Skip to main content

MINI REVIEW article

Front. Appl. Math. Stat., 02 December 2024
Sec. Mathematical Physics
This article is part of the Research Topic Quasi-Normal Modes, Non-Selfadjoint Operators and Pseudospectrum: an Interdisciplinary Approach View all 11 articles

Pseudospectra and eigenvalue asymptotics for disordered non-selfadjoint operators in the semiclassical limit

  • Institut de Recherche Mathématique Avancée - UMR 7501, Université de Strasbourg et CNRS, Strasbourg, France

The purpose of this note is to review certain recent results concerning the pseudospectra and the eigenvalues asymptotics of non-selfadjoint semiclassical pseudo-differential operators subject to small random perturbations.

1 Introduction

The spectral theory of non-selfadjoint operators acting on a Hilbert space is an established and highly developed subject. Non-selfadjoint operators are prevalent naturally in a wide range of modern problems. For instance, in the field of quantum mechanics, the study of scattering systems naturally leads to the notion of quantum resonances. These can be described as the complex values of the meromorphic. The continuation of the scattering matrix or of the cut-off resolvent of the Hamiltonian to the non-physical sheet of the complex plane. Alternatively, through a complex deformation of the initial Hamiltonian, these resonances can be characterized as the genuine complex valued eigenvalues of a non-selfadjoint operator [1, 3, 60]. We recommend the reader to reference [18] for an in-depth discussion of the mathematics of scattering poles. Another aspect of quantum mechanics is the examination of a small system that is linked to a larger environment. The effective dynamics of the small systems are governed by a non-selfadjoint operator: the Lindbladian [39].

A major obstacle to the spectral analysis of non-selfadjoint operators is the possible strong spectral instability of their spectrum with respect to small perturbations. This phenomenon, sometimes referred to as the pseudospectral effect, was initially considered to be a drawback, as it could lead to the origin of immense numerical errors, see Embree and Trefethen [19] and the references therein. However, a recent line of research has also demonstrated that the pseudospectral effect can provide novel insights into the spectral distribution of non-selfadjoint operators that are subjected to small generic perturbations.

2 Spectral instability of non-selfadjoint operators

We commence by recalling the definition of the pseudospectrum of a linear operator, a crucial concept that which quantifies its spectral instability. This notion appears to have originated in the second half of the 20th century in various contexts, see reference [65] for a historic overview. It quickly became an important notion in numerical analysis, as it allows us to quantify how much eigenvalues can spread out under the influence of small perturbations, see references [64, 65] and the book [19]. We follow here the latter reference.

Let H be a complex Hilbert space (assumed separable for simplicity) with norm ||·|| and scalar product (·|·). Let P:HH be a closed densely defined linear operator, with resolvent set ρ(P) and spectrum Spec(P) = ℂ\ρ(P).

Definition 1. For any ε > 0, we define the ε-pseudospectrum of P by

Specε(P):=Spec(P){zρ(P);||(P-z)-1||>ε-1}.    (1)

We note that some authors define the ε-pseudospectrum with a ⩾ rather than a >. We, however, follow here reference [19]. It is noteworthy that with this choice of non-strict inequality results in the Specε(P) being an open set in ℂ.

For P selfadjoint (or even normal), the spectral theorem implies that

Specε(P)Spec(P)+D(0,ε),    (2)

where D(0, ε) ⊂ ℂ denotes the open disk with radius ε centered at 0. For P non-selfadjoint, the pseudospectrum of P can be much larger, as illustrated by the following example.

Example 1. For N ≫ 1, consider the Jordan block matrix

PN=(01000010100):NN.    (3)

The spectrum of PN is given by {0}. Consider the vector e+=(1,z,,zN-1), |z| ⩽ r < 1. Then,

||(PN-z)e+||=|z|N=O(e-N|logr|)||e+||.

So, Theorem 2 shows that for any ε > 0 and any r ∈ ]0, 1[ we have that for N > 1 sufficiently large

D(0,r)Specε(PN).

An immediate consequence of Equation 1 is the property that pseudospectra are nested. More precisely,

Specε2(P)Specε1(P),  ε1>ε2>0.    (4)

The set (Equation 1) describes a region of spectral instability of the operator P, since any point in the ε-pseudospectrum of P lies within the spectrum of a certain ε-perturbation of P [19].

Theorem 1. Let ε > 0. Then

Specε(P)=QB(H)Q<1Spec(P+εQ).    (5)

Proof. See reference [19, p. 31].

A third, equivalent definition of the ε-pseudospectrum of P is provided by the existence of approximate solutions to the eigenvalue problem (Pz)u = 0.

Theorem 2. Let ε > 0 and z ∈ ℂ. Then the following statements are equivalent:

1. z ∈ Specε(P);

2. z ∈ Spec(P) or there exists a uzD(P) such that ||(Pz)uz|| < ε||uz||, where D(P) denotes the domain of P.

Proof. See reference [19, p. 31].

Such a state uz is referred to as an ε-quasimode, or simply a quasimode of Pz.

Spectral instability of semiclassical pseudo-differential operators

Although the notion of ε-pseudospectrum defined in Definition 1 is valid in the context of semiclassical pseudo-differential operators, we present here a somewhat different, but still related notion, which is more suited to the semiclassical setting. Here, the term “semiclassical” implies that our operators are dependent on a parameter h ∈ ]0, 1] (often referred to as “Planck's parameter”), and that our focus we will be on the asymptotic (semiclassical) regime h ↘ 0. This small parameter will provide us with a natural threshold for defining the pseudospectrum, and thereby measuring the spectral instability. The following discussion is based on the studies of Davies [13] and Dencker et al. [16].

Let d ⩾ 1 and h ∈ ]0, 1]. An order function mC(2d;[1,[), is a function satisfying the following growth condition:

C01, N0>0:  m(ρ)C0ρ-μN0m(μ),  ρ,μ2d,    (6)

where ρ-μ:=1+|ρ-μ|2 denotes the “Japanese brackets.” We will also sometimes write (x, ξ) = ρ ∈ ℝ2d, so that ξ ∈ ℝd. To such an order function m, we may associate a semiclassical symbol class [17, 71]. We assert that a smooth function pC(ρ2d,]0,1]h) belongs to the symbol class S(m) if for any multiindex α ∈ ℕ2d when there exists a constant Cα > 0 such that

|ραp(ρ;h)|Cαm(ρ),       ρ2d,h]0,1].    (7)

We recommend the reader for further reading on semiclassical analysis to [17, 41, 71].

