \(M_{z,w, \phi }\) denote the function \(M_{z,w}\) defined in (3.13) evaluated at a point \((z,w,\phi )\) with

    $$\begin{aligned} z\in {\mathbb {C}}_{+} \quad \text{ and } \quad (w,\phi )\in \big ({\mathbb {C}}_{+}\times \big ({(}0,1/2{]}\cap {\mathbb {Q}}\big )\big )\cup \big (({\mathbb {C}}_{+}\cup {\mathbb {R}}) \times \big ((1/2,1)\cap {\mathbb {Q}}\big )\big ). \end{aligned}$$
    (3.38)

    Then \(M_{z,w,\phi }\) can be continuously extended to the set

    $$\begin{aligned} z\in {\mathbb {C}}_{+} \quad \text{ and } \quad (w,\phi )\in {\mathbb {R}}\times (0,1) \end{aligned}$$
    (3.39)

    in the following sense: for any \((w,\phi ) \in {\mathbb {R}}\times (0,1)\) and any sequence \(\{(w_{n},\phi _{n}) \}_{n\ge 1} \subset \big ({\mathbb {C}}_{+}\times \big ({(}0,1/2{]}\cap {\mathbb {Q}}\big )\big )\cup \big (({\mathbb {C}}_{+}\cup {\mathbb {R}}) \times \big ((1/2,1)\cap {\mathbb {Q}}\big )\big )\) with \((w_{n},\phi _{n})\rightarrow (w,\phi )\) as \(n\rightarrow \infty \), there exists an analytic matrix-valued function \(M_{z,w,\phi }:{\mathbb {C}}_+ \rightarrow {\mathbb {C}}^{(6k + 2l)\times (6k + 2l)}\), \( z\mapsto M_{z,w,\phi }\), such that

    $$\begin{aligned} M_{z,w_{n},\phi _{n}} \rightarrow M_{z,w,\phi } \end{aligned}$$
    (3.40)

    uniformly on compact z-subsets of \({\mathbb {C}}_+\) as \(n\rightarrow \infty \).

    For \(\phi \in (0,1)\) and \(w= E \in {\mathbb {R}}\), denote by \(M_{z,E}:= M_{z,E,\phi }\) the function defined in (3.40) omitting explicitly the dependence on \(\phi \).

  2. (b)

    For all \(\phi \in (0,1)\), \(\gamma >0\) and \(E\in {\mathbb {R}}\)

    $$\begin{aligned} M_{z,E}(1,1) = \mathrm {i}\,\frac{4+ \gamma ^2 \nu _0^2+ \gamma ^2 \xi _0^2}{4 \gamma \xi _0}\,z^{-1/2} + O(1)\, \text{ as } z\rightarrow 0 , \end{aligned}$$
    (3.41)

    with constants \(\xi _{0} := \xi _{0} ( \phi , \gamma ) \) and \(\nu _{0} := \nu _{0} ( \phi , \gamma ) \) given in (3.49) in the proof below.

  3. (c)

    For all \(\phi \in (0,1)\), \(\gamma >0\) and \(|E| \le E_{0}:=E_{0}(\phi ,\gamma )\)

    $$\begin{aligned} M_{z,E}(1,1) = \frac{4\xi _0}{(\xi _0^2 + \gamma ^2 E^2+4 )}(z-1)^{-1/2} + O\left( 1\right) \text{ as } z\rightarrow 1 , \end{aligned}$$
    (3.42)

    with constants \(E_{0}\) and \( \xi _{0}:= \xi _{0} ( \phi , \gamma ) <0\) given in (3.52) and (3.54) correspondingly.

The branch of the square root is chosen to be continuous on \({\mathbb {C}}{\setminus } (\mathrm {i}(-\infty ,0])\) such that \(\sqrt{1}=1\).

Proof

The analysis of (3.22) for \(\phi \in (0,1)\) will follow similar steps as the analysis of (2.32) for the \(\phi =1\) case as performed in Sect. 2.4, and we will omit the details of some straightforward albeit tedious calculations. In the first step below, we analyze the solution to (3.22) for rational \(\phi \in (0,1)\) and \(w= E + \mathrm {i}\,\phi \, \epsilon \) with \(E\in {\mathbb {R}}\) and \(\epsilon >0\) small enough. Using the same procedure that led to (2.84), we rewrite this equation as \(\Delta =0\) with

$$\begin{aligned}&\Delta := \Big (Z_1+(1-2\phi )M_{3} + M_{3} \sigma _1(M_{3}\sigma _1-E+\mathrm {i}\epsilon \phi \sigma _3 )(Z_1+M_{3})\Big )\nonumber \\&\qquad \frac{1}{Z_1-Z_2}(Z_2+M_{3}) -\phi M_{3} , \end{aligned}$$
(3.43)

and the two \(2 \times 2\)-matrices

$$\begin{aligned} Z_1 := -\frac{1}{2 \gamma ^2 z}(I_2+\sigma _3) -\frac{1}{ \gamma }\sigma _2 , \qquad Z_2 := -2(I_2-\sigma _3) -\frac{1}{ \gamma }\sigma _2 . \end{aligned}$$
(3.44)

