1 Introduction
This paper resolves open computational spectral problems related to geometric features of spectra of operators. In other words, we consider the following problem:
Are there algorithms that given a bounded
1 operator
\(A\in {\mathcal {B}}(l^2({\mathbb {N}}))\), approximate key geometric features (e.g. spectral gaps, notions of sizes and capacity, measures, topological features such as fractal dimensions, etc.) of the set
\(\textrm{Sp}(A)\) from a matrix representation of
A?
To answer this question, we use the newly established solvability complexity index (SCI) hierarchy [
18,
51,
91], a classification tool that determines the boundaries of what is computationally possible. Classifying spectral problems and providing a library of optimal algorithms
2 remains largely uncharted territory in the foundations of computational mathematics. In exploring this territory, there will, necessarily, have to be many different types of algorithms, as different structures on the various classes of operators and different spectral properties require different techniques.
A famous example of the above question is the almost Mathieu operator on
\(l^2({\mathbb {Z}})\) (see Sect.
4.4):
$$\begin{aligned} (H_{\alpha }x)_n=x_{n-1}+x_{n+1}+2\lambda \cos (2\pi n\alpha )x_n, \end{aligned}$$
which induces the Hofstadter butterfly [
92]. The almost Mathieu operator plays an important role in physics [
104], arising in the study of the quantum Hall effect [
160], and has become a laboratory for exploring the spectral properties of ergodic Schrödinger operators [
95]. When
\(\alpha \) is irrational, the Lebesgue measure of the spectrum is
\(4\left| 1-\left| \lambda \right| \right| \). This formula was conjectured based on the numerical work of Aubry and André [
8] and became one of B. Simon’s problems for the twenty-first century [
146]. It was later proven by Avila and Krikorian [
11]. Similarly, M. Kac’s “Ten Martini Problem”, that the spectrum is a Cantor set for all irrational
\(\alpha \) and
\(\lambda >0\), was conjectured by Azbel [
13] and also became one of B. Simon’s problems. This problem attracted a host of numerical and analytical work (see the summary in [
104]), before being proven by Avila and Jitomirskaya [
9]. In both of these examples, we see a crucial interplay between computation, conjecture, and mathematical proof. The above geometric features of spectra play an important role in the physics of the underlying quantum system [
90,
99,
100,
147]. The almost Mathieu operator is by no means unique in this regard, and there is a growing literature on computational studies of geometric features of spectra in diverse areas of physics [
14,
68,
83,
94,
103,
106,
110,
120,
125,
133,
138,
139,
156,
161].
However, there is a current lack of rigorous computational theory and convergence analysis, and no known algorithms can tackle general cases. Moreover, the foundations of computation (i.e. what is and what is not computationally possible) for computing geometric features of spectra are almost entirely unexplored. We solve these open problems and others by providing algorithms that compute geometric features of spectra and by classifying the computational problems in the SCI hierarchy.
1.1 The SCI Hierarchy
The SCI hierarchy has recently been used to resolve the problem of computing spectra of general bounded operators in infinite dimensions [
18,
91] and is now being used to explore the foundations of computation in many diverse areas of mathematics [
2,
15,
16,
19‐
23,
30,
52,
53,
55,
57,
59,
60,
64,
140,
141,
166].
3 Whilst for some classes of operators one can compute spectra with error control [
54,
60,
64], a potentially surprising consequence is that, for general operators, one needs several successive limits to compute the spectrum. Since traditional approaches are dominated by techniques based on one limit, this explains why many computational spectral problems remain unsolved and opens the door to an infinite classification theory. Moreover, this phenomenon is not just restricted to spectral problems but is shared by other areas of computational mathematics. An example is S. Smale’s problem of root-finding of polynomials with rational maps [
149], which also requires several successive limits as established by McMullen [
115,
116] and Doyle and McMullen [
70]. These results can be expressed in terms of the SCI hierarchy [
18], which generalises Smale’s seminal work [
148,
150] with Blum et al. [
28,
29,
66], and his program on the foundations of scientific computing and the existence of algorithms. Many other problems in the foundations of computations, such as the work by Weinberger [
167], can also be viewed in the context of the SCI hierarchy.
The SCI hierarchy is further motivated by computer-assisted proofs. Computer-assisted proofs are rapidly becoming an essential part of modern mathematics [
86] and, perhaps surprisingly, non-computable problems can be used in computer-assisted proofs. Examples include the recent proof of Kepler’s conjecture (Hilbert’s 18th problem) [
87,
88] on optimal packings of 3-spheres, led by T. Hales, and the Dirac–Schwinger conjecture on the asymptotic behaviour of ground states of certain Schrödinger operators, proven in a series of papers by Fefferman and Seco [
72‐
80]. Both of these proofs rely on computing non-computable problems. This apparent paradox can be explained by the SCI hierarchy (the
\(\Sigma ^A_1\) and
\(\Pi _1^A\) classes described below become available for computer-assisted proofs); Hales, Fefferman and Seco implicitly prove
\(\Sigma ^A_1\) classifications in the SCI hierarchy in their papers. Some of the problems we consider also lie in
\(\Sigma ^A_1\cup \Pi _1^A\), meaning that they can be used for computer-assisted proofs.
1.2 The Problems Addressed in this Paper
The algorithms we provide are sharp in the SCI hierarchy, meaning that they realise the boundaries of what computers can achieve. Table
1 provides a summary of the main SCI classifications of this paper. The main theorems are contained in Sect.
3, including further motivations and classifications for different classes of operators. We provide resolutions to the following problems:
(i)
Computing spectral radii, essential spectral radii, polynomial operator norms and capacity of spectra. The spectral radius is perhaps the most basic geometric property of spectra and arises in stability analysis. We show that computing the spectral radius is high up in the SCI hierarchy for non-normal operators. In fact, it has the same classification in the SCI hierarchy for general bounded operators as that of computing the spectrum itself. Classifications are given for different types of operators (e.g. known column decay, control on resolvent norms) and also for the essential spectral radius. In many cases, the problem of computing polynomial operator norms is easier in the sense of SCI hierarchy. We also consider the problem of computing the logarithmic capacity of the spectrum, following the work of Halmos [
89], which has applications in orthogonal polynomials, approximation theory and when studying the convergence of Krylov methods (see, for example, the work of Nevanlinna [
121‐
123] and Miekkala and Nevanlinna [
117]).
(ii)
Computing essential numerical ranges, gaps in essential spectra, and determining whether spectral pollution occurs on sets. We provide classification results for the essential numerical range, which also hold in the case of unbounded operators. In connection with computing spectra, there has been a substantial effort in studying the finite section method and locating gaps in essential spectra of operators (see the discussion in Sect.
3.4). When using the finite section method to approximate spectra of self-adjoint operators, spurious eigenvalues, known as spectral pollution, can occur anywhere within these gaps. Paradoxically, we show that determining if spectral pollution occurs on a given set is strictly harder in the sense of the SCI hierarchy than computing the spectrum itself. Hence, computing a failure flag for the finite section method is, in a certain sense, strictly harder than solving the original problem for which it was designed. Moreover, we establish the SCI of detecting gaps in essential spectra of self-adjoint operators, a problem that arises in areas such as perturbation theory and defect models.
(iii)
Computing Lebesgue measure of spectra and pseudospectra, and determining if the spectrum is Lebesgue null. An important property of the spectrum is its Lebesgue measure, with recent progress in the field of Schrödinger operators with random or almost periodic potentials [
9,
11,
12,
17,
135]. If the spectrum of an operator is Lebesgue null; then, this implies the absence of absolutely continuous spectra,
4 which is related to transport properties if the operator represents a Hamiltonian. Whilst results are known for specific one-dimensional examples such as the almost Mathieu operator [
11] or the Fibonacci Hamiltonian [
154], very little is known in the general case or higher dimensions. This is reflected by the difficulty of performing rigorous numerical studies, despite many examples studied in the physics literature (see the references in [
10,
24,
147]). We provide the first algorithms for computing the Lebesgue measure of spectra and pseudospectra, and determining whether the spectrum is Lebesgue null, for many different classes of operators.
(iv)
Computing fractal dimensions of spectra. Fractal dimensions of spectra are important in many applications. For example, in quantum mechanics, they lead to upper bounds on the spreading of wavepackets and are related to time-dependent quantities associated with wave functions [
90,
99,
100]. Fractal spectra appear in a wide variety of contexts, such as exciting new results in multilayer materials (e.g. bilayer graphene) [
68,
83,
94,
133], strained materials [
120,
139] or quasicrystals [
14,
103,
106,
156]. Another well-studied area where fractal spectral properties appear is optics [
125,
138], following the analytical and numerical work of Berry and coauthors [
25‐
27]. Despite the physical importance of fractal dimensions, analytical results are known only for a limited number of specific models. Moreover, there are currently no algorithms for computing fractal dimensions of spectra for general operators, or even tridiagonal self-adjoint operators. We provide the first algorithms for computing the box-counting and Hausdorff dimensions of spectra for many different classes of operators.
1.3 Contributions to the SCI Hierarchy Itself
Our final contribution is a new tool to prove lower bounds (impossibility results) in the SCI hierarchy. This is crucial for some of the classifications of the above problems and holds regardless of the model of computation. We show that for a certain special class of combinatorial problems, the SCI hierarchy is equivalent to the Baire hierarchy from descriptive set theory. (This equivalence does not hold in general.) By embedding these combinatorial problems into spectral problems,
5 we provide the first technique for dealing with problems that have SCI greater than three and also greatly simplify the proofs of results lower down in the SCI hierarchy. However, it should be stressed that this is not a paper on descriptive set theory or mathematical logic. Our discussion is entirely self-contained and written for a wide audience from a primarily computational background.
1.4 Outline of Paper
In Sect.
2, we provide a brief summary of the SCI hierarchy and define the classes of operators for the interpretation of Table
1 and the main results. A detailed discussion of the SCI hierarchy is delayed until Sect.
5.1. In Sect.
3, we summarise our main results on the classification of computational spectral problems. Computational examples are then given in Sect.
4. For example, we provide numerical evidence that a portion of the spectrum of the graphical Laplacian on an infinite Penrose tile is Lebesgue null and fractal, with a fractal dimension of approximately 0.8, and that the whole spectrum has a logarithmic capacity of approximately 2.26. Mathematical preliminaries, including definitions of the SCI hierarchy and the new tool to provide lower bounds in the SCI hierarchy, are presented in Sect.
5. Proofs of our results are given in Sects.
6–
9. To make the paper self-contained, we include a short appendix on the results/algorithms of [
64], which are used in some of our proofs. Pseudocode for many of the new algorithms is provided in “Appendix B”.