Let the symbol pS(m), m ⩾ 1, be a “classical” symbol, which satisfies an asymptotic expansion in the limit h → 0:

p(ρ;h)~p0(ρ)+hp1(ρ)+  in   S(m),    (8)

where each pjS(m) is independent of h. We assume that there exists a z0 ∈ ℂ and a C0 > 0 such that

|p0(ρ)-z0|m(ρ)/C0,  ρT*d.    (9)

Here, T*d ≃ ℝ2d denotes the cotangent space of ℝd. In this case, we call p0 the (semiclassical) principal symbol of p. We then define two subsets of ℂ associated with p0:

Σ:=Σ(p0):=p0(T*d)¯,Σ:={zΣ; (ρj)j1s.t.|ρj|,p0(ρj)z}.    (10)

Here, the p0(T*d)¯ denotes the closure of the set p0(T*d), and we will use this notation in the sequel. The set Σ is the classical spectrum, and Σ can be called the classical spectrum at infinity of the h-Weyl quantization of p was defined by

Phu(x):=pw(x,hDx)u(x)                 =1(2πh)deih(x-y)·ξp(x+y2,ξ;h)u(y)dydξ,                          uS(d),    (11)

seen as an oscillatory integral in ξ. The operator Ph maps SS, and by duality S′ → S′, continuously.

3.1 Semiclassical pseudospectrum

Similar to Dencker et al. [16], we define for a symbol pS(m) as in Equation 8 the sets

Λ±(p):={p(ρ); ±12i{p¯,p}(ρ)<0}Σ,    (12)

where {·, ·} denotes the Poisson bracket. It should be noted that the condition 12i{p¯,p}0 is the classical analog of the [Ph*,Ph]0. As in Dencker et al. [16], we call the set

Λ(p):=Λ-Λ+¯    (13)

the semiclassical pseudospectrum.

Theorem 3 ([16]). Suppose that n ⩾ 2, Cb(T*d)p~p0+hp1+ , and p0-1(z) is compact for a dense set of values z ∈ ℂ. If Ph=pw(x,hDx), then

Λ(p0)\ΣΛ+(p0)¯

and for every z ∈ Λ+(p0) and every ρ0T*d with

p0(ρ0)=z,  12i{p¯0,p0}(ρ0)<0,

there exists 0e+L2(d) such that

||(Ph-z)e+||=O(h)||e+||,  WFh(e+)={ρ0}.    (14)

1

If, in addition, p has a bounded holomorphic continuation to to {ρ ∈ ℂ2d, |Imρ| ⩽ 1/C}, then Equation 14 holds with the h replaced by exp(−1/(Ch)).

If n = 1, then the same conclusion holds, provided that in addition to the general assumptions, each component of ℂ\Σ has a nonempty intersection with ∁Λ(p).2

This result can be extended to unbounded symbols pS(T*d, m), as shown in Equation 8, and the corresponding operators Ph with principal symbol p0, by applying Theorem 3 to P~h=(Ph-z0)-1(Ph-z), with principal symbol p~0Cb(T*d) and z0 as in Equation 9 and z0z. Indeed, note that z ∈ Σ(p0) if and only if 0Σ(p~0), and that ρp0-1(z) with ±{Rep0, Imp0}(ρ) < 0 is equivalent to ρp~0-1(0) with ±{Rep~0,Imp~0}(ρ)<0. Furthermore, a quasimode u as in Theorem 3 for P~h then provides, after a possible truncation, a quasimode for Phz in the same sense.

By replacing Ph with its formal adjoint, Ph*, and thus p with p¯, Theorem 3 yields that for every z ∈ Λ(p) and every ρ0T*d with

p0(ρ0)=z,  12i{p¯0,p0}(ρ0)>0,

there exists 0e-L2(d) such that

||(Ph-z)*e-||=O(h)||e-||,  WFh(e-)={ρ0}.

The additional statements of Theorem 3 regarding symbols that permit a holomorphic extension to a complex neighborhood of ℝ2d, and the case where n = 1 hold as well.

Example 2. The case study to be considered is the case of the non-selfadjoint Harmonic oscillator

Ph=(hDx)2+ix2

is seen as an unbounded operator L2(ℝ) → L2(ℝ). The principal symbol for Ph is given by p(x, ξ) = ξ2 + ix2S(T*ℝ, m), with a weight function m(x, ξ) = 1 + ξ2 + x2. We equip Ph with the domain H(m):=(Ph+1)-1L2(), where the operator on the right is the pseudo-differential inverse of Ph + 1. This choice of domain renders Ph a closed and densely defined operator. Using, for instance, the method of complex scaling, it can be observed that the spectrum of Ph is determined by

Spec(Ph)={eiπ/4(2n+1)h;n}.    (15)

Furthermore, Σ is the closed first quadrant in the complex plane, whereas Σ = ∅. For ρ = (x, ξ) ∈ T*ℝ, we find that

12i{p¯,p}(x,ξ)=2ξ·x.    (16)

Thus, for every z°3 there exist points

ρ+j(z)=(-1)j(-|Rez|,|Imz|),ρ-j(z)=(-1)j(-|Rez|,-|Imz||),  j=1,2,

such that

±12i{p¯,p}(ρ±j(z))<0,  j=1,2.

Using the WKB method, it is possible to construct quasimodes of the form e+j(x;h)=a+j(x;h)eiϕ+j(x)/h with a+j(x;h)Cc() admitting an asymptotic expansion a+j(x;h)~a+,0j(x)+ha+,1j(x)+  with WFh(e+j)={ρ+j(z)} and

||(Ph-z)e+j||=O(e-1/Ch),    (17)

see Davies [13, 14] for an explicit computation, and Dencker et al. [16] for a more general construction.

In fact, the works of Davies [13, 14] provide an explicit WKB construction for a quasimode u for one-dimensional non-selfadjoint Schrödinger operators Ph-z=(hDx)2+V(x)-z on L2(ℝ) with VC(ℝ) complex-valued and z = V(a)+η2, for some a ∈ ℝ, η > 0. Furthermore, one assumes that ImV′(a) ≠ 0. These studies served as the foundation for the quasimode construction of non-selfadjoint (pseudo-)differential operators. Zworski [69] compared Davies' quasimode construction under the condition on the gradient of ImV to a quasimode construction under a non-vanishing condition of the Poisson bracket 12i{p¯,p}. Furthermore, Zworski [69] established the link to the famous commutator condition of Hörmander [32, 33]. A full generalization of the quasimode construction under a non-vanishing condition of the poisson bracket, see Theorem 3, was then achieved by Dencker et al. [16]. Finally, Pravda-Starov [4648] improved these results by modifying a quasimode construction by Moyer and Hörmander, see reference [34, Lemma 26.4.14], for adjoints of operators that do not satisfy the Nirenberg-Tréves condition (Ψ) for local solvability.

For a quasimode construction for non-selfadjoint boundary value problems, we recommend the reader refer to the study of Galkowski [20].

It is noteworthy, that Equation 14 (or Equation 17 in the aforementioned example) implies that if the resolvent (Ph-z)-1 exists then its norm is larger than any power of h when h → 0, or even larger than e1/Ch in the analytical case. Each family (e+j(z,h)) is an h-quasimode of Phz, or for short a quasimode of Phz.