Step 1: Expansion around \(z=0\). We construct an expansion of \(M_3\) as a power series in \(t=\sqrt{z}\) in a neighborhood of \(z=0\). We make an ansatz compatible with the symmetries (3.24) and (3.25), namely

$$\begin{aligned} {\widetilde{M}}_3(t) = \left( \begin{array}{cc} \frac{\mathrm {i}\xi }{4 \gamma t} &{}\quad \frac{\nu }{2} + \frac{\xi {t}}{2}(1+\frac{\epsilon }{ \gamma } f )+\epsilon \frac{\mathrm {i}\xi ^2}{4} \\ \frac{\nu }{2} - \frac{\xi {t}}{2} (1+\frac{\epsilon }{ \gamma } f)-\epsilon \frac{\mathrm {i}\xi ^2}{4} &{}\quad \mathrm {i}\gamma \xi {t}\end{array} \right) , \end{aligned}$$
(3.45)

for the two functions \(\xi =\xi (t),\nu =\nu (t)\) of t to be determined and where the parameters \(E,\gamma , \phi ,\epsilon \) are considered fixed. Later, we will show that with the right choice of functions \(\xi , \nu \) this ansatz coincides with the solution of the Dyson equation, i.e., that \( {\widetilde{M}}_3(t)= {M}_3(z)\). Here, f solves the equation \(q_0=0\) for sufficiently small \(\epsilon \ge 0\), where

$$\begin{aligned} q_0 = f+ \frac{ \gamma ^2 }{2}\det \Bigg [ \left( \begin{array}{cc} -\frac{1}{ \gamma ^2 t^2} &{}\quad \frac{\mathrm {i}}{ \gamma } \\ -\frac{\mathrm {i}}{ \gamma } &{}\quad 0 \end{array} \right) + {\widetilde{M}}_3\Bigg ]+\frac{\mathrm {i}\gamma \xi }{2{t}} \end{aligned}$$
(3.46)

is a polynomial in all variables \(t,E, \gamma ,\xi ,\nu ,\epsilon \) and f, in which it is quadratic. The choice (3.46) for f ensures that the symmetry condition (3.25) is satisfied.

We plug (3.45) into (3.43) and, after a mechanical but very long calculation that uses \(q_0=0\) in the (2, 1)- and (2, 2)-entries of \(\Delta \), find

$$\begin{aligned} \Delta = \left( \begin{array}{cc} -\frac{\mathrm {i}}{16 \gamma t}q_{1,\epsilon ,t} &{}\quad \frac{1}{4}q_{2,\epsilon ,t} \\ -\frac{1}{4}q_{2,\epsilon ,t} &{}\quad \frac{ \mathrm {i}\gamma {t}}{4}q_{1,\epsilon ,t} + (\mathrm {i}\gamma t^2+\mathrm {i}t \epsilon f{t}- \frac{ \gamma \epsilon \xi {t}}{2})q_{2,\epsilon ,t} \end{array} \right) , \end{aligned}$$
(3.47)

where

$$\begin{aligned} q_{1,\epsilon ,t} =q_1+\epsilon {\widetilde{q}}_1 + \frac{t}{4 \gamma ^3 } p_{1}, \qquad q_{2,\epsilon ,t} =q_2 + \frac{t}{ \gamma } p_{2}, \end{aligned}$$

and \(p_1,p_2\) are polynomials in \(t,E, \gamma ,\phi ,\xi ,\nu , \epsilon \) and \(q_1=q_{1,0,0},q_2=q_{2,0,0},{\widetilde{q}}_1\) are the following explicitly defined functions of the unknowns \((\xi ,\nu )\):

$$\begin{aligned} q_1= & {} \xi (\xi ^2 - \nu ^2+2E\nu +4(\phi -1)), \quad q_2= \xi ^2 (\nu -E)+2\nu \phi , \nonumber \\ {\widetilde{q}}_1= & {} -\mathrm {i}\,\xi q_2+\epsilon \, \Big (\frac{\xi ^{5}}{4}+\xi ^3\phi \Big )\, . \end{aligned}$$
(3.48)

In particular, the equation \(\Delta =0\) is equivalent to \((q_{1,\epsilon ,t},q_{2,\epsilon ,t})=0\) and thus to \((q_1,q_2) = 0\) in the limit \(t \rightarrow 0\) and \(\epsilon \rightarrow 0\). This, in turn, fixes the values for \(\nu _0=\nu |_{t=0}\) and \(\xi _0=\xi |_{t=0}\) through

$$\begin{aligned} \nu _0 = \frac{E\xi _0^2 }{ \xi _0^2+2\phi }, \quad r := \xi _0^6 + (E^2+8\phi -4) \xi _0^4+ 4 \phi (E^2+ 5\phi -4) \xi _0^2+16(\phi -1)\phi ^2=0, \end{aligned}$$
(3.49)

where we choose the positive solution \(\xi _0\) for \(r=0\). The fact that \(r=0\) has a unique positive solution \(\xi _0>0\) is an explicit elementary calculation. The positivity of \(\xi _0\) will ensure the positive definiteness of \(\mathrm {Im}{\widetilde{M}}_3\) for \(z \in {\mathbb {C}}_+\).