Table 1
Summary of the main results for the readable information \(\Lambda _1\) consisting of matrix values
Computing the spectral radius. | Varies. e.g. Normal operators: \(\in \Sigma _1^A\), \(\not \in \Delta _1^G\) | |
Controlled resolvent: \(\in \Sigma _2^A\), \(\not \in \Delta _2^G\) |
General bounded operators: \(\in \Pi _3^A\), \(\not \in \Delta _3^G\) |
Computing the essential spectral radius. | Varies. e.g. Most classes: \(\in \Pi _2^A\), \(\not \in \Delta _2^G\) | |
General bounded operators: \(\in \Pi _3^A\), \(\not \in \Delta _3^G\) |
Computing polynomial operator norms. | With bounded dispersion: \(\in \Sigma _1^A\), \(\not \in \Delta _1^G\) | |
Without bounded dispersion: \(\in \Sigma _2^A\), \(\not \in \Delta _2^G\) |
Computing the capacity of the spectrum. | With bounded dispersion: \(\in \Pi _2^A\), \(\not \in \Delta _2^G\) | |
Without bounded dispersion: \(\in \Pi _3^A\), \(\not \in \Delta _3^G\) |
Computing gaps in the essential spectrum. | \(\in \Sigma _3^A\), \(\not \in \Delta _3^G\) | |
Computing the essential numerical range. | \(\in \Pi _2^A\), \(\not \in \Delta _2^G\) | |
Determining whether spectral pollution can occur on a set (i.e. failure of finite section method). | \(\in \Sigma _3^A\), \(\not \in \Delta _3^G\) | |
Computing the Lebesgue measure of the spectrum. | Varies. e.g. | |
Bounded dispersion: \(\in \Pi _2^A\), \(\notin \Delta _2^G\) |
Self-adjoint and general bounded: \(\in \Pi _3^A\), \(\notin \Delta _3^G\) |
Computing the Lebesgue measure of the pseudospectrum. | Varies. e.g. | |
Bounded dispersion: \(\in \Sigma _1^A\), \(\notin \Delta _1^G\) |
Self-adjoint and general bounded: \(\in \Sigma _2^A\), \(\notin \Delta _2^G\) |
Determining whether the Lebesgue measure of the spectrum is zero. | Varies. e.g. | |
Bounded dispersion: \(\in \Pi _3^A\), \(\notin \Delta _3^G\) |
Self-adjoint and general bounded: \(\in \Pi _4^A\), \(\notin \Delta _4^G\) |
Computing the box-counting dimension of the spectrum (when it exists). | Varies. e.g. | |
Bounded dispersion:\(\in \Pi _2^A\), \(\notin \Delta _2^G\) |
Self-adjoint: \(\in \Pi _3^A\), \(\notin \Delta _3^G\) |
Computing the Hausdorff dimension of the spectrum. | Varies. e.g. | |
Bounded dispersion:\(\in \Sigma _3^A\), \(\notin \Delta _3^G\) |
Self-adjoint: \(\in \Sigma _4^A\), \(\notin \Delta _4^G\) |
5 Mathematical Preliminaries and Combinatorial Problems in the SCI Hierarchy
In this section, we begin by providing formal definitions of the SCI hierarchy. We then link the SCI hierarchy, in a certain specific case, to the Baire hierarchy on a suitable topological space. As well as being interesting in its own right, this provides a useful method of providing canonical problems high up in the SCI hierarchy. In particular, the results we prove hold for towers of
general algorithms (see Definition
5.1) without the restrictions of arithmetic operations or notions of recursivity etc. This will be used extensively in the proofs of lower bounds for spectral problems that have
\(\textrm{SCI}>2\), where we typically reduce the problems discussed here to the given spectral problem. It should be stressed that such links to existing hierarchies only exist in special cases when
\(\Omega \) and
\({\mathcal {M}}\) are particularly well-behaved. Even when such a link does exist, the induced topology on
\(\Omega \) is often too complicated, unnatural or strong to be useful from a computational viewpoint. We also take the view that, for problems of scientific interest, the mappings
\(\Lambda \) and metric space
\({\mathcal {M}}\) are often given to us apriori from the corresponding applications and are typically not compatible with topological viewpoints of computation.
5.1 The SCI Hierarchy
We begin by defining the solvability complexity index (SCI) hierarchy, allowing us to show that our algorithms realise the boundary of what computers can achieve. We have already presented the definition of a computational problem
\(\{\Xi ,\Omega ,{\mathcal {M}},\Lambda \}\) in §
2.1. Recall that the goal is to find algorithms that approximate the function
\(\Xi \). More generally, the main pillar of our framework is the concept of a tower of algorithms, which is needed to describe problems that need several successive limits in the computation. However, first one needs the definition of a general algorithm.
The definition of a general algorithm is more general than the definition of a Turing machine [
164] or a BSS machine [
28]. A general algorithm has no restrictions on the operations allowed. The only restriction is that it can only take a finite amount of information, though it is allowed to
adaptively choose the finite amount of information it reads depending on the input. Condition (iii) ensures that the algorithm consistently reads the information. With a definition of a general algorithm, we can define the concept of towers of algorithms.
In addition to a general tower of algorithms, we focus on arithmetic towers.
Given the definitions above we can now define the key concept, namely the solvability complexity index:
We will let
\(\textrm{SCI}(\Xi ,\Omega )_{\textrm{A}}\) and
\(\textrm{SCI}(\Xi ,\Omega )_{\textrm{G}}\) denote the SCI with respect to an arithmetic tower and a general tower, respectively. Note that a general tower means just a tower of algorithms as in Definition
5.2, where there are no restrictions on the mathematical operations. Thus, clearly
\(\textrm{SCI}(\Xi ,\Omega )_{\textrm{A}} \ge \textrm{SCI}(\Xi ,\Omega )_{\textrm{G}}\). The definition of the SCI immediately induces the SCI hierarchy:
When there is additional structure on the metric space, such as in the spectral case when one considers the Attouch–Wets or the Hausdorff metric, one can extend the SCI hierarchy. For non-empty closed sets, we consider the
Attouch–Wets metric defined by
$$\begin{aligned} d_{\textrm{AW}}(C_1,C_2)=\sum _{n=1}^{\infty } 2^{-n}\min \left\{ {1,\underset{\left| x\right| \le n}{\sup }\left| \textrm{dist}(x,C_1)-\textrm{dist}(x,C_2)\right| }\right\} , \end{aligned}$$
(5.1)
for
\(C_1,C_2\in \textrm{Cl}({\mathbb {C}}),\) where
\(\textrm{Cl}({\mathbb {C}})\) denotes the set of closed non-empty subsets of
\({\mathbb {C}}\). This generalises the familiar Hausdorff metric to unbounded closed sets and corresponds to local uniform converge on compact subsets of
\({\mathbb {C}}\).
Note that to build a \(\Sigma _1\) algorithm, it is enough (by taking subsequences of n) to construct \(\Gamma _n(A)\) such that \(\Gamma _{n}(A) \subset {\mathcal {N}}_{E_n(A)}(\Xi (A))\) with some computable \(E_n(A)\) that converges to zero. The same idea can be applied to the real line with the usual metric, or \(\{0,1\}\) with the discrete metric (we interpret 1 as “Yes”).
Suppose we are given a computational problem
\(\{\Xi , \Omega , {\mathcal {M}}, \Lambda \}\), and that
\(\Lambda = \{f_j\}_{j \in \beta }\), where
\(\beta \) is some index set that can be finite or infinite. Obtaining
\(f_j\) may be a computational task on its own, which is exactly the problem in most areas of computational mathematics. In particular, for
\(A \in \Omega \),
\(f_j(A)\) could be the number
\(e^{\frac{\pi }{j} i }\) for example. Hence, we cannot access
\(f_j(A)\), but rather
\(f_{j,n}(A)\) where
\(f_{j,n}(A) \rightarrow f_{j}(A)\) as
\(n \rightarrow \infty \). Or, just as for problems that are high up in the SCI hierarchy, it could be that we need several successive limits, in particular one may need mappings
\(f_{j,n_m,\ldots , n_1}: \Omega \rightarrow {\mathbb {D}} + i{\mathbb {D}}\), where
\({\mathbb {D}}\) denotes the dyadic rational numbers, such that
$$\begin{aligned} \lim _{n_m \rightarrow \infty } \ldots \lim _{n_1 \rightarrow \infty } \Vert f_{j,n_m,\ldots , n_1}(A) - f_j(A)\Vert _{\infty } = 0 \quad \forall j\in \beta ,\forall A \in \Omega . \end{aligned}$$
(5.2)
In particular, we may view the problem of obtaining
\(f_j(A)\) as a problem in the SCI hierarchy, where
\(\Delta _1\) classification would correspond to the existence of mappings
\(f_{j,n}: \Omega \rightarrow {\mathbb {D}} + i {\mathbb {D}}\) such that
$$\begin{aligned} \Vert f_{j,n}(A) - f_j(A)\Vert _{\infty } \le 2^{-n} \quad \forall j\in \beta ,\forall A \in \Omega . \end{aligned}$$
(5.3)
This idea is formalised in the following definition.
We want algorithms that can handle all computational problems \(\{\Xi ,\Omega ,{\mathcal {M}},{{\hat{\Lambda }}}\}\) when \({{\hat{\Lambda }}} \in {\mathcal {L}}^m(\Lambda )\). To formalise this, we define a computational problem with \(\Delta _m\)-information.
The SCI and the SCI hierarchy, given \(\Delta _m\)-information, are then defined in the standard obvious way.
5.2 Recalling Some Results from Descriptive Set Theory
We briefly recall the definition of the Borel hierarchy as well as some well-known theorems from descriptive set theory. It is beyond the scope of this paper to provide an extensive discussion of descriptive set theory, but we refer the reader to [
98, Chapter 2] for an excellent introduction that covers the main ideas.
Let
X be a metric space and define
$$\begin{aligned} \Sigma _1^0(X)=\{U\subset X:U\text { is open}\},\quad \Pi _1^0(X)=\sim \Sigma _1^0(X)=\{F\subset X:F\text { is closed}\}, \end{aligned}$$
where for a class
\({\mathcal {U}}\),
\(\sim {\mathcal {U}}\) denotes the class of complements (in
X) of elements of
\({\mathcal {U}}\). Inductively define
$$\begin{aligned} \Sigma ^0_\xi (X)&=\{\cup _{n\in {\mathbb {N}}}A_n:A_n\in \Pi ^0_{\xi _n},\xi _n<\xi \},\text { if }\xi >1,\\ \Pi _\xi ^0(X)&=\sim \Sigma _\xi ^0(X),\quad \Delta _\xi ^0(X)=\Sigma ^0_\xi (X)\cap \Pi _\xi ^0(X). \end{aligned}$$
The full Borel hierarchy extends to all
\(\xi <\omega _1\) (
\(\omega _1\) being the first uncountable ordinal) by transfinite induction but we do not need this here.
Given metric spaces
X and
Y, the Baire hierarchy is defined as follows. A function
\(f:X\rightarrow Y\) is of Baire class 1, written
\(f\in {\mathcal {B}}_1\), if it is
\(\Sigma _2^0(X)\)-measurable. For
\(1<\xi <\omega _1\), a function
\(f:X\rightarrow Y\) is of Baire class
\(\xi \), written
\(f\in {\mathcal {B}}_\xi \), if it is the pointwise limit of a sequence of functions
\(f_n\) in
\({\mathcal {B}}_{\xi _n}\) with
\(\xi _n<\xi \). The following Theorem is well-known [
98, Section 24] and provides a useful link between the Borel and Baire hierarchies.
The assumption that X is zero-dimensional in the last statement is important. Without any assumptions, the final statement of the theorem is false, as is easily seen by considering \(X={\mathbb {R}}\). Examples of zero-dimensional spaces include products of the discrete space \(\{0,1\}\) or the Cantor space. Any such space is necessarily totally disconnected, meaning that the connected components in the space are the one-point sets (the converse is true for locally compact Hausdorff spaces). Our primary interest will be when Y is equal to \(\{0,1\}\) or [0, 1], both with their natural topologies.
5.3 Linking the SCI Hierarchy to the Baire Hierarchy in a Special Case
The generated topology can be very perverse and not every class of algorithms is closed under search. However, we do have the following useful theorem when \(\Omega \) (and \(\Lambda \)) is a particularly simple discrete space.
5.4 Combinatorial Problems High up in the SCI Hierarchy
We can now combine the results of the previous two subsections and obtain combinatorial problems high up in the SCI hierarchy. Let
\(k\in {\mathbb {N}}_{\ge 2}\) and let
\(\Omega _k\) denote the collection of all infinite arrays
\(\{a_{m_1,\ldots ,m_k}\}_{m_1,\ldots ,m_k\in {\mathbb {N}}}\) with entries
\(a_{m_1,\ldots ,m_k}\in \{0,1\}\). As usual,
\(\Lambda _k\) is the set of component-wise evaluations/projections. Consider the formulas
$$\begin{aligned} P(a,m_1,\ldots ,m_{k-2})&={\left\{ \begin{array}{ll} 1,\quad \text {if }\, \exists i \, \forall j\, \exists n>j \text { s.t. }a_{m_1,\ldots ,m_{k-2},n,i}=1\\ 0,\quad \text {otherwise} \end{array}\right. },\\ Q(a,m_1,\ldots ,m_{k-2})&={\left\{ \begin{array}{ll} 1,\quad \text {if } \, \forall ^{\infty }i\, \forall \, j\, \exists n>j \text { s.t. }a_{m_1,\ldots ,m_{k-2},n,i}=1\\ 0,\quad \text {otherwise} \end{array}\right. }, \end{aligned}$$
where
\(\forall ^{\infty }\) means “for all but a finite number of”. In words,
P decides whether the corresponding matrix has a column with infinitely many 1’s, whereas
Q decides whether the matrix has only finitely many columns with only finitely many 1’s. For
\(R=P\) or
Q, consider the problem function for
\(a\in \Omega _k\)$$\begin{aligned} \Xi _{k,R}(a)={\left\{ \begin{array}{ll} \exists m_1\text { } \forall m_2\text { }\ldots \text { }\forall m_{k-2} R(a,m_1,\ldots ,m_{k-2}),\quad \text {if }k\text { is even}\\ \forall m_1\text { } \exists m_2\text { }\ldots \text { }\forall m_{k-2} R(a,m_1,\ldots ,m_{k-2}),\quad \text {otherwise} \end{array}\right. }, \end{aligned}$$
that is, so that all quantifier types alternate.