From the quasimode Equation 14, it is easy to observe an operator Q of unity norm and a parameter δ=O(h), such that the perturbed operator Ph + δQ has an eigenvalue at z. For instance, if we call the error r+ = (Phz)e+, we may take the rank 1 operator δQ=-r+(e+)*. According to Theorem 3, it can be observed that the interior of the set Λ(p), situated away from the set Σ, is a zone of strong spectral instability for Ph. For this reason, we may refer to the semiclassical pseudospectrum Λ(p) also as the (h-) pseudospectrum of Ph. Finally, we recommend the reader also to the refer studies of Pravda-Starov [4648] for further refinement of the notion of semiclassical pseudospectrum.

3.2 Outside the semiclassical pseudospectrum

When

z\Σ(p),

then by condition (Equation 9), we have (p0(ρ) − z) ⩾ m(ρ)/C for some sufficiently large C > 0 and so we know that the inverse (Ph-z)-1 is a pseudo-differential operator with principal symbol (p0-z)-1S(1/m)S(1). Hence, (Ph-z)-1 maps L2L2 and

||(Ph-z)-1||=O(1)    (18)

uniformly in h > 0. Therefore, from the semiclassical point of view, we may consider ℂ\Σ as a zone of spectral stability.

3.3 At the boundary of the semiclassical pseudospectrum

At the boundary of the semiclassical pseudospectrum, a transition occurs between the zone of strong spectral instability and stability. Indeed, at the boundary we find an improvement over the resolvent bounds, assuming some additional non-degeneracy:

Splitting a symbol pCb(T*d) into real and imaginary part, p = p1 + ip2, we consider the iterated Poisson bracket

pI:={pi1,{pi2,{,{pik-1,pik}} }}

where I ∈ {1, 2}k, and |I| = k is called the order of the Poisson bracket. The order of p at ρ ∈ T*d is given by

k(ρ):=max{j;pI(ρ)=0, 1<|I|j}.

The order of z0 ∈ Σ\Σ is the maximum of k(ρ) for ρp-1(z0).

Theorem 4. See Dencker et al. [16, 56] Assume that Cb(T*d)p~p0+hp1+ . Let Ph=pw(x,hDx) and let z0 ∈ ∂Σ(p0)\Σ(p0). Assume that dp0 ≠ 0 at every point in p0-1(z0), and that z0 has a finite order k ⩾ 1 for p. Then, k is equal and h > 0 is small enough for

||(Ph-z)-1||Ch-kk+1.

In particular, there exists a c0 > 0, such that h > 0 is small enough for

{z;|z-z0|c0hkk+1}Spec(Ph)=.

This result was proven in dimension 1 by Zworski [70], and in certain cases by Boulton [8]. Further refinements have been obtained from Sjöstrand [56]. Similar to the discussion after Theorem 3, we can extend Theorem 4 to unbounded symbols pS(T*d, m) and their corresponding quantizations.

Example 3. Recall the non-selfadjoint Harmonic oscillator Ph=(hDx)2+ix2 from Example 2. Here ∂Σ = ℝ+iℝ+, so we see by Equation 16 that for 0 ≠ z0 ∈ Σ

12i{p¯,p}(ρ)={Rep,Imp}(ρ)=0,  ρp-1(z0).

However,

either{Rep,{Rep,Imp}}(ρ)=4ξ20,or{Imp,{Rep,Imp}}(ρ)=-4x20,

indicating that z0 is of order 2 for p = ξ2 + ix2, and Theorem 4 reveals that

||(Ph-z0)-1||Ch-23.

In order for a the ε-pseudospectrum of Ph to reach the boundary of Σ, we require ε > h2/3/C.

3.4 Pseudospectra and random matrices

In this section, we present a brief discussion on pseudospectra for large N × N random matrices. One may interpret the 1/N, where N ≫ 1, as an analog to the semiclassical parameter. By recalling the example of the non-selfadjoint harmonic oscillator, as illustrated in Example 2, we see that pseudospectra can be very large in general. However, in a generic setting, they are typically much smaller.

Let M ∈ ℂN×N be a complex N × N matrix and let s1(M) ⩾ … ⩾ sN(M) ⩾ 0 denotes its singular values, which are the eigenvalues of M*M ordered in a decreasing manner and counting multiplicities. It should be noted that if Mz is bijective for some z ∈ ℂ, then

||(M-z)-1||=sN(M-z)-1.

In view of Equation 1, the ε-pseudospectrum of M is then characterized by the condition that z ∈ Specε(M)

zSpecε(M)    sN(M-z)<ε.

A classical result from Sankar et al. [51, Lemma 3.2] (stated there for real Gaussian random matrices) indicates that with a high probability, the smallest singular value of a deformed random matrix is not too small.

Theorem 5 ([51]). There exists a constant C > 0 such that the following holds true. Let N ⩾ 2, let X0 be an arbitrary complex N × N matrix, and let Q be an N × N complex Gaussian random matrix, whose entries are all independent copies of a complex Gaussian random variable q~N(0,1). Subsequently, for any δ > 0

P(sN(X0+δQ)<δt)CNt2.

Proof. For real matrices the proof can be found in Sankar et al. [51, Lemma 3.2], see also reference [63, Theorem 2.2]. For complex matrices a proof is presented for instance in Vogel [66, Appendix A].

Theorem 5 states us that any fixed z ∈ ℂ is not included in the ε-pseudospectrum of X + δQ with a probability ⩾ 1 − CNε2δ−2. This result suggests that the pseudospectrum of random matrices is typically not too large. Theorem 5 has received many extensions. For instance Rudelson and Vershynin [50] consider the case of random matrices with iid (independent and identically distributed) sub-Gaussian entries. Tao and Vu [62] consider iid entries with a nonzero variance. Cook [12] considers the case of random matrices whose of entries have an inhomogeneous variance profile under appropriate assumptions. We conclude this section by noting the following, quantitative outcome obtained by Tao and Vu.

Theorem 6 ([63]). Let q be a random variable with a mean zero and a bounded second moment, and let γ ⩾ 1/2, A ⩾ 0 be constants. Then, there exists a constant C > 0, depending on q, γ, and A such that the following holds true. Let Q be the random matrix of size N, whose entries are independent and identically distributed copies of q, and let X0 be a deterministic matrix satisfying ||X0||Nγ. Then,

P(sn(X0+Q)n-γ(2A+2)+1/2)C(n-A+o(1)+P(||Q||nγ)).    (19)

Example 4. Consider the case where q is a random variable satisfying the moment conditions

𝔼[q]=0,  𝔼[|q|2]=1,  𝔼[|q|4]<+.    (20)

Form [37] reveals that Equation 20 implies that 𝔼[||Q||] ⩽ CN1/2, which, using Markov's inequality, yields that for any ε > 0

P[QCN1/2+ε]C-1N-1/2-ε𝔼[||Q||]N-ε.    (21)

In this case (Equation 19) becomes

P(sn(X0+Q)n-(ε+1/2)(2A+2)+1/2)C(n-A+o(1)+N-ε).    (22)

4 Eigenvalue asymptotics for non-selfadjoint (random) operators

Consider the operator Ph=pw(x,hDx) depicted in Equations 8, 11, which is viewed as an unbounded operator L2(ℝd) → L2(ℝd). We equip Ph with the domain H(m):=(Ph-z0)-1L2(d). It should be noted that (Ph-z0)-1 exists for h > 0 that is sufficiently small by the elipticity condition (Equation 9). We will denote by ||u||m: = ||(Phz0)u|| the associated norm on H(m). Although this norm depends on the selection of the symbol p0z0, it is equivalent to the norm defined by any operator with an elliptic principal symbol qS(m), so that the space H(m) solely depends on the order function m. Since H(m) contains the Schwartz functions S(d), it is dense in L2(ℝd).