We compute the Jacobian of the function \((q_1, q_2)\) from (3.48) as

$$\begin{aligned} J(\xi ,\nu ):= & {} \det \nabla _{\xi ,\nu }(q_1,q_2) = 3\xi ^4+ ((3 \nu ^2-6 \nu E+4 E^2)+10\phi -4)\xi ^2\nonumber \\&\quad +2\phi (4\phi -4+2E\nu -\nu ^2). \end{aligned}$$
(3.50)

Using that at \(t=0\) we have \(q_1=0\) and \( \xi _0\ne 0\), we can eliminate the quadratic terms in \(\nu \) and obtain

$$\begin{aligned} J(\xi _0,\nu _0) = 2\xi _0^2(3 \xi _0^2 +2E^2 + 10\phi -8). \end{aligned}$$

Again an elementary calculation using the defining equation \(r=0\) for \(\xi _0\) shows that \(J(\xi _0, \nu _0)\) never vanishes. Thus, the ansatz (3.45) solves the Dyson equation in a small neighborhood of \(z=t^2=0\). Furthermore, for \(t = u (1+\mathrm {i}u )\) with sufficiently small \( u >0\) it is easy to see that its imaginary part is positive definite. By using a regularization argument analogous to the one from Step 3 of Sect. 2.4 and combining it with the uniqueness of solutions to the Dyson equation with positive definite imaginary part, this implies that \({\widetilde{M}}_3(t) = M_3(z)\) for all \(\sqrt{z}=t = u (1+\mathrm {i}u )\), \(\phi \in (0,1)\) and small enough \(\epsilon >0\). Since both functions are analytic this also implies equality for \(z=t^2\) in the complex upper half plane intersected with a neighborhood of \(z=0\).

Step 2: Extending \(M_{z,w}\) to \(w= E \in {\mathbb {R}}\) for \(\phi \in (0,1) \). Fix \(\phi \in (0,1) \), \(\gamma >0\) and \(E\in {\mathbb {R}}\). For \(\epsilon _0>0\) and a coordinate pair \((i,j)\in \{1,\ldots , 6k + 2l\}^2\), consider the family of functions \(\{M_{z,E+\mathrm {i}\,\epsilon _n \phi _{n} }\,(i,j)\}_{n\ge 1}\) analytic in z with \(\{\phi _{n},\epsilon _n\}_{n\ge 1}\subset ((0,1)\cap {\mathbb {Q}}) \times (0,\epsilon _0)\) and \((\phi _{n},\epsilon _{n})\rightarrow (\phi ,0)\) as \(n\rightarrow \infty \). It follows from part (ii) of Lemma 3.7 that for small enough \(\epsilon _0\) and any small \(\theta >0\) the family of functions \(\{M_{z,E+\mathrm {i}\,\epsilon _n \phi _{n}}\,(i,j)\}_{n\ge 1}\) is uniformly bounded on the set \(\{z \in {\mathbb {C}}_+ \, : \, |z| \ge \theta , |1-z| \ge \theta \}\), and thus locally bounded on \({\mathbb {C}}_+\).

It was established in Step 1 above that for any \(\gamma >0\), \(E\in {\mathbb {R}}\), \(\phi \in (0,1)\) and \(\epsilon >0\) small enough the solution to Eq. (3.22) with positive semidefinite imaginary part can be explicitly given as \({\widetilde{M}}_3(\sqrt{z})\) [see (3.45)] in the neighborhood of the origin, i.e., on the set \(\{|z|\le \delta , \mathrm {Im}z >0\}\) for \(\delta > 0\) sufficiently small. Moreover, Step 1 shows that for any \(z\in \{|z|\le \delta , \mathrm {Im}z >0\}\), there exists a well-defined limit \(\lim _{(\phi _{n},\epsilon _{n})\rightarrow (\phi , 0)} {\widetilde{M}}_3(\sqrt{z})\). Together with (3.19) and (3.20), this implies that on the set \(\{|z|\le \delta , \mathrm {Im}z >0\}\) the limit \(M_{z,E}:= \lim _{(\phi _{n},\epsilon _{n})\rightarrow (\phi , 0)} M_{z,E+\mathrm {i}\,\epsilon _n \phi _{n}} \) exists as well.