For our applications to spectral problems, we will use
\({\tilde{\Omega }}\) to denote
\(\Omega _k\) and consider
$$\begin{aligned} \begin{aligned} {\tilde{\Xi }}_1=\Xi _{2,P},\quad {\tilde{\Xi }}_2=\Xi _{2,Q},\\ {\tilde{\Xi }}_3=\Xi _{3,P},\quad {\tilde{\Xi }}_4=\Xi _{3,Q}. \end{aligned} \end{aligned}$$
(5.6)
Theorem
5.19 holds for a much wider class of decision problems, but these four are the only ones we shall use in the sequel. The decision problems
\({{\tilde{\Xi }}}_1\) and
\({{\tilde{\Xi }}}_2\) were shown to have
\(\textrm{SCI}_G=3\) in [
18], but only with regard to the discrete space
\({\mathcal {M}}=\{0,1\}\) and the proof used a somewhat complicated Baire category argument. Theorem
5.19 is much more general, can be extended to arbitrarily large SCI, and has a much slicker proof, making clear a beautiful connection with the Baire hierarchy for well-behaved
\(\Omega \).
6 Proofs Concerning Spectral Radii, Essential Spectral Radii, Capacity and Operator Norms
Here we prove the theorems found in Sects.
3.1–
3.3. First, we briefly recall
\(\Sigma _1^A\) algorithms for spectral problems presented in [
64], that are sharp in the SCI hierarchy. The algorithms constructed in [
64] are shown as pseudocode in “Appendix A”, where we also refer the reader to a more detailed account. The following was proven in [
64] and was generalised in [
60] to unbounded operators:
We now turn to the proof of Theorem
3.3, dealing first with the evaluation set
\(\Lambda _1\). Suppose that
\(\{{\tilde{\Gamma }}_{n_k,\ldots ,n_1}\}\) is a
\(\Pi ^A_k\) tower of algorithms to compute the spectrum of a class of operators, where the output is a finite set for each
\(n_1,\ldots ,n_k\). It is then clear that
$$\begin{aligned} \Gamma _{n_k,\ldots ,n_1}(A)=\sup _{z\in {\tilde{\Gamma }}_{n_k,\ldots ,n_1}(A)}\left| z\right| +\frac{1}{2^{n_k}} \end{aligned}$$
provides a
\(\Pi ^A_k\) tower of algorithms for the spectral radius. Strictly speaking, the above may not be an arithmetic tower owing to the absolute value. But it can be approximated to arbitrary precision (from above say), the error of which can be absorbed in the first limit. In what follows, we always assume this is done without further comment. Similarly if
\(\{{\tilde{\Gamma }}_{n_k,\ldots ,n_1}\}\) provides a
\(\Sigma ^A_k\) tower of algorithms for the spectrum (and outputs a finite set for each
\(n_1,\ldots ,n_k\)),
$$\begin{aligned} \Gamma _{n_k,\ldots ,n_1}(A)=\sup _{z\in {\tilde{\Gamma }}_{n_k,\ldots ,n_1}(A)}\left| z\right| -\frac{1}{2^{n_k}} \end{aligned}$$
provides a
\(\Sigma ^A_k\) tower of algorithms for the spectral radius. If we only have a height
k tower with no
\(\Sigma _k\) or
\(\Pi _k\) type error control for the spectrum, then taking the supremum of absolute values shows that we get a height
k tower for the spectral radius.
The fact that
\(\{\Xi _r,\Omega _\textrm{D}\} \in \Sigma ^A_1\),
\(\{\Xi _r,\Omega _f\cap \Omega _g\} \in \Sigma ^A_1\),
\(\{\Xi _r,\Omega _g\} \in \Sigma ^A_2\),
\(\{\Xi _r,\Omega _f\} \in \Pi ^A_2\) and
\(\{\Xi _r,\Omega _{\textrm{B}}\} \in \Pi ^A_3\) hence follow from Theorems
6.1 and the results of [
18]. It is clear that
\(\{\Xi _r,\Omega _\textrm{D}\}\notin \Delta ^G_1\) and this also shows that
\(\{\Xi _r,\Omega _\textrm{N}\}\notin \Delta ^G_1\) and
\(\{\Xi _r,\Omega _f\cap \Omega _g\}\notin \Delta ^G_1\). Hence, we must show the positive result that
\(\{\Xi _r,\Omega _\textrm{N}\} \in \Sigma ^A_1\) and prove the lower bounds
\(\{\Xi _r,\Omega _g\}\notin \Delta ^G_2\),
\(\{\Xi _r,\Omega _f\}\notin \Delta ^G_2\) and
\(\{\Xi _r,\Omega _{\textrm{B}}\}\notin \Delta ^G_3\).
Proof of Theorem 3.3for \(\Lambda _1\) Throughout this proof, we use the evaluation set
\(\Lambda _1\), which we drop from the notation for convenience.
Step 1 \(\{\Xi _r,\Omega _\textrm{N}\} \in \Sigma ^A_1\). Recall that the spectral radius of a normal operator
\(A\in \Omega _{\textrm{B}}\) is equal to its operator norm. Consider the finite section matrices
\(P_nAP_n\in {\mathbb {C}}^{n\times n}\). It is straightforward to show that
$$\begin{aligned} \left\| P_nAP_n\right\| \uparrow \left\| A\right\| \quad \text { as }n\rightarrow \infty . \end{aligned}$$
The norm
\(\left\| P_nAP_n\right\| \) is the square root of the largest eigenvalue of the semi-positive definite self-adjoint matrix
\((P_nAP_n)^*(P_nAP_n)\). This can be estimated from below to an accuracy of 1/
n using Corollary 6.9 of [
60], which then yields a
\(\Sigma _1^A\) algorithm for
\(\{\Xi _r,\Omega _\textrm{N}\}\).
Step 2 \(\{\Xi _r,\Omega _g\}\notin \Delta ^G_2\). Recall that we assumed the existence of a
\(\delta \in (0,1)\) such that
\(g(x)\le (1-\delta )x\). Let
\(\epsilon >0\), then it is easy to see that the matrices
$$\begin{aligned} S_{\pm }(\epsilon )=\begin{pmatrix} 1 &{} 0 \\ \pm \epsilon &{} 1 \end{pmatrix} \end{aligned}$$
have norm bounded by
\(1+\epsilon +\epsilon ^2\) and are clearly inverse of each other. Choose
\(\epsilon \) small such that
\((1+\epsilon +\epsilon ^2)^2\le 1/(1-\delta )\). If
\(B\in {\mathbb {C}}^{2\times 2}\) is normal, it follows that
\({\hat{B}}:=S_{+}(\epsilon )BS_{-}(\epsilon )\) lies in
\(\Omega _g\) and has the same spectrum as
B. We choose
$$\begin{aligned} {\hat{B}}=S_{+}(\epsilon )\begin{pmatrix} 1 &{} -\epsilon \\ -\epsilon &{} 0 \end{pmatrix}S_{-}(\epsilon )=\begin{pmatrix} 1+\epsilon ^2 &{} -\epsilon \\ \epsilon ^3 &{} -\epsilon ^2 \end{pmatrix}. \end{aligned}$$
The crucial property of
\({\hat{B}}\) is that the first entry
\(1+\epsilon ^2\) is strictly greater in magnitude than the two eigenvalues
\((1\pm \sqrt{1+4\epsilon ^2})/2\).
Now suppose for a contradiction that a height one tower,
\(\{\Gamma _n\}\), solves the problem. We will gain a contradiction by showing that
\(\Gamma _n(A)\) does not converge for an operator of the form,
$$\begin{aligned} A=\bigoplus _{r=1}^\infty A_{l_r},\quad A_{m}:=\begin{pmatrix} 1+\epsilon ^2&{} &{} &{} &{}-\epsilon \\ &{}0&{} &{} &{} \\ &{} &{}\ddots &{} &{} \\ &{} &{} &{}0&{} \\ \epsilon ^3&{} &{} &{} &{}-\epsilon ^2\\ \end{pmatrix} \in {\mathbb {C}}^{m\times m}, \end{aligned}$$
where we only consider
\(l_k\ge 3\). Each
\(A_m\) is unitarily equivalent to the matrix
\({\hat{B}}\oplus 0\in {\mathbb {C}}^{m\times m}\) and has spectrum equal to
\(\{0,(1\pm \sqrt{1+4\epsilon ^2})/2\}\). Any
A of the above form is unitarily equivalent to a direct sum of an infinite number of
\({\hat{B}}\)’s and the zero operator and hence lies in
\(\Omega _g\). Now suppose that
\(l_1,\ldots ,l_k\) have been chosen and consider the operator
$$\begin{aligned} B_k=A_{l_1}\oplus \dots \oplus A_{l_k}\oplus C,\quad C=\textrm{diag}\{1+\epsilon ^2,0,\ldots \}. \end{aligned}$$
The spectrum of
\(B_k\) is
\(\{0,(1\pm \sqrt{1+4\epsilon ^2})/2,1+\epsilon ^2\}\), and hence, there exists
\(\eta >0\) and
\(n(k)\ge k\) such that
\(\Gamma _{n(k)}(B_k)>(1+\sqrt{1+4\epsilon ^2})/2+\eta \). But
\(\Gamma _{n(k)}(B_k)\) can only depend on the evaluations of the matrix entries
\(\{B_k\}_{ij}=\langle B_ke_j,e_i \rangle \) with
\(i,j\le N(B_k,n(k))\) (as well as evaluations of the function
g) into account. If we choose
\(l_{k+1}>N(B_k,n(k))\) then by the assumptions in Definition
5.1,
\(\Gamma _{n(k)}(A)=\Gamma _{n(k)}(B_k)>(1+\sqrt{1+4\epsilon ^2})/2+\eta \). But
\(\Gamma _n(A)\) must converge to
\((1+\sqrt{1+4\epsilon ^2})/2\), a contradiction.
Step 3 \(\{\Xi _r,\Omega _f\}\notin \Delta ^G_2\). Suppose for a contradiction that a height one tower,
\(\{\Gamma _n\}\), solves the problem. We will gain a contradiction by showing that
\(\Gamma _n(A)\) does not converge for an operator of the form
$$\begin{aligned} A=\bigoplus _{r=1}^\infty C_{l_r}\oplus A_{l_r},\quad{} & {} A_{m}:=\begin{pmatrix} 0&{} 1&{} &{} &{}\\ &{}0&{} 1&{} &{} \\ &{} &{}\ddots &{} \ddots &{} \\ &{} &{} &{}&{} 1\\ &{} &{} &{} &{}0\\ \end{pmatrix} \in {\mathbb {C}}^{m\times m},\quad \\ {}{} & {} C_m=\textrm{diag}\{0,0,\ldots ,0\}\in {\mathbb {C}}^{m\times m}, \end{aligned}$$
where we assume that
\(l_r\ge r\) to ensure that the spectrum of
A is equal to the unit disc
\(B_1(0)\). Note that the function
\(f(n)=n+1\) will do for the bounded dispersion with
\(c_n=0\). Now suppose that
\(l_1,\ldots ,l_k\) have been chosen and consider the operator
$$\begin{aligned} B_k=\big (C_{l_1}\oplus A_{l_1}\big )\oplus \cdots \oplus \big (C_{l_k}\oplus A_{l_k}\big )\oplus C,\quad C=\textrm{diag}\{0,0,\ldots \}. \end{aligned}$$
The spectrum of
\(B_k\) is
\(\{0\}\) and hence there exist
\(n(k)\ge k\) such that
\(\Gamma _{n(k)}(B_k)<1/4\). But
\(\Gamma _{n(k)}(B_k)\) can only depend on the evaluations of the matrix entries
\(\{B_k\}_{ij}=\langle B_ke_j,e_i \rangle \) with
\(i,j\le N(B_k,n(k))\) (as well as evaluations of the function
f) into account. If we choose
\(l_{k+1}>N(B_k,n(k))\), then by the assumptions in Definition
5.1,
\(\Gamma _{n(k)}(A)=\Gamma _{n(k)}(B_k)<1/4\). But
\(\Gamma _n(A)\) must converge to 1, a contradiction.