Let us verify that Ph equipped with domain H(m) is closed. Let (Phz0)ujv and uju in L2. Since (Ph-z0):H(m)L2 is bijective, it follows that uj(Ph-z0)-1v in H(m) and also in L2. So u=(Ph-z0)-1v. In summary, Ph equipped with the domain H(m) is a densely defined closed linear operator.

Recall Equation 10, and let

Ω\Σ    (23)

be open, relatively compact, not entirely contained in Σ and so that Ω¯\Σ. Using the ellipticity assumption (Equation 9), it was proven in reference [25, Section 3] that

• Spec(Ph) ∩ Ω is discrete for h > 0 small enough,

• For all ε > 0 there exists an h(ε) > 0 such that

Spec(Ph)ΩΣ+D(0,ε),  0<hh(ε),

where D(0, ε) denotes the disc in ℂ of radius ε and centered at 0.

4.1 The selfadjoint setting

If Ph above is selfadjoint, which implies in particular that p is real-valued, we have the classical Weyl asymptotics. We follow here Dimassi and Sjöstrand [17] for a brief review.

Theorem 7. Let Ω be as in Equation 23. For every h-independent interval I ⊂ Ω ∩ ℝ with Vol2d(I)=0,

#(Spec(Ph)I)=1(2πh)d(p0-1(I)dxdξ+o(1)),  h0.    (24)

This result is, in increasing generality, attributed to Chazarin [10], Helffer and Robert [26, 27], Petkov and Robert [45] and Ivrii [35]. See also Dimassi and Sjöstrand [17] for an overview. We highlight two special cases: when I = [a, b], a < b, and a, b are not critical points of p0, then the error term becomes O(h), see Chazarin [10], Helffer-Robert [26], and Ivrii [35]. When additionally the unions of periodic Hp0 trajectories4 in the energy shell p0-1(a) and p0-1(b) are of the Liouville measure 0, then the error term is of the form

h(p0=ap1(ρ)La(dρ)-p0=bp1(ρ)Lb(dρ))+o(h),    (25)

where Lλ denotes the Liouville measure on p0-1(λ). See Petkov and Robert [45] and Ivrii [35] and Dimassi and Sjöstrand [17] for details. Let us also highlight that similar results obtained from Theorem 7 are also valid for compact smooth manifolds, see, for instance, Grigis and Sjöstrand [21, Chapter 12] and the references therein.

The corresponding results in the setting of self-adjoint partial differential operators in the high energy limit go back to the seminal study of Weyl [68] and have a long and very rich history. These are, however, beyond the scope of this review.

Example 5. The guiding example to keep in mind is the self-adjoint Harmonic oscillator

Ph=(hDx)2+x2:L2()L2()

seen as an unbounded operator. The principal symbol of Ph is represented by p(x, ξ) = ξ2 + x2S(T*ℝ, m), and the weight function m(x, ξ) = 1 + ξ2 + x2. Ph is represented by the domain H(m):=(Ph+1)-1L2(), where the operator on the right is the pseudo-differential inverse of Ph + 1. This choice of domain makes Ph a densely defined closed operator. It is widely acknowledged (see, for instance, reference [71, Theorem 6.2]) that the spectrum of Ph is determined by

Spec(Ph)={(2n+1)h;n}.

Counting the points (2n + 1)h contained in an interval [a, b], 0 ⩽ a < b < ∞, gives

#(Spec(Ph)[a,b])=b-a2h+O(1).

Since Vol2({aξ2+x2b})=π(b-a), we confirm Theorem 7 for the Harmonic oscillator.

4.2 The non-self-adjoint setting

The natural counterpart of Theorem 7 for non-self-adjoint operators would be eigenvalue asymptotics in a complex domain Ω ⋐ ℂ as in Equation 23. Recall the non-self-adjoint Harmonic oscillator Ph from Example 2 with principal symbol p(x, ξ) = ξ2 + ix2. In this case, Σ = {z ∈ ℂ; Rez, Imz ⩾ 0} and Σ = ∅. Any ∅ ≠ Ω ⋐ Σ away from the line eiπ/4+, indicates the view of Equation 15 that

#(Spec(Ph)Ω)=0.

On the other hand,

12πhp-1(Ω)dxdξ>0.

This example suggests that a direct generalization of Theorem 7 to non-self-adjoint operators with a complex valued principal symbol cannot hold.

Let us comment on two settings where a form of Weyl asymptotics is known to hold: Upon assuming analyticity, one may recover a sort of Weyl asymptotics. More precisely, as shown in the studies of Melin and Sjöstrand [43], Sjöstrand [53], Hitrik and Sjöstrand [2830], Hitrik et al. [31], and Rouby [49], the discrete spectrum of certain analytic non-self-adjoint pseudo-differential operators is confined to curves in Σ. Moreover, one can recover eigenvalue asymptotics using Bohr-Sommerfeld quantization conditions.

The second setting occurs when the non-self-adjointness of the operator Ph arises not from the principal symbol p0 (assumed to be real-valued), but from the subprincipal symbol p1. For instance, when studying the damped wave equation on a compact Riemannian manifold X, one is led to study the eigenvalues of the corresponding stationary operator

Ph(z)=-h2Δ+2iha(x)z,  aC(X;).

Here, Δ denotes the Laplace-Beltrami operator on X, and we call z ∈ ℂ an eigenvalue of Ph(z) if there exists a corresponding L2 function u is present in the kernel of Ph(z)−z. In fact, such a u is smooth by elliptic regularity. Using Fredholm theory, one can show that these eigenvalues form a discrete set in ℂ.

The principal part of Ph = Ph(z) is given by −h2Δ, and thus is self-adjoint. The principal symbol is p0(x,ξ)=|ξ|x2 (the norm here is with respect to the Riemannian metric on X). However, the subprincipal part is complex valued and non-self-adjoint.

Lebeau [38] has established that there exists a± ∈ ℝ, wherein for every ε > 0 there exist a finite number of eigenvalues such that

Imzh[a--ε,a++ε].