Combining the above information, we see that for any index pair (ij), \(\{M_{z,E+\mathrm {i}\,\epsilon _n \phi _{n}}\,(i,j)\}_{n\ge 1}\) is a family of analytic functions, locally bounded on \({\mathbb {C}}_+\) that converges on \(\{|z|\le \delta , \mathrm {Im}z >0\}\) to \(M_{z,E}(i,j)\). Applying the Vitali–Porter theorem (see, e.g., Section 2.4 in [34]), we conclude that for any \(z\in {\mathbb {C}}_+\) the limit \(M_{z,E}\,(i,j):=\lim _{(\phi _{n},\epsilon _{n})\rightarrow (\phi , 0) }M_{z,E+\mathrm {i}\,\epsilon _n \phi _{ n }}\,(i,j)\) exists, the convergence holds uniformly on the compact subsets of \({\mathbb {C}}_+\) and, as a result, the function \(z\mapsto M_{z,E}\,(i,j)\) is analytic on \({\mathbb {C}}_+\). Taking \(M_{z,E}\,(i,j)\) as the entries of the matrix-valued function \(M_{z,E}\) defines the solution to the Dyson equation (3.21) for \(\phi \in (0,1)\) and \(w= E \in {\mathbb {R}}\).

To compute the asymptotic behavior of \(M_{z,E}(1,1)\), we use (3.45) with \(\epsilon =0\) [see (3.48)], (3.20) and \(t = \sqrt{z}\) to find \(M_1\) and its upper left corner element \(M_{z,E}(1,1)\) in the neighborhood of \(z=0\)

$$\begin{aligned} M_{z,E}(1,1) = \mathrm {i}\frac{4+ \gamma ^2 \nu _0^2+ \gamma ^2 \xi _0^2}{4 \gamma \xi _0} \, z^{-1/2} + O(1). \end{aligned}$$

Step 3: Expansion around \(z=1\). We apply exactly the same procedure as in Step 2 of Lemma 2.10, just we insert the parameters \(\phi \) and \(\gamma \) into the identities (2.83) and (2.84) and follow them through the analysis. Here, we record the final result of this elementary calculation. Our ansatz is

$$\begin{aligned} {\widetilde{M}}_3 = \left( \begin{array}{cc} \frac{\xi }{4 \gamma ^2t}&{}\quad \frac{E}{2}+\frac{\nu t}{4 \gamma } - \frac{\mathrm {i}\xi }{2 \gamma }(t +t^{-1}) \\ \frac{E}{2}+\frac{\nu t}{4 \gamma } + \frac{\mathrm {i}\xi }{2 \gamma }(t +t^{-1}) &{} \quad \xi (t +t^{-1}) \end{array} \right) . \end{aligned}$$
(3.51)

The expansion in \(t=\sqrt{z-1}\) gives that for all \(|E| \le E_0\) with

$$\begin{aligned} E_0 := \frac{\sqrt{2}}{ \gamma }\Big ( \gamma ^2 (1-2\phi ) -1+\sqrt{ 1 + \gamma ^4 (1 - 2 \phi )^2 + 2 \gamma ^2 (1+2\phi )}\Big )^{1/2} \end{aligned}$$
(3.52)

the upper-left component of \(M_{z,E}\) is given by

$$\begin{aligned} M(1,1) = \frac{4\xi _0}{(\xi _0^2 + \gamma ^2 E^2+4 )t} + O(1), \end{aligned}$$
(3.53)

where \(\xi _{0}\) is defined by

$$\begin{aligned} \xi _0 = - \gamma \sqrt{E_0^2-E^2} . \end{aligned}$$
(3.54)

This finishes the proof of the lemma. \(\square \)

3.5 Explicit Solution for \(\phi \rightarrow 0\)

Lemma 3.9

Let \(\gamma >0\), \(w= E \in {\mathbb {R}}\) and \(\phi = 0\). Then, the Dyson equation (3.21)–(3.22) admits a solution \({\mathbb {C}}_{+}\ni z\mapsto M_{z,E} \in {\mathbb {C}}^{8\times 8}\) with \(\mathrm {Im}M_{z,E}\ge 0\) and the upper-left entry is explicitly given by

$$\begin{aligned} M_{z,E}(1,1) = \frac{ \gamma ^2 (4-E^2)-(1+ \gamma ^2 )^2+\mathrm {i}\gamma (1+ \gamma ^2 )}{ \gamma ^2 (4-E^2)+((1+ \gamma ^2 )^2-(4-E^2) \gamma ^2 )z} \sqrt{\frac{4-E^2}{z(1-z)}}. \end{aligned}$$
(3.55)

Moreover, this solution can be continuously extended to the set \(\phi \in [0,\phi _{0}]\), \(|1-z|\ge \theta \), \(\theta \le |z| \le \theta ^{-1}\) for \(E\in {\mathbb {R}}\) and sufficiently small \(\theta >0\) and \(\phi _0 = \phi _0(\theta ) >0\).