Step 4 \(\{\Xi _r,\Omega _{\textrm{B}}\}\notin \Delta ^G_3\). Suppose for a contradiction that
\(\{\Gamma _{n_2,n_1}\}\) is a height two (general) tower and without loss of generality, assume it to be nonnegative. We use the results of Sect.
5. Let
\(({\mathcal {M}},d)\) be the space [0, 1] with the usual metric (note in particular this is not discrete so we use Remark
5.20), let
\({{\tilde{\Omega }}}\) denote the collection of all infinite matrices
\(\{a_{i,j}\}_{i,j\in {\mathbb {N}}}\) with entries
\(a_{i,j}\in \{0,1\}\) and recall the problem function
$$\begin{aligned} {{\tilde{\Xi }}}_1(\{a_{i,j}\}):\text { Does }\{a_{i,j}\}\text { have a column containing infinitely many nonzero entries?} \end{aligned}$$
Theorem
5.19 in Sect.
5 shows that
\(\textrm{SCI}({{\tilde{\Xi }}}_1,{{\tilde{\Omega }}})_{G} = 3\). We will gain a contradiction by using the supposed height two tower to solve
\(\{{{\tilde{\Xi }}}_1,{{\tilde{\Omega }}}\}\).
Without loss of generality, identify
\(\Omega _{\textrm{B}}\) with
\({\mathcal {B}}(X)\) where
\(X=\bigoplus _{j=1}^{\infty }X_j\) in the
\(l^2\)-sense with
\(X_j=l^2({\mathbb {N}})\). Now let
\(\{a_{i,j}\}\in {{\tilde{\Omega }}}\) and define
\(B_j\in {\mathcal {B}}(X_j)\) with the matrix representation
$$\begin{aligned} (B_{j})_{k,i}= {\left\{ \begin{array}{ll} 1, &{} \text {if } k=i\text { and }a_{k,j}=0\\ 1, &{} \text {if } k<i\text { and }a_{l,j}=0\text { for }k<l<i\\ 0, &{} \text {otherwise } 0\le n\le 1. \end{array}\right. } \end{aligned}$$
Let
\({\mathcal {I}}_j\) be the index set of all
i where
\(a_{i,j}=1\).
\(B_j\) acts as a unilateral shift on
\(\overline{\textrm{span}}\{e_k:k\in {\mathcal {I}}_j\}\) and the identity on its orthogonal complement. It follows that
$$\begin{aligned} \textrm{Sp}(B_j)= {\left\{ \begin{array}{ll} 1, &{} \text {if } {\mathcal {I}}_j=\emptyset \\ \{0,1\}, &{} \text {if } {\mathcal {I}}_j\text { is finite and non-empty}\\ {\mathbb {D}}\quad (\text {the unit disc}),&{} \text {if } {\mathcal {I}}_j\text { is infinite}. \end{array}\right. } \end{aligned}$$
For the matrix
\(\{a_{i,j}\}\) define
\(A\in \Omega _{\textrm{B}}\) by
$$\begin{aligned} A=\bigoplus _{j=1}^{\infty }(B_j-\frac{1}{2}I_j), \end{aligned}$$
where
\(I_j\) denotes the identity operator on
\({\mathbb {C}}^{j\times j}\), then
\(\mathrm {Sp(A)}=\overline{\cup _{j=1}^\infty \textrm{Sp}(B_j)}-\frac{1}{2}\).
Hence, we see that
$$\begin{aligned} \Xi _r(A)= {\left\{ \begin{array}{ll} \frac{1}{2}, &{} \text {if }{{\tilde{\Xi }}}_1(\{a_{i,j}\})=0\\ \frac{3}{2}, &{} \text {if }{{\tilde{\Xi }}}_1(\{a_{i,j}\})=1. \end{array}\right. } \end{aligned}$$
We then set
\({\tilde{\Gamma }}_{n_2,n_1}(\{a_{i,j}\})=\min \{\max \{\Gamma _{n_2,n_1}(A)-1/2,0\},1\}\). It is clear that this defines a generalised algorithm mapping into [0, 1]. In particular, given
N we can evaluate
\(\{A_{k,l}:k,l\le N\}\) using only finitely many evaluations of
\(\{a_{i,j}\}\), where we can use a bijection between canonical bases of
\(l^2({\mathbb {N}})\) and
\(\bigoplus _{j=1}^{\infty }X_j\) to view
A as acting on
\(l^2({\mathbb {N}})\). But then
\(\{{{\tilde{\Gamma }}}_{n_2,n_1}\}\) provides a height two tower for
\(\{{{\tilde{\Xi }}}_1,{{\tilde{\Omega }}}\}\), a contradiction.
\(\square \)
Proof of Theorem 3.3for \(\Lambda _2\) Here we prove the changes for
\(\Xi _r\) when we consider the evaluation set
\(\Lambda _2\). It is clear that the classifications in
\(\Sigma _1^A\) do not change. It is also easy to use the algorithm in Theorem
6.1 (now using
\(\Lambda _2\) to collapse the first limit and approximate
\(\gamma _n\)—see “Appendix A”) to prove
\(\{\Xi _r,\Omega _g,\Lambda _2\}\in \Sigma _1^A\). Similarly we can use the algorithm for the spectrum of operators in
\(\Omega _f\) for
\(\Omega _{\textrm{B}}\) using
\(\Lambda _2\) to collapse the first limit and hence
\(\{\Xi _r,\Omega _{\textrm{B}},\Lambda _2\}\in \Pi _2^A\). Since
\(\Omega _f\subset \Omega _{\textrm{B}}\), it follows that we only need to prove
\(\{\Xi _r,\Omega _f,\Lambda _2\}\not \in \Delta _2^G\). This can be proven using exactly the same example and a similar argument to step 3 of the proof of Theorem
3.3 (hence omitted).
\(\square \)
Proof of Theorem 3.6
We begin by proving the results for \(\Lambda _1\). For the lower bounds, it is enough to show that \(\{\Xi _{er},\Omega _\textrm{D},\Lambda _1\}\not \in \Delta _2^G\) and \(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\not \in \Delta _3^G\). For the upper bounds, we must show that \(\{\Xi _{er},\Omega _f,\Lambda _1\}\in \Pi _2^A\), \(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\in \Pi _3^A\) and \(\{\Xi _{er},\Omega _\textrm{N},\Lambda _1\}\in \Pi _2^A\). The lower bounds for \(\Lambda _2\) follow from \(\{\Xi _{er},\Omega _\textrm{D},\Lambda _1\}\not \in \Delta _2^G\) and for the upper bounds it is enough to prove \(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _2\}\in \Pi _2^A\).
Step 1 \(\{\Xi _{er},\Omega _\textrm{D},\Lambda _1\}\not \in \Delta _2^G\). This is the same argument as in step 3 of the proof of Theorem
3.3; however, now we replace
\(A_m\) by
\(A_m=\textrm{diag}\{1,1,\ldots ,1\}\in {\mathbb {C}}^{m\times m}\) and use the fact that
\(\Xi _{er}(B_k)=0\). It follows that given the proposed height one tower
\(\{\Gamma _n\}\) and the constructed
A,
\(\Xi _{er}(A)=1\) but
\(\Gamma _{n(k)}(A)<1/4\), the required contradiction.
Step 2 \(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\not \in \Delta _3^G\). This is the same argument as step 4 of the proof of Theorem
3.3.
Step 3 \(\{\Xi _{er},\Omega _f,\Lambda _1\}\in \Pi _2^A\),
\(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\in \Pi _3^A\) and
\(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _2\}\in \Pi _2^A\).
\(\{\Xi _{er},\Omega _f,\Lambda _1\}\in \Pi _2^A\) follows immediately from the existence of a
\(\Pi _2^A\) tower of algorithms for the essential spectrum of operators in
\(\Omega _f\) proven in [
18]. The output of this tower is a finite collection of rectangles with complex rational vertices; hence, we can gain an approximation of the maximum absolute value over this output to any given precision. This can be used to construct a
\(\Pi _2^A\) tower for
\(\{\Xi _{er},\Omega _f,\Lambda _1\}\). Similarly,
\(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\in \Pi _3^A\) follows from the
\(\Pi _3^A\) tower of algorithms for
\(\{\textrm{Sp}_{\textrm{ess}},\Omega _{\textrm{B}},\Lambda _1\}\) constructed in [
18]. Finally, we can use
\(\Lambda _2\) to collapse the first limit of the algorithm for the essential spectrum in [
18], giving a
\(\Pi _2^A\) algorithm, and this can be used to show
\(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _2\}\in \Pi _2^A\).
Step 4 \(\{\Xi _{er},\Omega _\textrm{N},\Lambda _1\}\in \Pi _2^A\). A
\(\Pi _2^A\) tower is constructed in the proof of Theorem
3.10 for the essential numerical range,
\(W_e(A)\), of normal operators (using
\(\Lambda _1\)), and this outputs a finite collection of points. For normal operators
A,
\(W_e(A)\) is the convex hull of the essential spectrum and hence
\(\sup _{z\in W_e(A)}\left| z\right| \) is equal to
\(\Xi _{er}(A)\). Hence, a
\(\Pi _2^A\) tower for
\(\{\Xi _{er},\Omega _\textrm{N},\Lambda _1\}\) follows by taking the maximum absolute value over the tower for
\(W_e(A)\).
\(\square \)
Now choose a sequence of rational numbers
\(\{z_j\}_{j\in {\mathbb {N}}}\in [-1,1]\) that is also dense in
\([-1,1]\) and let
\(B_j=B(z_1,\ldots ,z_j)\). For each column of a given
\(\{a_{i,j}\}\in {{\tilde{\Omega }}}\), let the infinite matrix
\(C^{(j)}\) be defined as follows. If
\(k,l<j+1\) then
\(C^{(j)}_{kl}=z_{k}\delta _{k,l}\). Let
r(
i) denote the row of the
ith one of the column
\(\{a_{i,j}\}_{i\in {\mathbb {N}}}\) (with
\(r(i)=\infty \) if
\(\sum _{m}a_{m,j}<i\) and
\(r(0)=0\)). If
\(r(i)<\infty \) then for
\(k\le l\) define
$$\begin{aligned} C^{(j)}_{kl}={\left\{ \begin{array}{ll} a_p\delta _{k,l-(r(i)-r(i-1)-1)}, &{} p=1,\ldots ,j,l=r(i)+j\cdot (2i-1)+p-1\\ -z_p\delta _{k,l}, &{} p=1,\ldots ,j,l=r(i)+j\cdot (2i-1)+p-1\\ z_p\delta _{k,l}, &{} p=1,\ldots ,j,l=r(i)+2j\cdot i+p-1\\ 0, &{} \text {otherwise}, \end{array}\right. } \end{aligned}$$
and extend
\(C^{(j)}_{kl}\) below the diagonal to a symmetric matrix. The key property of this matrix is that if the column
\(\{a_{i,j}\}_{i\in {\mathbb {N}}}\) has infinitely many 1 s, then its is unitarily equivalent to an infinite direct sum of infinitely many
\(B_j\) together with the zero operator acting on some subspace (whose dimension is equal to the number of zeros in the column). In this case
\(\textrm{Sp}(C^{(j)})=\{-1,1,0\}\) or
\(\{-1,1\}\). On the other hand, if
\(\{a_{i,j}\}_{i\in {\mathbb {N}}}\) has finitely many 1 s, then
\(C^{(j)}\) is unitarily equivalent the direct sum of a finite number of
\(B_j\), the diagonal operator
\(\textrm{diag}\{z_1,\ldots ,z_k\}\) and the zero operator acting on some subspace. In this case
\(\{z_1,\ldots ,z_j\}\subset \textrm{Sp}(C^{(j)})\). Let
\(A=\bigoplus _{j=1}^{\infty }C^{(j)}\), then it is clear that if
\({{\tilde{\Xi }}}_2(\{a_{i,j}\})=1\), then
\(\textrm{Sp}(A)\) is a finite set, otherwise it is the entire interval
\([-1,1]\).