Remark 1. In fact Lebeau provided precise expressions for a± in terms of the infimum and the supremum over the co-sphere bundle S*X of the long time average of the damping function a evolved via the geodesic flow. Further refinements have been obtained by Sjöstrand [52], and when X is negatively curved by Anantharaman [2] and Jin [36].

Additionally, Markus and Matsaev [40] and Sjöstrand [52] have demonstrated the following analog of the Weyl law. For 0 < E1 < E2 < ∞ and for C > 0 sufficiently large

#(Spec(Ph)([E1,E2]+i[-Ch,Ch]))=1(2πh)d(p0-1([E1,E2])dxdξ+O(h)).    (26)

Finer results have been obtained by Anantharaman [2] and Jin [36] when X is negatively curved.

4.3 Probabilistic Weyl asymptotics

In a series of studies by Hager [2325] and Sjöstrand [54, 55], the authors proved a Weyl law, with overwhelming probability, for the eigenvalues in a compact set Ω ⋐ ℂ as in Equation 23 for randomly perturbed operators

Pδ=Ph+δQω,  0<δ=δ(h)1,    (27)

where Ph is as per in Section 3, and the random perturbation Qω is one of the following two types.

4.3.1 Random matrix

Let N(h) → ∞ sufficiently fast as h → 0. Let qj,k, 0 ⩽ j, k < N(h) be independent copies of a complex Gaussian random variable α~N(0,1). We consider the random matrix

Qω=0j,k<N(h)qj,kejek*,    (28)

where {ej}jL2(d) is an orthonormal basis and ejek*u=(u|ek)ej for uL2(ℝ). The condition on N(h) is determined by the requirement that the microsupport of the vectors in the orthonormal system {ej}j < N(h), “covers” the compact set p0-1(Ω)T*d, where p0 is the principal symbol of Ph. For instance, we could consider the first N(h) eigenfunctions (ordered according to increasing eigenvalues) of the Harmonic oscillator Ph=-h2Δ+x2 on ℝd. The number N(h) is then determined by the condition that the semiclassical wavefront sets of ej, jN(h), are disjoint from p0-1(Ω). Alternatively, as in Hager and Sjöstrand [25], one may also take N(h) = ∞; however, then one must conjugate Qω by suitable elliptic Hilbert–Schmidt operators. We recommend the reader to Hager and Sjöstrand [25] for further information.

4.3.2 Random potential

We take N(h) and an orthonormal family (ek)k ∈ ℕ as above. Let v be real or complex random vector in ℝN(h) or ℂN(h), respectively, with joint probability law

v*(dP)=Zh-11B(0,R)(v)eϕ(v)L(dv),    (29)

where Zh > 0 is a normalization constant, B(0, R) is either the real ball ⋐ℝN(h) or the complex ball ⋐ℂN(h) of radius R = R(h) ≫ 1, and centered at 0, L(dv) denotes the Lebesgue measure on either ℝN(h) or ℂN(h) and ϕ ∈ C1 with

||vϕ||=O(h-κ4)    (30)

uniformly, for an arbitrary but fixed value of κ4 ⩾ 0. In Hager [24] the case of non-compactly supported probability law was considered. More precisely, the entries of the random vector v were supposed to be independent and identically distributed (iid) complex Gaussian random variables ~N(0,1). In Sjöstrand [54, 55], the law Equation 29 was considered. For the sake of simplicity, we will not elaborate here the precise conditions on the ek, R(h), and N(h), in this case, but refer the reader to Sjöstrand [54, 55]. However, one example of a random vector v with law (Equation 30) is a truncated complex or real Gaussian random variables with expectation 0, and uniformly bounded covariances. In fact, the methods in Sjöstrand [54, 55] can be extended to non-compactly supported probability distributions, provided sufficient decay conditions at infinity are assumed. For instance, iid complex Gaussian random variables, as in the one dimensional case [24], are permissable. Finally, we conclude that the methods in Sjöstrand [54, 55] can probably also be modified to allow for the case of more general independent and identically distributed random variables. We define the random function as

Vω=0j<N(h)vjej.    (31)

We call this perturbation a “random potential,” even though Vω is complex valued. When we consider this type of perturbation, we will make the additional symmetry assumption:

p(x,ξ;h)=p(x,-ξ;h).    (32)

Let Ω ⋐ ℂ be an open simply connected set as in Equation 23. For z ∈ Ω and 0 ⩽ t ≪ 1 we set

Vz(t)=Vol{ρT*d;|p0(ρ)-z|2t}.    (33)

Let Γ ⋐ Ω be open with C2 boundary and make the following non-flatness assumption

κ]0,1], such that Vz(t)=O(tκ),uniformly for zneigh(Γ),  0t1.    (34)

The above mentioned works have yielded the following result.

Theorem 8 (Probabilistic Weyl's law). Let Ω be as in Equation 23. Let Γ ⋐ Ω be open with C2 boundary. Let Phδ be a randomly perturbed operators as in Equation 27 with e−1/Ch ≪ δ ⩽ hθ with θ > 0 sufficiently large. Then, in the limit h → 0,

#(Spec(Phδ)Γ)=1(2πh)d(p0-1(Γ)dxdξ+o(1))with probability1-Chη,    (35)

for some fixed η > 0.

The studies [2325, 54, 55] also provide an explicit control over θ, the error term in Weyl's law, and the error term in the probability estimate. Theorem 8 is remarkable because such Weyl laws are typically a feature of self-adjoint operator, whereas in the non-selfadjoint case they generally fail. Indeed, as laid out in Section 4.2, the discrete spectrum of the (unperturbed) non-selfadjoint operator Ph is usually localized to curves in the pseudospectrum Σ, see Melin and Sjöstrand [43], Hitrik and Sjöstrand [2831], and Rouby [49]. In contrast, Theorem 8 shows that a “generic” perturbation of size O(h) is sufficient for the spectrum to “fill out” Σ.

To illustrate this phenomenon, recall the non-selfadjoint harmonic oscillator Ph=-h2x2+ix2 on ℝ from Example 2. Its spectrum is given by {e/4(2n + 1)h; n ∈ ℕ} [14] on the line eiπ/4+. The Theorem 8 shows that a “generic” perturbation of arbitrarily small size is sufficient to produce spectrum roughly equidistributed in any fixed compact set in its classical spectrum Σ, which is in this case the upper right quadrant of ℂ.