Proof

At \(\epsilon =0\) and setting \(\phi = 0\), the Eq. (3.22) simplifies to a quadratic matrix equation for \(M_3 \sigma _1\), where \(M_3\) satisfies the symmetry constraints (3.24) and (3.25), i.e.,

$$\begin{aligned} -\frac{1}{M_{3}} = -E\,\sigma _{1} + \sigma _{1} M_{3} \sigma _{1} , \qquad M_3 = \left( \begin{array}{cc} \frac{\mathrm {i}\xi }{4 \gamma } &{}\quad \frac{\nu }{2}+\frac{ z \xi }{2} \\ \frac{\nu }{2}-\frac{ z \xi }{2} &{}\quad z \gamma \mathrm {i}\xi \end{array} \right) , \end{aligned}$$
(3.56)

for two functions \(\xi \) and \(\nu \) that are easily computed to be \(\nu =E\) and

$$\begin{aligned} \xi = \sqrt{\frac{4-E^2}{z(1-z)}}. \end{aligned}$$
(3.57)

The choice of root is determined by \(\mathrm {Re}\xi >0\). Inserting \(M_{3}\) into (3.20) leads to

$$\begin{aligned} M_{z,E}(1,1) = \frac{ \gamma ^2 (4-E^2)-(1+ \gamma ^2 )^2+\mathrm {i}\gamma (1+ \gamma ^2 ) \xi }{ \gamma ^2 (4-E^2)+((1+ \gamma ^2 )^2-(4-E^2) \gamma ^2 )z}. \end{aligned}$$
(3.58)

We now construct the solution of (3.22) perturbatively for \(\phi \in (0,\phi _0)\) with some sufficiently small \(\phi _0=\phi _0(\theta )>0\) with \(|1-z|\ge \theta \) and \(\theta \le |z| \le \theta ^{-1}\). Recall that (3.22) is equivalent to \(\Delta =0\) with \(\Delta \) defined as in (3.43). In particular,

$$\begin{aligned} \Delta ={\widetilde{\Delta }} (Z_1+M_{3})\frac{1}{Z_1-Z_2}(Z_2+M_{3}) , \end{aligned}$$
(3.59)

implicitly defining \({\widetilde{\Delta }}\) that satisfies

$$\begin{aligned} {\widetilde{\Delta }}|_{\phi =0}=1+ M_{3}( \sigma _1M_{3}\sigma _1- \sigma _1E ). \end{aligned}$$
(3.60)

Similarly as we did in the proof of Lemma 3.8 instead of considering the equation \(\Delta =0\) for a solution \(M_{3} \in {\mathbb {C}}^{2\times 2}\), we can equivalently consider it as an equation for the two unknown functions \(\xi \) and \(\nu \) from the ansatz (3.56) for \(M_{3}\). Since clearly the factors \(Z_1+M_{3}\) and \(Z_2+M_{3}\) in (3.59) have bounded inverses when \(M_{3}\) is the explicit solution at \(\phi =0\) from (3.56) with (3.57), we can equivalently consider \({\widetilde{\Delta }}=0\) as the equation for \(\xi \) and \(\nu \). Plugging the ansatz (3.56) into (3.60), we see that \({\widetilde{\Delta }}_{11}=0\) and \({\widetilde{\Delta }}_{12}=0\) already imply \({\widetilde{\Delta }}=0\) and that

$$\begin{aligned} {\widetilde{\Delta }}_{11}|_{\phi =0}= & {} \frac{1}{4} (4 + \nu ^2 + 2 z \nu \xi - z \xi ^2 + z^2 \xi ^2 - 2 E (\nu + z \xi )),\\ {\widetilde{\Delta }}_{12}|_{\phi =0}= & {} -\frac{\mathrm {i}(E - \nu ) \xi }{4 \gamma }. \end{aligned}$$

We compute the determinant of the Jacobian of the function \((\xi ,\nu )\rightarrow ( {\widetilde{\Delta }}_{11},{\widetilde{\Delta }}_{12})\) to be

$$\begin{aligned} \det \nabla _{\xi ,\nu }({\widetilde{\Delta }}_{11},{\widetilde{\Delta }}_{12})|_{\phi =0,\nu =E} = \frac{\mathrm {i}(z-1)z\xi ^2}{8\gamma }. \end{aligned}$$

Since \(\xi \) from (3.57) does not vanish we infer that (3.22) is linearly stable as an equation for \(\xi \), \(\nu \) for small enough parameters \(\phi \) in a vicinity of the explicit solution \(M_{3}\) from (3.56) with (3.57) for \(\phi =0\). \(\square \)

From (3.55), we read off the density of transmission eigenvalues \( \rho (\lambda )= \rho _{E, \phi = 0}(\lambda ) :=\frac{1}{\pi }\lim _{\eta \downarrow 0}\mathrm {Im}M_{\lambda + \mathrm {i}\eta ,E}(1,1)\). The corresponding Fano factor for \(\phi = 0\) is now computable as

$$\begin{aligned} F(E,\gamma ) = 1-\frac{\int \lambda ^2 \rho (\lambda ) \mathrm {d} \lambda }{\int \lambda \rho (\lambda ) \mathrm {d} \lambda } = \frac{1+ \gamma ^2 }{2(1+ \gamma ^2 + \gamma \sqrt{4-E^2})}. \end{aligned}$$
(3.61)

For \(\gamma =1\) and \(E=0\), we recover the density (1.5) and the Fano factor \(F=\frac{1}{4}\) obtained in [7], see Sect. 4 for more details.