Now we use the following facts for bounded self-adjoint operators A. If \(\textrm{Sp}(A)\) is a finite set then \(\Xi _{cap}(A)=0\), whereas if \(\textrm{Sp}(A)=[-1,1]\) then \(\Xi _{cap}(A)=1/2\) (this can be proven easily using the minimal \(l^\infty \) norm property of monic Chebyshev polynomials). We then define \({\tilde{\Gamma }}_{n_2,n_1}(\{a_{i,j}\})=\min \{\max \{1-2\Gamma _{n_2,n_1}(A),0\},1\}\). It is clear that this defines a generalised algorithm. In particular, given N we can evaluate \(\{A_{k,l}:k,l\le N\}\) using only finitely many evaluations of \(\{a_{i,j}\}\), where we can use a bijection between canonical bases of \(l^2({\mathbb {N}})\) and \(\bigoplus _{j=1}^{\infty }X_j\) to view A as acting on \(l^2({\mathbb {N}})\). We also have the convergence \(\lim _{n_2\rightarrow \infty }\lim _{n_1\rightarrow \infty }{\tilde{\Gamma }}_{n_2,n_1}(\{a_{i,j}\})={{\tilde{\Xi }}}_2(\{a_{i,j}\})\), a contradiction.
Step 4 \(\{\Xi _{cap},\Omega _{\textrm{D}},\Lambda _2\}\not \in \Delta ^G_2\). This is the same argument as in step 3 of the proof of Theorem
3.3. However, we now replace
\(A_m\) by
\(A_m=\textrm{diag}\{d_1,d_2,\ldots ,d_m\}\in {\mathbb {C}}^{m\times m}\), where
\(\{d_m\}\) is a dense subsequence of
\([-1,1]\), and use the fact that
\(\Xi _{cap}(B_k)=0\). It follows that given the proposed height one tower
\(\{\Gamma _n\}\) and the constructed
A,
\(\Xi _{cap}(A)=1/2\) but
\(\Gamma _{n(k)}(A)<1/4\), the required contradiction.
Step 5 \(\{\Xi _{r,p},\Omega _{\textrm{SA}},\Lambda _2\}\not \in \Delta ^G_2\). Recall that we are given some polynomial
p of degree at least two. We assume without loss of generality that the zeros of
p are
\(\pm 1\) and
\(\left| p(0)\right| >1\) (the more general case is similar). The argument is similar to step 3 of the proof of Theorem
3.3, but we spell it out since it uses Lemma
6.3. Suppose for a contradiction that a height one tower,
\(\{\Gamma _n\}\), solves the problem. We will gain a contradiction by showing that
\(\Gamma _n(A)\) does not converge for an operator of the form,
$$\begin{aligned} A=\bigoplus _{r=1}^\infty B(z_1,\ldots ,z_{l_r}), \end{aligned}$$
and define
$$\begin{aligned} C=\textrm{diag}\{z_1,z_2,\ldots \}\in \Omega _{\textrm{B}}. \end{aligned}$$
We assume that
\(l_r\ge r\) to ensure that the spectrum of
A is equal to
\(\{-1,1\}\) and hence
\(\Xi _{r,p}(A)=0\). Now suppose that
\(l_1,\ldots ,l_k\) have been chosen and consider the operator
$$\begin{aligned} B_k=B(z_1)\oplus \cdots \oplus B(z_1,\ldots ,z_{l_k})\oplus C. \end{aligned}$$
The spectrum of
\(B_k\) is
\([-1,1]\) so that
\(\Xi _{r,p}(B_k)>1\) and hence there exists
\(n(k)\ge k\) such that
\(\Gamma _{n(k)}(B_k)>1/4\). But
\(\Gamma _{n(k)}(B_k)\) can only depend on the evaluations of the matrix entries
\(\{B_k\}_{ij}=\langle B_ke_j,e_i \rangle \) with
\(i,j\le N(B_k,n(k))\) (as well as evaluations of the function
f) into account. If we choose
\(l_{k+1}>N(B_k,n(k))\) then by the assumptions in Definition
5.1,
\(\Gamma _{n(k)}(A)=\Gamma _{n(k)}(B_k)>1/4\). But
\(\Gamma _n(A)\) must converge to 0, a contradiction.
\(\square \)
7 Proofs Concerning Essential Numerical Ranges, Essential Spectra and Spectral Pollution
Proof of Theorem 3.10for \(\Xi _{we}\) For the lower bounds, it is enough to note that
\(\{\Xi _{we},\Omega _\textrm{D},\Lambda _2\}\not \in \Delta _2^G\) by the same argument as step 1 of the proof of Theorem
3.6. The construction is exactly the same but yields
\(d_{\textrm{H}}(\Gamma _{n(k)}(A),\{0\})\le 1/2\), whereas
\(\Xi _{we}(A)=[0,1]\). Hence, the proposed height one tower cannot converge. To construct a
\(\Pi _2^A\) tower for general operators, we need the following Lemma:
It is well known that for
\(A\in \Omega _{\textrm{B}}\),
$$\begin{aligned} \overline{W(P_nA|_{P_n{\mathcal {H}}})}&\uparrow \overline{W(A)},\\ \overline{W((I-P_n)A|_{(I-P_n){\mathcal {H}}})}&\downarrow W_{e}(A). \end{aligned}$$
Given
A, let
\(\Gamma _{n_2,n_1}(A)\) be a finite collection of points produced by the algorithm in Lemma
7.1 applied to
\(B=(I-P_{n_2})P_{n_1+n_2+1}A|_{P_{n_1+n_2+1}(I-P_{n_2}){\mathcal {H}}}\) and
\(\epsilon =1/n_1\). The above limits show that
\(\{\Gamma _{n_2,n_1}\}\) provides a
\(\Pi _2^A\) tower for
\(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\).
\(\square \)
Proof of Theorem 3.10for \(\Xi _{poll}^{{\mathbb {F}}}\) We will prove that
\(\{\Xi _{poll}^{{\mathbb {R}}},\Omega _\textrm{D},\Lambda _i\}\not \in \Delta ^G_3\) and
\(\{\Xi _{poll}^{{\mathbb {C}}},\Omega _{\textrm{B}},\Lambda _1\}\in \Sigma ^A_3\). The construction of towers for
\(\Xi _{poll}^{{\mathbb {R}}}\) is similar, as are the arguments for lower bounds.
Step 1 \(\{\Xi _{poll}^{{\mathbb {C}}},\Omega _{\textrm{B}},\Lambda _1\}\in \Sigma ^A_3\). Let
\(\{{{\tilde{\Gamma }}}_{n_2,n_1}\}\) be the
\(\Pi _2^A\) tower for
\(\{\Xi _{er},\Omega _{\textrm{B}},\Lambda _1\}\) constructed above. Recall the definition
$$\begin{aligned} \gamma _{n_2,n_1}(z;A)=\min \{\sigma _{\textrm{inf}}(P_{n_1}(A-zI){|_{P_{n_2}{\mathcal {H}}}}),\sigma _{\textrm{inf}}(P_{n_1}(A^*-{\bar{z}}I){|_{P_{n_2}{\mathcal {H}}}})\} \end{aligned}$$
and that this can be approximated to any given accuracy in finitely many arithmetic operations and comparisons (see also “Appendix A”). We assume that we approximate from below to an accuracy of
\(1/n_1\) and call this approximation
\(\tilde{\gamma }_{n_2,n_1}\). The function
\(\gamma _{n_2,n_1}(z;A)\) is Lipschitz continuous with Lipschitz constant bounded by 1. Define the set
$$\begin{aligned} V_{n_1}=\bigcup _{m=1}^{n_1}U_{m}, \end{aligned}$$
where
\(U_{m}\) are the approximations to the open set
U. By taking squares of distances to ball centres, we can decide whether a point
\(z\in {\mathbb {Q}}+i{\mathbb {Q}}\) has
\(\textrm{dist}(z,V_{n_1})<\eta \) for any given
\(\eta \in {\mathbb {Q}}_+\). Let
\(\Upsilon _{n_2,n_1}(A,U)\) be the finite collection of all
\(z\in {{\tilde{\Gamma }}}_{n_2,n_1}(A)\) with
\(\textrm{dist}(z,V_{n_1})<1/n_2-1/n_1\). If
\(\Upsilon _{n_2,n_1}(A,U)\) is empty then set
\(Q_{n_2,n_1}(A,U)=0\), otherwise set
$$\begin{aligned} Q_{n_2,n_1}(A,U):=\sup _{z\in \Upsilon _{n_2,n_1}(A,U)} {{\tilde{\gamma }}}_{n_2,n_1}(z;A)-\frac{1}{n_1}. \end{aligned}$$
The above remarks show that this can be computed using finitely many arithmetic operations and comparisons.
Let
\(W_{n_2}=\overline{W((I-P_{n_2})A|_{(I-P_{n_2}){\mathcal {H}}})}\) and
\(W_{n_2,n_1}=W((I-P_{n_2})P_{n_1+n_2+1}A|_{P_{n_1+n_2+1}(I-P_{n_2}){\mathcal {H}}})\). We claim that the set
\(\Upsilon _{n_2,n_1}(A,U)\) converges to
$$\begin{aligned} \Upsilon _{n_2}(A,U):=\overline{\left\{ z\in W_{n_2}:\textrm{dist}(z,{\overline{U}})<\frac{1}{n_2}\right\} }, \end{aligned}$$
as
\(n_1\rightarrow \infty \), meaning also if
\(\Upsilon _{n_2}(A,U)\) is empty then
\(\Upsilon _{n_2,n_1}(A,U)\) is empty for large
\(n_1\). If
\(z\in \Upsilon _{n_2,n_1}(A,U)\), then there exists
\({\hat{z}}\in W_{n_2,n_1}\subset W_{n_2}\) with
\(\left| z-{\hat{z}}\right| \le 1/{n_1}\). Since
$$\begin{aligned} \textrm{dist}(z,{\overline{U}})\le \textrm{dist}(z,V_{n_1})<1/n_2-1/n_1, \end{aligned}$$
it follows that
\(\textrm{dist}({\hat{z}},{\overline{U}})<{1}/{n_2}\) and hence
\(\Upsilon _{n_2}(A,U)\) is non-empty. So to prove convergence, we only need to deal with the case
\(\Upsilon _{n_2}(A,U)\ne \emptyset \). The above argument also shows that any limit point of a subsequence
\(z_{m(j)}\in \Upsilon _{n_2,m(j)}(A,U)\) must lie in
\(\Upsilon _{n_2}(A,U)\). Hence to prove the claim, we need to only prove that for any
\(z\in \Upsilon _{n_2}(A,U)\), there exists
\(z_{n_1}\) that are contained in
\(\Upsilon _{n_2,n_1}(A,U)\) for large
\(n_1\) and converge to
z.
Let
\(z\in W_{n_2} \) with
\(\textrm{dist}(z,{\overline{U}})<{1}/{n_2}\), then there exists
\(\epsilon >0\) and
\(j>0\) such that
\(\textrm{dist}(z,U_j)<{1}/{n_2}-\epsilon \). There also exists
\(z_{n_1}\in {{\tilde{\Gamma }}}_{n_2,n_1}(A)\) with
\(z_{n_1}\rightarrow z\). It must hold for
\(n_1>j\) that
$$\begin{aligned} \textrm{dist}({z}_{n_1},V_{n_1})\le \textrm{dist}({z}_{n_1},V_j)&\le \left| z_{n_1}-z\right| +\textrm{dist}(z,U_j)\\&<\left| z_{n_1}-z\right| +\frac{1}{n_2}-\epsilon . \end{aligned}$$
This last quantity is smaller than
\(1/n_2-1/n_1\) for large
\(n_1\) and hence
\(z_{n_1}\in \Upsilon _{n_2,n_1}(A,U)\) for large
\(n_1\). It follows for any
\(z\in \Upsilon _{n_2}(A,U)\), there exists
\(z_{n_1}\) that are contained in
\(\Upsilon _{n_2,n_1}(A,U)\) for large
\(n_1\) and converge to
z.