As observed in Christiansen and Zworski [11], the real analytic p condition (Equation 34) consistently holds for some κ > 0. Similarly, when p is truly analytical and such that Σ ⊂ ℂ has non-empty interior, then

zΩ: dpp-1(z)0    (4.12) holds with  κ>1/2.    (36)

For smooth p, we have that when for every z ∈ ∂Ω

dp,dp¯ are linearly independent at every point of p-1(z),then  (4.12) holds with κ=1.    (37)

Observe that dp and dp¯ are linearly independent at ρ when {p,p¯}(ρ)0, where {a, b} = ∂ξa·∂xb − ∂xa·∂ξb denotes the Poisson bracket. Moreover, in dimension d = 1, the condition {p,p¯}0 on p−1(z) is equivalent to dp, with dp¯ being linearly independent at every point of p−1(z). However, in dimensions d > 1, this cannot in hold general, as the integral of {p,p¯} with respect to the Liouville measure on p−1(z) vanishes on every compact connected component of p−1(z), see reference [42, Lemma 8.1]. Furthermore, condition (Equation 37) cannot hold when z ∈ ∂Σ. However, some iterated Poisson brackets may not have zero there. For example, it has been observed in [25, Example 12.1] that if

ρp-1(Ω): {p,p¯}(ρ)0 or {p,{p,p¯}}(ρ)0,then  (4.12) holds with  κ=34.    (38)

4.3.3 Related results

Theorem 8 has also been extended to the case of elliptic semiclassical differential operators on compact manifolds by Sjöstrand [55], to the Toeplitz quantization of the torus by Christiansen and Zworski [11] and Vogel [66], and to general Berezin-Toeplitz quantizations on compact Kähler manifolds by Oltman [44] in the context of complex Gaussian noise. A further extension of Theorem 8 has been achieved by Becker, Oltman and the author in Becker et al. [6]. There we prove a probabilistic Weyl law for the non-selfadjoint off-diagonal operators of the Bistritzer-MacDonald Hamiltonian [7] for twisted bilayer graphene, see also Cancés et al. [9] and Watson et al. [67], subject to random tunneling potentials. This probabilistic Weyl has an interesting physical consequence as it demonstrates the instability of the so-called magic angels for this model of twisted bilayer graphene. Similar results have been achieved in random matrix theory. The case of Toeplitz matrices is represented by symbols on T2 of the form naneinξ, (x, ξ) ∈ T2, has been conducted in a series of recent studies by Śniady [61], Davies and Hager [15], Guionnet et al. [22], Basak et al. [4, 5], Sjöstrand and the author of this text [5759]. Such symbols amount to the case of symbols which are constant in the x variable. In these studies the non-selfadjointness of the problem, however, does not come from the symbol itself, but from the boundary conditions destroying it. The periodicity of the symbol in x is achieved by allowing for a discontinuity. Nevertheless, these studies demonstrate that by adding a small random matrix, the limit of the empirical eigenvalues counting measure μN of the perturbed operator converges in probability (or even almost surely in some cases) to p*().

Author contributions

MV: Writing – original draft, Writing – review & editing.

Funding

The author(s) declare financial support was received for the research, authorship, and/or publication of this article. I am partly supported by the ANR Grant ADYCT ANR-20-CE40-0017.

Conflict of interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Footnotes

1. ^This implies that the semiclassical wavefront set of e+ is defined by ρ0. In other words, the state e+ is concentrated in position and frequency near the point ρ0. See, for instance, Zworski [71] for a definition. For u = (u(h))h ∈ (0, 1) a bounded family in L2(ℝd), its semiclassical wavefront set WFh(u) denotes the phase space region where u is h-microlocalized:

WFh(u)=def{(x,ξ)T*d;  aCc(T*d),  a(x,ξ)=1,aw(x,hDx)u(h)L2=O(h),}

where aw denotes the Weyl quantization of a, and ∁U denotes the complement of a given set U.

2. ^∁Λ(p) denotes the complement of the set Λ(p).

3. ^° denotes the interior of the set Σ.

4. ^Hp0 denotes the Hamilton vector field induced by p0.

References

1. Aguilar J, Combes J-M. A class of analytic perturbations for one-body schrödinger Hamiltonians. Commun Math Phys. (1971) 22:269–79. doi: 10.1007/BF01877510

Crossref Full Text | Google Scholar

2. Anantharaman N. Spectral deviations for the damped wave equation. Geom Funct Anal. (2010) 20:3. doi: 10.1007/s00039-010-0071-x

Crossref Full Text | Google Scholar

3. Balslev E, Combes J-M. Spectral properties of many-body Schrödinger operators with dilation analytic interactions. Comm Math Phys. (1971) 22:280–94. doi: 10.1007/BF01877511

Crossref Full Text | Google Scholar

4. Basak A, Paquette E, Zeitouni O. Regularization of non-normal matrices by Gaussian noise - the banded Toeplitz and twisted Toeplitz cases. Forum of Mathematics, Sigma. (2019) 7:e3. doi: 10.1017/fms.2018.29

Crossref Full Text | Google Scholar

5. Basak A, Paquette E, Zeitouni O. Spectrum of random perturbations of Toeplitz matrices with finite symbols. Trans Amer Math Soc. (2020) 373:4999–5023. doi: 10.1090/tran/8040

Crossref Full Text | Google Scholar

6. Becker S, Oltman I, Vogel M. Absence of small magic angles for disordered tunneling potentials in twisted bilayer graphene. arXiv preprint arXiv:2402.12799. (2024).

Google Scholar

7. Bistritzer R, MacDonald A. Moiré bands in twisted double-layer graphene. PNAS. (2011) 108:12233–7. doi: 10.1073/pnas.1108174108

PubMed Abstract | Crossref Full Text | Google Scholar

8. Boulton L. Non-self-adjoint harmonic oscillator, compact semigroups and pseudospectra. J Operator Theory. (2002) 47:413–29.

Google Scholar

9. Cancès E, Garrigue L, Gontier D. A simple derivation of moiré-scale continuous models for twisted bilayer graphene. Phys Rev B. (2023) 107:155403. doi: 10.1103/PhysRevB.107.155403

Crossref Full Text | Google Scholar

10. Chazarain J. Spectre d'un hamiltonien quantique et mécanique classique. Comm PDE. (1980) 5:595–644. doi: 10.1080/0360530800882148

Crossref Full Text | Google Scholar

11. Christiansen TJ, Zworski M. Probabilistic Weyl laws for quantized tori. Comm Math Phys. (2010) 299:305–34. doi: 10.1007/s00220-010-1047-2

Crossref Full Text | Google Scholar

12. Cook N. Lower bounds for the smallest singular value of structured random matrices. Ann Probab. (2018) 46:3442–500. doi: 10.1214/17-AOP1251

Crossref Full Text | Google Scholar

13. Davies EB. Pseudo–spectra, the harmonic oscillator and complex resonances. Proc Royal Soc London A. (1999) 455:585–99. doi: 10.1098/rspa.1999.0325

Crossref Full Text | Google Scholar

14. Davies EB. Semi-classical states for non-self-adjoint Schrödinger operators. Comm Math Phys. (1999) 200:35–41. doi: 10.1007/s002200050521

Crossref Full Text | Google Scholar

15. Davies EB, Hager M. Perturbations of Jordan matrices. J Approx Theory. (2009) 156:82–94. doi: 10.1016/j.jat.2008.04.021

PubMed Abstract | Crossref Full Text | Google Scholar

16. Dencker N, Sjöstrand J, Zworski M. Pseudospectra of semiclassical (pseudo-) differential operators. Commun Pure Appl Mathem. (2004) 57:384–415. doi: 10.1002/cpa.20004

Crossref Full Text | Google Scholar

17. Dimassi M, Sjöstrand J. Spectral Asymptotics in the Semi-Classical Limit. London Mathematical Society Lecture Note Series 268, Cambridge: Cambridge University Press (1999).