3.6 Proof of Parts (i), (iii) and (iv) of Theorem 1.4

In this section, we collect the results established in Sects. 3.13.5 and complete the proof of Theorem 1.4. Recall that Theorem 1.3 was proven in Lemma 2.5 for \(\phi = 1\) and Lemma 3.4 for \(\phi \in (0,1)\).

Proof of part (i) of Theorem 1.4. The extension of \(\rho _{w,\phi }(d\lambda )\) to \(w= E \in {\mathbb {R}}\) for \(\phi \in {(}0,1/2{]}\cap {\mathbb {Q}}\) as well as the limit (1.12) and the extension of \(\rho _{E,\phi }(d\lambda )\) to irrational \(\phi \in (0,1)\) follows from the equivalence between the weak convergence of measures defined by (3.16) and the pointwise convergence of \(M_{z,w}\) established in Lemma 3.4 and part (a) of Lemma 3.8 for the corresponding limits.

The weak limit \(\lim _{\phi \downarrow 0}\rho _{E,\phi }(d\lambda ) = \rho _{E,\phi =0}(\lambda )d\lambda \) follows from the continuity of \( \phi \mapsto M_{z,E}\) at \(\phi = 0\) for all \(z\in {\mathbb {C}}_{+}\), which was established in Lemma 3.8 in the regimes \(|z|\le \theta \), \(|1-z|\le \theta \) and in Lemma 3.9 in the complementary regimes \(|z|\ge \theta \), \(\theta \le |1-z|\le \theta ^{-1}\). Since the Stieltjes transform of \(\rho _{E,\phi }\) is given by \(M_{z,E}(1,1)\), the exact expression for \(\rho _{E,0}\) can be derived as an inverse Stieltjes transform of \(M_{z,E}\) from (3.55).

Similarly as for the case \(\phi =1\) in Sect. 2.5, Lemma 2.11 and the integrability of \(M_{z,E}\) that can be deduced from Lemmas 3.7 and 3.8 yield the absolute continuity of \(\rho _{w,\phi }(d\lambda )=\rho _{w,\phi }(\lambda ) d\lambda \). \(\square \)

Proof of part (iii) of Theorem 1.4. Follows from the asymptotic behavior of \(M_{z,E}\) near \(z=0\) and \(z=1\) (3.41)–(3.42) established in Lemma 3.8, block structure of \(M_{z,E}\) (3.19) and the definition of the self-consistent density of states (3.18). \(\square \)

Proof of part (iv) of Theorem 1.4. Follows from the explicit formula for \(M_{z,E}\) in the regime \(\phi \rightarrow 0\) (3.58) established in Sect. 3.5, the block structure of \(M_{z,E}\) (3.19) and the definition of the self-consistent density of states (3.18). \(\square \)

4 Comparison with the Results of Beenakker and Brouwer

Consider the scattering matrix (1.9)

$$\begin{aligned} S(E) := I - 2 \gamma \mathrm {i}\,W^* (E\cdot I - H + \mathrm {i}\,\gamma W W^*)^{-1} W \in {\mathbb {C}}^{N_0\times N_0} \end{aligned}$$
(4.1)

with \(w= E \in {\mathbb {R}}\) in the regime \(\phi = N/M \rightarrow 0\) as \(M\rightarrow \infty \) with \(N_0 = 2N\). This model was studied in the case of the Gaussian entries by Beenakker and Brouwer in [7, 10], and one of the remarkable results of their theory is that in the experimentally relevant setting of the ideal coupling the limiting transmission eigenvalue density is given by the arcsine law (1.5) (see [8, Eq. (3.12)]. The ideal coupling assumption is formulated in terms of the matrix S(E) having zero mean [8, Eq. (3.8)]. Below we show that in the regime \(\phi \rightarrow 0\) the assumption \({\mathbb {E}}[S(E)] = 0\) is equivalent to \(E=0\) and \(\gamma = 1\). By plugging these values into (1.19) and (3.61), we recover the arcsine distribution and the corresponding Fano factor \(F(0,1) = 1/4\).

Since the results of the current section do not affect the main outcomes of this paper and are meant to be of expository nature, we will keep the presentation rather informal, focusing only on the crucial steps and omitting the technical details.