Define
$$\begin{aligned} Q_{n_2}(A,U):=\sup _{z\in \Upsilon _{n_2}(A,U)} \gamma _{n_2}(z;A), \end{aligned}$$
where we recall that
\(\gamma _{n_2}(z;A)=\min \{\sigma _{\textrm{inf}}((A-zI){|_{P_{n_2}{\mathcal {H}}}}),\sigma _{\textrm{inf}}((A^*-{\bar{z}}I){|_{P_{n_2}{\mathcal {H}}}})\}.\) If
\(z\in \Upsilon _{n_2,n_1}(A,U)\), then the above shows that there exists
\({\hat{z}}\in \Upsilon _{n_2}(A,U)\) with
\(\left| z-{\hat{z}}\right| \le 1/{n_1}\). It follows that
$$\begin{aligned} {{\tilde{\gamma }}}_{n_2,n_1}(z;A)-\frac{1}{n_1}&\le \gamma _{n_2,n_1}(z;A)-\frac{1}{n_1}\\&\le \gamma _{n_2,n_1}({\hat{z}};A)\le \gamma _{n_2}(z;A), \end{aligned}$$
where we have used the bound on the Lipschitz constant and the fact that
\(\gamma _{n_2,n_1}\) converge up to
\(\gamma _{n_2}\) (and uniformly on compact subsets of
\({\mathbb {C}}\)). It follows that
\(Q_{n_2,n_1}(A,U)\le Q_{n_2}(A,U)\) and this also covers the case that
\(\Upsilon _{n_2}(A,U)=\emptyset \) if we define the supremum over the empty set to be 0. The set convergence proven above and uniform convergence of
\({{\tilde{\gamma }}}_{n_2,n_1}\) implies that
\(Q_{n_2,n_1}(A,U)\) converges to
\(Q_{n_2}(A,U)\). It is also clear that the
\(\Upsilon _{n_2}(A,U)\) are nested and converge down to
\(W_e(A)\cap {\overline{U}}\) since
\(W_{n_2}\) converges down to
\(W_{e}(A)\). The functions
\(\gamma _{n_2}\) also converge down to
$$\begin{aligned} \gamma (z;A)=\left\| R(z,A)\right\| ^{-1} \end{aligned}$$
uniformly on compact subsets of
\({\mathbb {C}}\) and hence
\(Q_{n_2}(A,U)\) converges down to
$$\begin{aligned} Q(A,U)=\sup _{z\in W_e(A)\cap {\overline{U}}} \left\| R(z,A)\right\| ^{-1}. \end{aligned}$$
Define
$$\begin{aligned} \Gamma _{n_3,n_2,n_1}(A,U)=1-\chi _{[0,1/{n_3}]}(Q_{n_2,n_1}(A,U))\in \{0,1\}. \end{aligned}$$
The above show that
$$\begin{aligned} \lim _{n_1\rightarrow \infty } \Gamma _{n_3,n_2,n_1}(A,U)=1-\chi _{[0,1/{n_3}]}(Q_{n_2}(A,U))=:\Gamma _{n_3,n_2}(A,U). \end{aligned}$$
Since
\(\chi _{[0,1/n_3]}\) has right limits and
\(Q_{n_2}(A,U)\) are non-increasing,
$$\begin{aligned} \lim _{n_2\rightarrow \infty } \Gamma _{n_3,n_2}(A,U)=1-\chi _{[0,1/{n_3}]}(Q(A,U)\pm )=:\Gamma _{n_3}(A,U), \end{aligned}$$
where ± denotes one of the right or left limits (it is possible to have either). Now if
\(\Xi _{poll}^{{\mathbb {C}}}(A,U)=0\), then
\(\Gamma _{n_3}(A,U)=0\) for all
\(n_3\). But if
\(\Xi _{poll}^{{\mathbb {C}}}(A,U)=1\), then for large
\(n_3\),
\(\Gamma _{n_3}(A,U)=1\). Moreover, in this latter case,
\(\Gamma _{n_3}(A,U)=1\) signifies the existence of
\(z\in W_e(A)\cap {\overline{U}}\) with
\(\gamma (z;A)>0\) and hence
\(z\not \in \textrm{Sp}(A)\). Hence,
\(\{\Gamma _{n_3,n_2,n_1}\}\) provides a
\(\Sigma _3^A\) tower.
Step 2 \(\{\Xi _{poll}^{{\mathbb {R}}},\Omega _{\textrm{D}},\Lambda _2\}\not \in \Delta ^G_3\). We will argue for the case that
\(U=U_1={\mathbb {R}}\) and the restricted case is similar. Assume for a contradiction that this is false and that
\(\{{\widehat{\Gamma }}_{n_2,n_1}\}\) is a general height two tower for
\(\{\Xi _{poll}^{{\mathbb {R}}},\Omega _{\textrm{D}},\Lambda _2\}\). We follow the same strategy as the proof of Theorem
3.3 step 4 (recall also the results of Sect.
5). Let
\(({\mathcal {M}},d)\) be discrete space
\(\{0,1\}\) and
\({{\tilde{\Omega }}}\) denote the collection of all infinite matrices
\(\{a_{i,j}\}_{i,j\in {\mathbb {N}}}\) with entries
\(a_{i,j}\in \{0,1\}\) and consider the problem function
$$\begin{aligned} {{\tilde{\Xi }}}_1(\{a_{i,j}\}):\text { Does }\{a_{i,j}\}\text { have a column containing infinitely many nonzero entries?} \end{aligned}$$
For
\(j\in {\mathbb {N}}\), let
\(\{b_{i,j}\}_{i\in {\mathbb {N}}}\) be a dense subset of
\(I_j:=[1-1/2^{2j-1},1-1/{2^{2j}}]\). Given a matrix
\(\{a_{i,j}\}_{i,j\in {\mathbb {N}}}\in {{\tilde{\Omega }}}\), construct a matrix
\(\{c_{i,j}\}_{i,j\in {\mathbb {N}}}\) by letting
\(c_{i,j}=a_{i,j}b_{r(i,j),j}\) where
$$\begin{aligned} r(i,j)=\max \left\{ 1,\sum _{k=1}^{i}a_{k,j}\right\} . \end{aligned}$$
Now consider any bijection
\(\phi :{\mathbb {N}}\rightarrow {\mathbb {N}}^2\) and define the diagonal operator
$$\begin{aligned} A=\textrm{diag}(c_{\phi (1)},c_{\phi (2)},c_{\phi (3)},\ldots ). \end{aligned}$$
The algorithm
\({\widehat{\Gamma }}_{n_2,n_1}\) thus translates to an algorithm
\(\Gamma '_{n_2,n_1}\) for
\(\{{{\tilde{\Xi }}}_1,{{\tilde{\Omega }}}\}\). Namely, set
\(\Gamma '_{n_2,n_1}(\{a_{i,j}\}_{i\in {\mathbb {N}}})={\widehat{\Gamma }}_{n_2,n_1}(A)\). The fact that
\(\phi \) is a bijection shows that the lowest level
\(\Gamma '_{n_2,n_1}\) are generalised algorithms (and are consistent). In particular, given
N, we can find
\(\{A_{i,j}:i,j\le N\}\) using finitely many evaluations of the matrix values
\(\{c_{k,l}\}\) (the same is true for
\(A^*A\) and
\(AA^*\) since the operator is diagonal). But for any given
\(c_{k,l}\) we can evaluate this entry using only finitely many evaluations of the matrix values
\(\{a_{m,n}\}\) by the construction of
r. Finally, note that
$$\begin{aligned} \textrm{Sp}(A)=\{1\}\cup \left( \bigcup _{j:\{a_{i,j}\}_{i\in {\mathbb {N}}}\text { has infinitely many 1s}}I_j\right) \cup Q, \end{aligned}$$
where
Q lies in the discrete spectrum. The intervals
\(I_j\) are also separated. It follows that there is a gap in the essential spectrum if and only if there exists a column
\(\{a_{i,j}\}_{i\in {\mathbb {N}}}\) with infinitely many 1 s. Otherwise the essential spectrum is
\(\{1\}\). It follows that
\(\tilde{\Xi }(\{a_{i,j}\})=\Xi _{poll}^{{\mathbb {R}}}(A,{\mathbb {R}})\), and hence, we get a contradiction.
\(\square \)
7.1 Essential Numerical Range for Unbounded Operators
The essential numerical range (see (
3.1)) was first introduced for a bounded operator
A in [
152], as the closure of the numerical range of the image of
A in the Calkin algebra:
$$\begin{aligned} W_e(A)=\bigcap _{B\in \Omega _K}\overline{W(A+B)}. \end{aligned}$$
Other equivalent characterisations were then given in [
82]. The unbounded case is significantly different from the bounded case, and definitions that are equivalent in the bounded case may yield very different sets in the unbounded case. The definition for unbounded operators appeared in [
34] and required the development of several new ideas and tools. In this section, we let
\(\Omega _{{\mathcal {C}}}\) denote the set of closed operators
T with domain
\({\mathcal {D}}(T)\subset l^2({\mathbb {N}})\) such that the linear span of the canonical basis forms a core of
T. This latter condition ensures that we can use the usual matrix representation of the operator
T and hence the evaluation functions
\(\Lambda _1\). We follow [
34] and define
$$\begin{aligned} W_e(T)=\left\{ \lambda \in {\mathbb {C}}:\exists \{x_n\}_{n\in {\mathbb {N}}}\subset {\mathcal {D}}(T),\Vert x_n\Vert =1,x_n\xrightarrow {w}0,\lim _{n\rightarrow \infty }\langle Tx_n,x_n\rangle =\lambda \right\} .\nonumber \\ \end{aligned}$$
(7.1)
In [
34], it was shown that for any
\(T\in \Omega _{{\mathcal {C}}}\),
\(W_e(T)\) consists precisely of the essential spectrum of
T together with all possible spectral pollution that may arise by applying projection methods to find the spectrum of
T numerically. This result therefore generalises Theorems
3.8 and
3.9. The set
\(W_e(T)\) is closed and convex, but, unlike the case when
T is bounded,
\(W_e(T)\) may be empty. We first need two simple lemmas.
We have the following corollary, which shows that the SCI classification of computing \(W_e(T)\) for \(T\in \Omega _{{\mathcal {C}}}\) remains \(\Pi _2^A\) (one can make this precise by adding the empty set to the Attouch–Wets topology, but we omit the details).
8 Proofs Concerning Lebesgue Measure
We use the function \(\texttt {DistSpec}\) in “Appendix A”. For ease of notation, we suppress the dispersion function f in calling \(\texttt {DistSpec}\), but assume that we know \(\{c_n\}\) with \(D_{f,n}(A)\le c_n\) and \(c_n\rightarrow 0\) as \(n\rightarrow \infty \). However, the proof of convergence also works when using \(c_n=0\) (which does not necessarily bound \(D_{f,n}(A)\)). The key observation is the following:
Observation: If \(A\in \Omega _f\), then the function \(F_n(z):=\texttt {DistSpec}(A,n,z,f(n))+c_n\) converges uniformly to \(\left\| R(z,A)\right\| ^{-1}\) from above on compact subsets of \({\mathbb {C}}\). By taking successive minima, we can assume without loss of generality that \(F_n\) is non-increasing in n.
The other ingredient needed is the following proposition
Proof of Theorem 3.13
Step 1 \(\{\Xi _1^L,\Omega _f,\Lambda _i\},\{\Xi _1^L,\Omega _\textrm{D},\Lambda _i\}\in \Pi ^A_2\). It is enough to consider
\(\Lambda _1\). We will estimate
\(\textrm{Leb}(\textrm{Sp}(A))\) by estimating the Lebesgue measure of the resolvent set on the closed square
\([-C,C]^2\), where
\(\left\| A\right\| \le C\). We do not assume
C is known. For
\(n_1,n_2\in {\mathbb {N}}\), let
$$\begin{aligned} \texttt {Grid}(n_1,n_2)=\left( \frac{1}{2^{n_2}}{\mathbb {Z}}+\frac{1}{2^{n_2}}i{\mathbb {Z}}\right) \cap [-n_1,n_1]^2. \end{aligned}$$
Letting
B(
x,
r) and
D(
x,
r) denote the closed and open balls of radius
r around
x, respectively,
18 in
\({\mathbb {C}}\) (or
\({\mathbb {R}}\) where appropriate), we define
$$\begin{aligned} U(n_1,n_2,A)=[-n_1,n_1]\times [-n_1,n_1]\cap (\cup _{z\in \texttt {Grid}(n_1,n_2)}B(z,F_{n_1}(z))). \end{aligned}$$
Note that
\(\textrm{Leb}(U(n_1,n_2,A))\) can be computed up to arbitrary predetermined precision using only arithmetic operations and comparisons by Proposition
8.1. Using this, we can define
$$\begin{aligned} \Gamma _{n_2,n_1}(A)=4n_1^2-\textrm{Leb}(U(n_1,n_2,A)) \end{aligned}$$
where, without loss of generality, we assume that we have computed the exact value of the Lebesgue measure (since we can absorb this error in the first limit).