Google Scholar

18. Dyatlov S, Zworski M. Mathematical Theory of Scattering Resonances. London: American Mathematical Society (2019).

Google Scholar

19. Embree M, Trefethen LN. Spectra and Pseudospectra: The Behavior of Nonnormal Matrices and Operators. Princeton: Princeton University Press (2005). doi: 10.1515/9780691213101

Crossref Full Text | Google Scholar

20. Galkowski J. Pseudospectra of semiclassical boundary value problems. J Inst Math Jussieu. (2014) 2:405–49. doi: 10.1017/S1474748014000061

Crossref Full Text | Google Scholar

21. Grigis A, Sjöstrand J. Microlocal Analysis for Differential Operators, London Mathematical Society Lecture Note Series 196. Cambridge: Cambridge University Press (1994).

Google Scholar

22. Guionnet A, Wood PM, Zeitouni O. Convergence of the spectral measure of non-normal matrices. Proc AMS. (2014) 142:667–79. doi: 10.1090/S0002-9939-2013-11761-2

PubMed Abstract | Crossref Full Text | Google Scholar

23. Hager M. Instabilité Spectrale Semiclassique d'Opérateurs Non-Autoadjoints II. Ann Henri Poincare. (2006) 7:1035–64. doi: 10.1007/s00023-006-0275-7

Crossref Full Text | Google Scholar

24. Hager M. Instabilité spectrale semiclassique pour des opérateurs non-autoadjoints I: un modèle. Ann faculté des Sci Toulouse Sé. (2006) 6:243–80. doi: 10.5802/afst.1121

PubMed Abstract | Crossref Full Text | Google Scholar

25. Hager M, Sjöstrand J. Eigenvalue asymptotics for randomly perturbed non-selfadjoint operators. Mathem Ann. (2008) 342:177–243. doi: 10.1007/s00208-008-0230-7

Crossref Full Text | Google Scholar

26. Helffer B, Robert D. Comportement semi-classique du spectre des hamiltoniens quantiques elliptiques. Ann Inst Fourier, Grenoble. (1981) 31:169–223. doi: 10.5802/aif.844

PubMed Abstract | Crossref Full Text | Google Scholar

27. Helffer B, Robert D. Calcul fonctionnel par la transformation de mellin et opérateurs admissibles. J Funct Anal. (1983) 53:246–68. doi: 10.1016/0022-1236(83)90034-4

Crossref Full Text | Google Scholar

28. Hitrik M, Sjöstrand J. Non-selfadjoint perturbations of selfadjoint operators in 2 dimensions I. Ann Henri Poincaré. (2004) 5:1–73. doi: 10.1007/s00023-004-0160-1

Crossref Full Text | Google Scholar

29. Hitrik M, Sjöstrand J. Non-selfadjoint perturbations of selfadjoint operators in 2 dimensions II. Vanishing averages. Comm Partial Differ Equat. (2005) 30:1065–1106. doi: 10.1081/PDE-200064447

Crossref Full Text | Google Scholar

30. Hitrik M, Sjöstrand J. Non-selfadjoint perturbations of selfadjoint operators in 2 dimensions III a. One branching point. Canadian J Math. (2008) 60:572–657. doi: 10.4153/CJM-2008-028-3

PubMed Abstract | Crossref Full Text | Google Scholar

31. Hitrik M, Sjöstrand J, Vũ Ngọc S. Diophantine tori and spectral asymptotics for non-selfadjoint operators. Amer J Math. (2007) 129:105–82. doi: 10.1353/ajm.2007.0001

PubMed Abstract | Crossref Full Text | Google Scholar

32. Hörmander L. Differential Equations without Solutions. Math Annalen. (1960) 140:169–173. doi: 10.1007/BF01361142

Crossref Full Text | Google Scholar

33. Hörmander L. Differential operators of principal type. Math Annalen. (1960) 140:124–146. doi: 10.1007/BF01360085

Crossref Full Text | Google Scholar

34. Hörmander L. The Analysis of Linear Partial Differential Operators IV, Grundlehren der mathematischen Wissenschaften, vol. 275. Cham: Springer-Verlag (1985).

Google Scholar

35. Ivrii V. Microlocal Analysis and Precise Spectral Asymptotics. New York: Springer (1998). doi: 10.1007/978-3-662-12496-3

Crossref Full Text | Google Scholar

36. Jin L. Damped wave equations on compact hyperbolic surfaces. Comm Math Phys. (2020) 373:771–94. doi: 10.1007/s00220-019-03650-x

Crossref Full Text | Google Scholar

37. R. Latala. Some estimates of norms of random matrices. Proc Amer Math Soc. (2005) 133:1273–82. doi: 10.1090/S0002-9939-04-07800-1

PubMed Abstract | Crossref Full Text | Google Scholar

38. Lebeau G. Équation des ondes amorties, Algebraic and Geometric Methods in Mathematical Physics (Kaciveli, 1993). Dordrecht: Kluwer Academy Publication (1996). doi: 10.1007/978-94-017-0693-3_4

Crossref Full Text | Google Scholar

39. Lindblad G. On the generators of quantum dynamical semigroups. Commun Math Phys. (1976) 48:119–30. doi: 10.1007/BF01608499

Crossref Full Text | Google Scholar

40. Markus AS, Matsaev VI. Comparison theorems for spectra of linear operators and spectral asymptotics Trudy Moskov. Mat Obshch. (1982) 45:133–81.

Google Scholar

41. Martinez A. An Introduction to Semiclassical and Microlocal Analysis. Cham: Springer (2002). doi: 10.1007/978-1-4757-4495-8

Crossref Full Text | Google Scholar

42. Melin A, Sjöstrand J. Determinants of pseudodifferential operators and complex deformations of phase space. Methods Appl Anal. (2002) 9:177–237. doi: 10.4310/MAA.2002.v9.n2.a1

Crossref Full Text | Google Scholar

43. Melin A, Sjöstrand J. Bohr-Sommerfeld quantization condition for non-selfadjoint operators in dimension 2. Astérisque. (2003) 284:181–244.