For simplicity, we will assume in this section that H is Gaussian. Note that \(W\in {\mathbb {C}}^{M\times N_{0}}\), so

$$\begin{aligned} W = U \Gamma V^* \end{aligned}$$
(4.2)

with unitary matrices \(U\in {\mathbb {C}}^{M\times M}\) and \(V\in {\mathbb {C}}^{N_{0}\times N_{0}}\) and

$$\begin{aligned} \Gamma = \left( \begin{array}{c} {\widetilde{\Gamma }} \\ 0 \end{array} \right) , \quad {\widetilde{\Gamma }} = \mathrm {diag}(\gamma _{1},\ldots , \gamma _{N_0}) , \end{aligned}$$
(4.3)

where \(\gamma _{i}\) are the singular values of W and \(N_0 \ll M\).

Note that \(N_{0} = 2 N = 2 \phi M\), and thus, the eigenvalues of \(W^*W\) have Marchenko–Pastur distribution with parameter \(2\phi \). For \(\phi \rightarrow 0\), regime that we are interested in, the eigenvalues of \(W^*W\) will be concentrated around point 1, in the neighborhood of size \(O(\sqrt{\phi })\), so for simplicity of presentation we will omit the asymptotically small term \(O(\sqrt{\phi })\) and assume throughout these computations that all \(\gamma _i=1\), i.e., \({\widetilde{\Gamma }} = I_{N_0}\).

After applying the singular value decomposition to W and factoring out matrices U and \(U^*\) from the inverse, we get

$$\begin{aligned} S(E)&= I - 2 \gamma \mathrm {i}\,V \Gamma ^t U^* (E\cdot I - H + \mathrm {i}\,\gamma U \Gamma \Gamma ^{t} U^*)^{-1} U \Gamma V^* \end{aligned}$$
(4.4)
$$\begin{aligned}&= I - 2 \gamma \mathrm {i}\,V ({\widetilde{\Gamma }}^t, 0) \left( E\cdot I - {\widetilde{H}} + \mathrm {i}\,\gamma \left( \begin{array}{cc} {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t}&{}\quad 0 \\ 0 &{}\quad 0_{(M-N_{0})\times (M-N_0) } \end{array} \right) \right) ^{-1} \left( \begin{array}{c} {\widetilde{\Gamma }} \\ 0 \end{array}\right) V^* \end{aligned}$$
(4.5)
$$\begin{aligned}&\!=\! I \!+\! 2 \gamma \mathrm {i}\,V ({\widetilde{\Gamma }}^t, 0) \left( {\widetilde{H}} -E\cdot I - \mathrm {i}\,\gamma \left( \begin{array}{cc} {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t}&{}\quad 0 \\ 0 &{}\quad 0_{(M-N_{0})\times (M-N_0) } \end{array} \right) \right) ^{-1} \left( \begin{array}{c} {\widetilde{\Gamma }} \\ 0 \end{array}\right) V^* , \end{aligned}$$
(4.6)

where \({\widetilde{H}}\) remains a GUE matrix. Separate the upper-left \(N_0 \times N_0\) block of \({\widetilde{H}}\)

$$\begin{aligned} {\widetilde{H}} = \left( \begin{array}{cc} {\widetilde{H}}_{1}&{}\quad {\widetilde{H}}_{2} \\ {\widetilde{H}}_{2}^{*} &{}\quad {\widetilde{H}}_{3} \end{array} \right) \end{aligned}$$
(4.7)

and note that \(\frac{1}{2\phi }{\widetilde{H}}_{1}\) and \( \frac{1}{1-2\phi }{\widetilde{H}}_{3}\) are both independent GUE matrices. Now the inverse matrix in (4.6) can be rewritten as

$$\begin{aligned} \left( \begin{array}{cc} {\widetilde{H}}_{1} -E - \mathrm {i}\,\gamma {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t}&{}\quad {\widetilde{H}}_{2} \\ {\widetilde{H}}_{2}^* &{}\quad {\widetilde{H}}_{3} -E \end{array} \right) ^{-1} . \end{aligned}$$
(4.8)

Using the Schur complement formula we have that the upper-left \(N_{0}\times N_{0}\) block of (4.8), the only part that does not vanish after sandwiching (4.8) by \(({\widetilde{\Gamma }}^t, 0)\) and its transpose, is given by

$$\begin{aligned} \left( {\widetilde{H}}_{1} -E - \mathrm {i}\,\gamma {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t} - {\widetilde{H}}_{2} \Big ({\widetilde{H}}_{3} -E\Big )^{-1} {\widetilde{H}}_{2}^* \right) ^{-1} . \end{aligned}$$
(4.9)