\(\Gamma _{n_2,n_1}\) are arithmetical algorithms using the fact that
\(\texttt {DistSpec}\) is and the above proposition. The only non-trivial part is convergence. The algorithm is summarised in the routine
LebSpec
in § B.3.
We now show that the algorithm
LebSpec
converges and realises the
\(\Pi _2^A\) classification. There exists a compact set
K such that
\(\left\| R(z,A)\right\| ^{-1}> 1\) on
\(K^c\) and without loss of generality we can make
C larger,
\(C\in {\mathbb {N}}\) and take
\(K=[-C,C]^2\). For
\(n_1\ge C\)$$\begin{aligned} U(n_1,n_2,A)=([-C,C]^2\cap (\cup _{z\in \texttt {Grid}(n_1,n_2)}B(z,F_{n_1}(z))))\cup ([-n_1,n_1]^2\backslash [-C,C]^2), \end{aligned}$$
since
\(F_{n}(z)\ge \left\| R(z,A)\right\| ^{-1}\). It follows that for large
\(n_1\)$$\begin{aligned} \Gamma _{n_2,n_1}(A)=4C^2-\textrm{Leb}([-C,C]^2\cap (\cup _{z\in \texttt {Grid}(n_1,n_2)}B(z,F_{n_1}(z)))). \end{aligned}$$
As
\(n_1\rightarrow \infty \),
\([-C,C]^2\cap (\cup _{z\in \texttt {Grid}(n_1,n_2)}B(z,F_{n_1}(z)))\) converges to the closed set
$$\begin{aligned} K(n_2,A)=[-C,C]^2\cap (\cup _{z\in \texttt {Grid}(C,n_2)}B(z,\left\| R(z,A)\right\| ^{-1})) \end{aligned}$$
from above and hence
$$\begin{aligned} \lim _{n_1\rightarrow \infty }\Gamma _{n_2,n_1}(A)=4C^2-\textrm{Leb}(K(n_2,A)), \end{aligned}$$
from below. Consider the relatively open set
$$\begin{aligned} V(n_2,A)=[-C,C]^2\cap (\cup _{z\in \texttt {Grid}(C,n_2)}D(z,\left\| R(z,A)\right\| ^{-1})). \end{aligned}$$
Clearly,
\(\textrm{Leb}(K(n_2,A))=\textrm{Leb}(V(n_2,A))\) since the sets differ by a finite collection of circular arcs or points (recall we defined the open ball of radius zero to be the empty set). Hence, we must show that
$$\begin{aligned} \lim _{n_2\rightarrow \infty }\textrm{Leb}(V(n_2,A))=\textrm{Leb}(\rho _{C}(A)), \end{aligned}$$
where
\(\rho _{C}(A)=[-C,C]^2\backslash \textrm{Sp}(A)\). For
\(z\in \rho _C(A)\),
$$\begin{aligned} \textrm{dist}(z,\textrm{Sp}(A))\ge \left\| R(z,A)\right\| ^{-1} \end{aligned}$$
and hence we get
\(V(n_2,A)\subset \rho _{C}(A)\). Since
\(\rho _C(A)\) is relatively open, a simple density argument using the continuity of
\(\left\| R(z,A)\right\| ^{-1}\) yields
\(V(n_2,A)\uparrow \rho _{C}(A)\) as
\(n_2\rightarrow \infty \) since the grid refines itself. So we get
$$\begin{aligned} \textrm{Leb}(V(n_2,A))\uparrow \textrm{Leb}(\rho _{C}(A)). \end{aligned}$$
This proves the convergence and also shows that
\(\Gamma _{n_2}(A)\downarrow \Xi _1^L(A)\), thus yielding the
\(\Pi _2^A\) classification. The same argument works in the one-dimensional case when considering self-adjoint operators
\(\Omega _{\textrm{D}}\) and
\(\textrm{Leb}_{{\mathbb {R}}}\). We simply restrict everything to the real line and consider the interval
\([-C,C]\) rather than a square.
Step 2 \(\{\Xi _1^L,\Omega _f,\Lambda _i\},\{\Xi _1^L,\Omega _\textrm{D},\Lambda _i\}\notin \Delta ^G_2\). It is enough to consider \(\Lambda _2\). We will only show that \(\textrm{SCI}(\Xi _1^L,\Omega _\textrm{D},\Lambda _2)_{G} \ge 2\) for which we use \(\textrm{Leb}_{{\mathbb {R}}}\) and the two-dimensional case is similar. Suppose for a contradiction that there exists a height one tower \(\{\Gamma _n\}\), then \(\Lambda _{\Gamma _n}(A)\) is finite for each \(A\in \Omega _\textrm{D}\). Hence, for every A and n there exists a finite number \(N(A,n)\in {\mathbb {N}}\) such that the evaluations from \(\Lambda _{\Gamma _n}(A)\) only take the matrix entries \(A_{ij} = \left\langle Ae_j, e_i\right\rangle \) with \(i,j\le N(A,n)\) into account.
Pick any sequence
\(a_1,a_2,\ldots \) that is dense in the unit interval [0, 1]. Consider the matrix operators
\(A_m=\textrm{diag}\{a_1,a_2,\ldots ,a_m\}\in {\mathbb {C}}^{m\times m}\),
\(B_m=\textrm{diag}\{0,0,\ldots ,0\}\in {\mathbb {C}}^{m\times m}\) and
\(C=\textrm{diag}\{0,0,\ldots \}\). Set
\(A=\bigoplus _{m=1}^{\infty }(B_{k_m}\oplus A_{k_m})\), where we choose an increasing sequence
\(k_m\) inductively as follows. Set
\(k_1=1\) and suppose that
\(k_1,\ldots ,k_m\) have been chosen.
\(\textrm{Sp}(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C)=\{0,a_1,a_2,\ldots ,a_{k_m}\}\) and hence
\(\textrm{Leb}(\textrm{Sp}(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C))=0\) so there exists some
\(n_m\ge m\) such that if
\(n\ge n_m\) then
$$\begin{aligned} \Gamma _n(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C)\le \frac{1}{2}. \end{aligned}$$
Now let
\(k_{m+1}\ge \max \{N(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C,n_m),k_m+1\}\). Any evaluation function
\(f_{i,j}\in \Lambda \) is simply the
\((i,j)^{\textrm{th}}\) matrix entry and hence by construction
$$\begin{aligned} f_{i,j}(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C)=f_{i,j}(A), \end{aligned}$$
for all
\(f_{i,j}\in \Lambda _{\Gamma _{n_m}}(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C)\). By assumption (iii) in Definition
5.1, it follows that
\(\Lambda _{\Gamma _{n_m}}(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C)=\Lambda _{\Gamma _{n_m}}(A)\) and hence by assumption (ii) in the same definition that
\(\Gamma _{n_m}(A)=\Gamma _{n_m}(B_{k_1}\oplus A_{k_1}\oplus \cdots \oplus B_{k_m}\oplus A_{k_m}\oplus C)\le 1/2\). But
\(\lim _{n\rightarrow \infty }(\Gamma _n(A))=\textrm{Leb}(\overline{\{0,a_1,a_2,\ldots \}})=1\), a contradiction.
Step 3 \(\{\Xi _1^L,\Omega ,\Lambda _1\}\in \Pi ^A_3\) for
\(\Omega =\Omega _{\textrm{B}},\Omega _\textrm{SA}\),
\(\Omega _\textrm{N}\) or
\(\Omega _g\). We will deal with the case of
\(\Omega _{\textrm{B}}\). The cases of
\(\Omega _\textrm{N}\) and
\(\Omega _g\) then follow via
\(\Omega _{\textrm{N}}\subset \Omega _{g}\subset \Omega _{\textrm{B}}\) and the one-dimensional Lebesgue measure case for
\(\Omega _\textrm{SA}\) is similar. A careful analysis of the proof in step 1 yields that
-
\(\Gamma _{n_2,n_1}(A)\) converges to \(\Gamma _{n_2}(A)\) from below as \(n_1\rightarrow \infty \).
-
\(\Gamma _{n_2}(A)\) converges to \(\textrm{Leb}(\textrm{Sp}(A))\) monotonically from above as \(n_2\rightarrow \infty \).
We can ensure that the first limit converges from below by always slightly overestimating the Lebesgue measure of
\(U(n_1,n_2)\) (with error converging to zero) and using Proposition
8.1. These observations will be used later to answer question 3. We do not need to know
\(c_n\) for the above proof to work, but we will need it for the first of the above facts. A slight alteration of the proof/algorithm by inserting an additional successive limit deals with the general case.
Define the function
$$\begin{aligned} \gamma _{n,m}(z;A)=\min \{\sigma _{\textrm{inf}}(P_m(A-zI){|_{P_n{\mathcal {H}}}}),\sigma _{\textrm{inf}}(P_m(A^*-{\bar{z}}I){|_{P_n{\mathcal {H}}}})\}, \end{aligned}$$
where
\(\sigma _{\textrm{inf}}\) denotes the injection modulus/smallest singular value (see also “Appendix A”). One can show that
\(\gamma _{n,m}\) converges uniformly on compact subsets to
$$\begin{aligned} \gamma _{n}(z;A)=\min \{\sigma _{\textrm{inf}}((A-zI){|_{P_n{\mathcal {H}}}}),\sigma _{\textrm{inf}}((A^*-{\bar{z}}I){|_{P_n{\mathcal {H}}}})\}, \end{aligned}$$
as
\(m\rightarrow \infty \) and that this converges uniformly down to
\(\left\| R(z,A)\right\| ^{-1}\) on compact subsets as
\(n\rightarrow \infty \) [
91]. With a slight abuse of notation, we can approximate
\(\gamma _{n,m}(z;A)\) to within 1/
m by
\(\texttt {DistSpec}(A,n,z,m)\) (where the spacing of the search routine is 1/
m, see also “Appendix A”) so that this converges uniformly on compact subsets to
\(\gamma _{n}(z;A)\). In exactly the same manner as before, define
$$\begin{aligned} U(n_1,n_2,n_3,A)&=[-n_2,n_2]^2\cap (\cup _{z\in \texttt {Grid}(n_2,n_3)}B(z,\gamma _{n_2,n_1}(z;A))),\\ \Gamma _{n_3,n_2,n_1}(A)&=(2n_2)^2-\textrm{Leb}(U(n_1,n_2,n_3,A)). \end{aligned}$$
The stated uniform convergence means that the argument in step 1 carries through and we have a height three tower, realising the
\(\Pi ^A_3\) classification.
Step 4 \(\{\Xi _1^L,\Omega _\textrm{SA},\Lambda _1\}\notin \Delta ^G_3\). The proof is exactly the same argument as the proof of step 3 of Theorem
3.7. However, in this case to gain the contradiction, we then define
\({\tilde{\Gamma }}_{n_2,n_1}(\{a_{i,j}\})=\min \{\max \{1-\Gamma _{n_2,n_1}(A)/2,0\},1\}\) where
\(\{\Gamma _{n_2,n_1}\}\) is the supposed height two tower for
\(\{\Xi _1^L,\Omega _\textrm{SA},\Lambda _1\}\).