Google Scholar

44. Oltman I. A probabilistic Weyl-law for Berezin-Toeplitz operators. J Spectral Theory. (2023) 13:727–54. doi: 10.4171/jst/459

Crossref Full Text | Google Scholar

45. Petkov V, Robert D. Asymptotique semi-classique du spectre d'hamiltoniens quantiques et trajectoires classiques périodiques. Comm in PDE. (1985) 10:365–90. doi: 10.1080/03605308508820382

Crossref Full Text | Google Scholar

46. Pravda-Starov K. A general result about the pseudo-spectrum of Schrödinger operators. Proc R Soc London Ser A Math Phys Eng Sci. (2004) 460:471–7. doi: 10.1098/rspa.2003.1194

Crossref Full Text | Google Scholar

47. Pravda-Starov K. Étude du pseudo-spectre d'opérateurs non auto-adjoints. Ph.D. thesis (2006).

Google Scholar

48. Pravda-Starov K. Pseudo-spectrum for a class of semi-classical operators. Bull Soc Math France. (2008) 136:329–372. doi: 10.24033/bsmf.2559

Crossref Full Text | Google Scholar

49. Rouby O. Bohr-Sommerfeld quantization conditions for non-selfadjoint perturbations of selfadjoint operators in dimension one. Int Math Res Not IMRN. (2018) 7:2156–207. doi: 10.1093/imrn/rnw309

PubMed Abstract | Crossref Full Text | Google Scholar

50. Rudelson M, Vershynin R. The least singular value of a randomsquare matrix is O(n−1/2). C R Math Acad Sci Paris. (2008) 346:893–6. doi: 10.1016/j.crma.2008.07.009

Crossref Full Text | Google Scholar

51. Sankar A, Spielmann DA, Teng SH. Smoothed analysis of the condition numbers and growth factors of matrices. SIAM J, Matrix Anal Appl. (2006) 28:446–76. doi: 10.1137/S0895479803436202

Crossref Full Text | Google Scholar

52. Sjöstrand J. Asymptotic Distribution of Eigenfrequencies for Damped Wave Equations. Publ RIMS, Kyoto Univ. (2000) 36:573–611. doi: 10.2977/prims/1195142811

Crossref Full Text | Google Scholar

53. Sjöstrand J. Perturbations of selfadjoint operators with periodic classical flow. In: Wave Phenomena and Asymptotic Analysis. RIMS Kokyuroku (2003). p. 1315.

Google Scholar

54. Sjöstrand J. Eigenvalue distribution for non-self-adjoint operators with small multiplicative random perturbations. Ann Fac Sci Toulouse. (2009) 18:739–795. doi: 10.5802/afst.1223

PubMed Abstract | Crossref Full Text | Google Scholar

55. Sjöstrand J. Eigenvalue distribution for non-self-adjoint operators on compact manifolds with small multiplicative random perturbations. Ann Fac Toulouse. (2010) 19:277–301. doi: 10.5802/afst.1244

PubMed Abstract | Crossref Full Text | Google Scholar

56. Sjöstrand J. Resolvent estimates for non-selfadjoint operators via semigroups. In: Around the Research of Vladimir Maz'ya III, International Mathematical Series 13. New York: Springer (2010). p. 359–384. doi: 10.1007/978-1-4419-1345-6_13

Crossref Full Text | Google Scholar

57. Sjöstrand J, Vogel M. Large bi-diagonal matrices and random perturbations. J Spectral Theory. (2016) 6:977–1020. doi: 10.4171/jst/150

Crossref Full Text | Google Scholar

58. Sjöstrand J, Vogel M. General Toeplitz matrices subject to Gaussian perturbations. Ann Henri Poincaré. (2021) 22:49–81. doi: 10.1007/s00023-020-00970-w

Crossref Full Text | Google Scholar

59. Sjöstrand J, Vogel M. Toeplitz band matrices with small random perturbations. Indagationes Mathem. (2021) 32:275–322. doi: 10.1016/j.indag.2020.09.001

Crossref Full Text | Google Scholar

60. Sjöstrand J, Zworski M. Complex scaling and the distribution of scattering poles. J Am Mathem Soc. (1991) 43:4. doi: 10.2307/2939287

PubMed Abstract | Crossref Full Text | Google Scholar

61. Śniady P. Random regularization of brown spectral measure. J Functional Anal. (2002) 193:291–313. doi: 10.1006/jfan.2001.3935

Crossref Full Text | Google Scholar

62. Tao T, Vu V. Random matrices: the circular law. Commun Contemp Math. (2008) 10:261–307. doi: 10.1142/S0219199708002788

Crossref Full Text | Google Scholar

63. Tao T, Vu V. Smooth analysis of the condition number and the least singular value. Math Comp. (2010) 79:2333–52. doi: 10.1090/S0025-5718-2010-02396-8

PubMed Abstract | Crossref Full Text | Google Scholar

64. Trefethen LN. Pseudospectra of matrices. Numer Anal. (1992) 91:234–66.

Google Scholar

65. Trefethen LN. Pseudospectra of linear operators. SIAM Rev. (1997) 39:383–406. doi: 10.1137/S0036144595295284

Crossref Full Text | Google Scholar

66. Vogel M. Almost sure Weyl law for quantized tori. Comm Math Phys 378. (2020) 2:1539–85. doi: 10.1007/s00220-020-03797-y

Crossref Full Text | Google Scholar

67. Watson AB, Kong T, MacDonald AH, Luskin M. Bistritzer-macdonald dynamics in twisted bilayer graphene. J Math Phys. (2023) 64:031502. doi: 10.1063/5.0115771

Crossref Full Text | Google Scholar

68. Weyl H. Das asymptotische verteilungsgesetz der eigenwerte linearer partieller differentialgleichungen (mit einer anwendung auf die theorie der hohlraumstrahlung). Math Ann. (1912) 71:441–79. doi: 10.1007/BF01456804

Crossref Full Text | Google Scholar

69. Zworski M. A remark on a paper of E.B. Davies. Proc AMS. (2001) 129:2955–2957. doi: 10.1090/S0002-9939-01-05909-3

PubMed Abstract | Crossref Full Text | Google Scholar

70. Zworski M. Pseudospectra of semi-classical operators. Unpublished Leture, King's College (2001).

Google Scholar

71. Zworski M. Semiclassical Analysis, Graduate Studies in Mathematics 138. New York: American Mathematical Society (2012). doi: 10.1090/gsm/138

Crossref Full Text | Google Scholar

Keywords: semiclassical analysis, non-selfadjoint operators, random matrix, spectral theory, partial differential equation (PDE)

Citation: Vogel M (2024) Pseudospectra and eigenvalue asymptotics for disordered non-selfadjoint operators in the semiclassical limit. Front. Appl. Math. Stat. 10:1508973. doi: 10.3389/fams.2024.1508973

Received: 10 October 2024; Accepted: 11 November 2024;
Published: 02 December 2024.

Edited by:

Jose Luis Jaramillo, Université de Bourgogne, France

Reviewed by:

Wenfeng Chen, SUNY Polytechnic Institute, United States

Copyright © 2024 Vogel. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Martin Vogel, dm9nZWwmI3gwMDA0MDttYXRoLnVuaXN0cmEuZnI=

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.