The semicircular law for the Hermitian (GUE) matrix \(\frac{1}{1-2\phi }{\widetilde{H}}_{3}\) implies that

$$\begin{aligned} ({\widetilde{H}}_{3} -E)^{-1} \!=\! \frac{1}{1-2\phi }\Big (\frac{1}{1-2\phi }{\widetilde{H}}_{3} \!-\!\frac{1}{1-2\phi }E\Big )^{-1} \!\approx \! \frac{1}{1\!-\!2\phi }m_{sc}\Big (\frac{1}{1\!-\!2\phi }E\Big ) I_{M-N_{0}} , \end{aligned}$$
(4.10)

as \( M\rightarrow \infty \), where \(m_{sc}(z)\) denotes the Stieltjes transform of the semicircular distribution and “\(\approx \)” denotes that the corresponding equality holds asymptotically with a vanishing additive term and with high probability. Note that random matrices \({\widetilde{H}}_{1}\), \({\widetilde{H}}_{2}\) and \({\widetilde{H}}_{3}\) are independent. Therefore, by the concentration for quadratic forms (see, e.g., [15, Theorem C.1]) can be approximated by

$$\begin{aligned} {\widetilde{H}}_{2} \Big ({\widetilde{H}}_{3} -E \Big )^{-1} {\widetilde{H}}_{2}^* \approx m_{sc}\Big (\frac{1}{1-2\phi }E\Big ) I_{N_{0}} . \end{aligned}$$
(4.11)

From (4.9) and (4.11), it remains to check the limiting behavior of

$$\begin{aligned} \left( {\widetilde{H}}_{1} -E - \mathrm {i}\,\gamma {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t} - m_{sc}\Big (\frac{1}{1-2\phi }E\Big ) \right) ^{-1} . \end{aligned}$$
(4.12)

Recall that since \(\phi \) is small, we assumed that \({\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t} = I_{N_0}\). In this case, using again the semicircular law for the Hermitian matrix \(\frac{1}{2\phi }{\widetilde{H}}_{1}\), we have

$$\begin{aligned}&\left( {\widetilde{H}}_{1} -E - \mathrm {i}\,\gamma {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t} - m_{sc}\Big (\frac{1}{1-2\phi }E\Big ) \right) ^{-1} \nonumber \\&\quad \approx \left( {\widetilde{H}}_{1} -E - \mathrm {i}\,\gamma - m_{sc}\Big (\frac{1}{1-2\phi }E\Big ) \right) ^{-1} \end{aligned}$$
(4.13)
$$\begin{aligned}&\quad = \frac{1}{2\phi } \left( \frac{1}{2\phi }{\widetilde{H}}_{1} -\frac{1}{2\phi }\Big (E + \mathrm {i}\,\gamma + m_{sc}\Big (\frac{1}{1-2\phi }E\Big )\Big ) \right) ^{-1} \end{aligned}$$
(4.14)
$$\begin{aligned}&\quad \approx \frac{1}{2\phi } m_{sc}\left( \frac{1}{2\phi }\Big (E + \mathrm {i}\,\gamma + m_{sc}\Big (\frac{1}{1-2\phi }E\Big )\Big ) \right) . \end{aligned}$$
(4.15)

From the asymptotics \(z m_{sc}(z) \rightarrow -1, |z|\rightarrow \infty \), we get that

$$\begin{aligned} \left( {\widetilde{H}}_{1} -E - \mathrm {i}\,\gamma {\widetilde{\Gamma }}{\widetilde{\Gamma }}^{t} - m_{sc}\Big (\frac{1}{1-2\phi }E\Big ) \right) ^{-1} \approx -\frac{1}{E + \mathrm {i}\,\gamma + m_{sc}(E)} ,\quad \phi \rightarrow 0 . \end{aligned}$$
(4.16)

We conclude that as \(N_{0}, M \rightarrow \infty \), \(\phi \rightarrow 0\) (see (4.6) and (4.16))

$$\begin{aligned} {\mathbb {E}}[S(E)] \approx 1+ 2 \gamma \mathrm {i}\,\left( -\frac{1}{E + \mathrm {i}\,\gamma + m_{sc}(E)} \right) . \end{aligned}$$
(4.17)

Now, from \(m_{sc}(0) = \mathrm {i}\,\) we have that

$$\begin{aligned} {\mathbb {E}}[S(0)] \approx 1- \frac{2 \gamma \mathrm {i}\,}{\mathrm {i}\,\gamma + \mathrm {i}\,} . \end{aligned}$$
(4.18)

Finally, taking \(\gamma = 1\) gives \( {\mathbb {E}}[S(0)] \approx 0\).

Note that from (4.17) we have that in the limit \(\phi \rightarrow 0\)

$$\begin{aligned} {\mathbb {E}}[S(0)] = 0 \quad \Leftrightarrow \quad E + m_{sc}(E) = \mathrm {i}\,\gamma , \end{aligned}$$
(4.19)

and for \(E\in {\mathbb {R}}\), the expression \(E + m_{sc}(E)\) is purely imaginary if and only if \(E=0\)

$$\begin{aligned} E + m_{sc}(E) = E+\frac{-E + \sqrt{E^2-4}}{2} = \frac{E + \sqrt{E^2-4}}{2} . \end{aligned}$$
(4.20)

Therefore, for \(\phi \rightarrow 0\), \({\mathbb {E}}[S(E)]=0\) if and only if \(E=0\) and \(\gamma = 1\).