Step 5 \(\{\Xi _1^L,\Omega ,\Lambda _1\}\notin \Delta ^G_3\) for
\(\Omega =\Omega _{\textrm{B}},\Omega _{\textrm{N}}\), or
\(\Omega _g\). Since
\(\Omega _\textrm{N}\subset \Omega _g\subset \Omega _{\textrm{B}}\), we only need to deal with
\(\Omega _\textrm{N}\). We can use a similar argument as in step 4, but now replacing each
\(C^{(j)}\) by
$$\begin{aligned} D^{(j)}=\bigoplus _{k=1}^j ih_k C^{(j)}, \end{aligned}$$
where
\(h_1,h_2,\ldots \) is a dense sequence in [0, 1], and these operators act on
\(X_j=\bigoplus _{k=1}^j l^2({\mathbb {N}})\). This ensures that the spectrum of the operator yields a positive two-dimensional Lebesgue measure if and only if
\({{\tilde{\Xi }}}_2(\{a_{i,j}\})=0\). The rest of the argument is entirely analogous.
Step 6 \(\Delta ^G_2 \not \ni \{\Xi _1^L,\Omega ,\Lambda _2\}\in \Pi ^A_2\) for \(\Omega =\Omega _{\textrm{B}},\Omega _\textrm{SA}\), \(\Omega _\textrm{N}\) or \(\Omega _g\). The impossibility result follows by considering diagonal operators. For the existence of \(\Pi ^A_2\) algorithms, we can use the construction in step 3, but the knowledge of matrix values of \(A^*A\) allows us to skip the first limit and approximate \(\gamma _n\) directly. \(\square \)
Finally, we deal with the question of determining whether the Lebesgue measure is zero. Recall that for this problem,
\(({\mathcal {M}},d)\) denotes the set
\(\{0,1\}\) endowed with the discrete topology and we consider the problem function
$$\begin{aligned} \Xi _3^L(A)= {\left\{ \begin{array}{ll} 0,\quad \text { if }\textrm{Leb}(\textrm{Sp}(A))>0\\ 1,\quad \text { otherwise.} \end{array}\right. } \end{aligned}$$
Proof of Theorem 3.16
We will show that \(\{\Xi _3^L,\Omega _f,\Lambda _1\}\in \Pi _3^A\) and \(\{\Xi _3^L,\Omega _\textrm{D},\Lambda _2\}\notin \Delta _3^G\). The analogous statements \(\{\Xi _3^L,\Omega _\textrm{D},\Lambda _1\}\in \Pi _3^A\) and \(\{\Xi _3^L,\Omega _f,\Lambda _2\}\notin \Delta _3^G \) follow from similar arguments.
The lower bound argument can also be used when considering
\(\Lambda _2\) and
\(\Omega =\Omega _{\textrm{B}},\Omega _\textrm{SA}\),
\(\Omega _\textrm{N}\) or
\(\Omega _g\). We will also prove the lower bound
\(\{\Xi _3^L,\Omega _\textrm{SA},\Lambda _1\}\notin \Delta _4^G\). The remaining lower bounds for
\(\Lambda _1\) follow from a similar argument and construction as in step 5 of the proof of Theorem
3.13 to ensure we are dealing with two-dimensional Lebesgue measure. Finally, we prove that
\(\{\Xi _3^L,\Omega _{\textrm{B}},\Lambda _1\}\in \Pi _4^A\). The upper bounds for
\(\Omega =\Omega _\textrm{SA}\),
\(\Omega _\textrm{N}\) or
\(\Omega _g\) and
\(\Lambda _1\) follow an almost identical argument. When considering
\(\Lambda _2\), we can collapse the first limit in the same manner as we did for solving
\(\Xi _1^L\).
Step 1 \(\{\Xi _3^L,\Omega _f,\Lambda _1\}\in \Pi _3^A\). First we use the algorithm used to compute
\(\Xi _1^L\) in Theorem
3.13, which we shall denote by
\({\widetilde{\Gamma }}\), to build a height 3 tower for
\(\{\Xi _3^L,\Omega _f\}\). As above,
\(\Omega _f\) denotes the set of bounded operators with the usual assumption of bounded dispersion (now with known bounds
\(c_n\)). Recall that we observed
-
\({\widetilde{\Gamma }}_{n_2,n_1}(A)\) converges to \({\widetilde{\Gamma }}_{n_2}(A)\) from below as \(n_1\rightarrow \infty \).
-
\({\widetilde{\Gamma }}_{n_2}(A)\) converges to \(\textrm{Leb}(\textrm{Sp}(A))\) monotonically from above as \(n_2\rightarrow \infty \).
We can alter our algorithms, by taking maxima, so that we can assume without loss of generality that
\({\widetilde{\Gamma }}_{n_2,n_1}(A)\) converges to
\({\widetilde{\Gamma }}_{n_2}(A)\) monotonically from below as
\(n_1\rightarrow \infty \). Now let
$$\begin{aligned} \Gamma _{n_3,n_2,n_1}(A)=\chi _{[0,1/n_3]}({\widetilde{\Gamma }}_{n_2,n_1}(A)). \end{aligned}$$
Note that
\(\chi _{[0,1/n_3]}\) is left continuous on
\([0,\infty )\) with right limits. Hence by the assumed monotonicity
$$\begin{aligned} \lim _{n_1\rightarrow \infty }\Gamma _{n_3,n_2,n_1}(A)=\chi _{[0,1/n_3]}({\widetilde{\Gamma }}_{n_2}(A)). \end{aligned}$$
It follows that
$$\begin{aligned} \lim _{n_2\rightarrow \infty }\lim _{n_1\rightarrow \infty }\Gamma _{n_3,n_2,n_1}(A)=\chi _{[0,1/n_3]}(\textrm{Leb}(\textrm{Sp}(A))\pm ), \end{aligned}$$
where ± denotes one of the right or left limits (it is possible to have either). It is then easy to see that
$$\begin{aligned} \lim _{n_3\rightarrow \infty }\lim _{n_2\rightarrow \infty }\lim _{n_1\rightarrow \infty }\Gamma _{n_3,n_2,n_1}(A)=\Xi _3^L(A). \end{aligned}$$
It is also clear that the answer to the question is “No” if
\(\Gamma _{n_3}(A)=0\), which yields the
\(\Pi _3^A\) classification.
Step 2 \(\{\Xi _3^L,\Omega _\textrm{D},\Lambda _1\}\notin \Delta _3^G\). Assume for a contradiction that this is false and
\(\{{\widehat{\Gamma }}_{n_2,n_1}\}\) is a general height two tower for
\(\{\Xi _3^L,\Omega _\textrm{D}\}\). Let
\(({\mathcal {M}},d)\) be discrete space
\(\{0,1\}\) and
\({{\tilde{\Omega }}}\) denote the collection of all infinite matrices
\(\{a_{i,j}\}_{i,j\in {\mathbb {N}}}\) with entries
\(a_{i,j}\in \{0,1\}\) and consider the problem function
$$\begin{aligned} {{\tilde{\Xi }}}_1(\{a_{i,j}\}):\text { Does }\{a_{i,j}\}\text { have a column containing infinitely many nonzero entries?} \end{aligned}$$
For
\(j\in {\mathbb {N}}\), let
\(\{b_{i,j}\}_{i\in {\mathbb {N}}}\) be a dense subset of
\(I_j:=[1-1/2^{j-1},1-1/{2^j}]\). Given a matrix
\(\{a_{i,j}\}_{i,j\in {\mathbb {N}}}\in {{\tilde{\Omega }}}\), construct a matrix
\(\{c_{i,j}\}_{i,j\in {\mathbb {N}}}\) by letting
\(c_{i,j}=a_{i,j}b_{r(i,j),j}\) where
$$\begin{aligned} r(i,j)=\max \left\{ 1,\sum _{k=1}^{i}a_{k,j}\right\} . \end{aligned}$$
Now consider any bijection
\(\phi :{\mathbb {N}}\rightarrow {\mathbb {N}}^2\) and define the diagonal operator
$$\begin{aligned} A=\textrm{diag}(c_{\phi (1)},c_{\phi (2)},c_{\phi (3)},\ldots ). \end{aligned}$$
The algorithm
\({\widehat{\Gamma }}_{n_2,n_1}\) thus translates to an algorithm
\(\Gamma '_{n_2,n_1}\) for
\(\{{{\tilde{\Xi }}}_1,{{\tilde{\Omega }}}\}\). Namely, set
\(\Gamma '_{n_2,n_1}(\{a_{i,j}\}_{i\in {\mathbb {N}}})={\widehat{\Gamma }}_{n_2,n_1}(A)\). The fact that
\(\phi \) is a bijection shows that the lowest level
\(\Gamma '_{n_2,n_1}\) are generalised algorithms (and are consistent). In particular, given
N, we can find
\(\{A_{i,j}:i,j\le N\}\) using finitely many evaluations of the matrix values
\(\{c_{k,l}\}\). But for any given
\(c_{k,l}\), we can evaluate this entry using only finitely many evaluations of the matrix values
\(\{a_{m,n}\}\) by the construction of
r. Finally note that
$$\begin{aligned} \textrm{Sp}(A)=\left( \bigcup _{j:\sum _ia_{i,j}=\infty }I_j\right) \cup Q, \end{aligned}$$
where
Q is at most countable. Hence,
$$\begin{aligned} \textrm{Leb}_{{\mathbb {R}}}(\textrm{Sp}(A))=\sum _{j:\sum _ia_{i,j}=\infty }\frac{1}{2^j}. \end{aligned}$$
It follows that
\({{\tilde{\Xi }}}_1(\{a_{i,j}\})=\Xi _3^L(A)\), and hence, we get a contradiction.
Step 3 \(\{\Xi _3^L,\Omega _\textrm{SA},\Lambda _1\}\notin \Delta _4^G\). Suppose for a contradiction that
\(\{\Gamma _{n_3,n_2,n_1}\}\) is a height three tower of general algorithms for the problem
\(\{\Xi _3^L,\Omega _\textrm{SA},\Lambda _1\}\). Let
\(({\mathcal {M}},d)\) be the space
\(\{0,1\}\) with the discrete metric, let
\({\tilde{\Omega }}\) denote the collection of all infinite arrays
\(\{a_{m,i,j}\}_{m,i,j\in {\mathbb {N}}}\) with entries
\(a_{m,i,j}\in \{0,1\}\) and consider the problem function
$$\begin{aligned} \begin{aligned} {\tilde{\Xi }}_4(\{a_{m,i,j}\}):\text { For every { m}, does }\{a_{m,i,j}\}_{i,j}&\text { have (only) finitely many columns}\\&\quad \quad \quad \quad \quad \quad \quad \text {with (only) finitely many 1's?} \end{aligned} \end{aligned}$$
Recall that it is shown in Sect.
5 that
\(\textrm{SCI}({\tilde{\Xi }}_4,{\tilde{\Omega }})_{G} = 4\). We will gain a contradiction by using the supposed height three tower to solve
\(\{{\tilde{\Xi }}_4,{\tilde{\Omega }}\}\).
The construction follows step 3 of the proof of Theorem
3.7 closely. For fixed
m, recall the construction of the operator
\(A_m:=A(\{a_{m,i,j}\}_{i,j})\) from that proof, the key property being that if
\(\{a_{m,i,j}\}_{i,j}\) has (only) finitely many columns with (only) finitely many 1’s then
\(\textrm{Sp}(A_m)\) is a finite subset of
\([-1,1]\), otherwise it is the whole interval
\([-1,1]\). Now consider the intervals
\(I_m=[1-2^{m-1},1-2^{m}]\) and affine maps,
\(\alpha _m\), that act as a bijection from
\([-1,1]\) to
\(I_m\). Without loss of generality, identify
\(\Omega _\textrm{SA}\) with self-adjoint operators in
\({\mathcal {B}}(X)\) where
\(X=\bigoplus _{i=1}^{\infty }\bigoplus _{j=1}^{\infty }X_{i,j}\) in the
\(l^2\)-sense with
\(X_{i,j}=l^2({\mathbb {N}})\). We then consider the operator
$$\begin{aligned} T(\{a_{m,i,j}\}_{m,i,j})=\bigoplus _{m=1}^\infty \alpha _{m}(A_m). \end{aligned}$$
The same arguments in the proof of Theorem
3.7 show that the map
$$\begin{aligned} {{\tilde{\Gamma }}}_{n_3,n_2,n_1}(\{a_{m,i,j}\}_{m,i,j})=\Gamma _{n_3,n_2,n_1}(T(\{a_{m,i,j}\}_{m,i,j})) \end{aligned}$$
defines a general tower using the relevant pointwise evaluation functions of the array
\(\{a_{m,i,j}\}_{m,i,j}\). If it holds that
\({\tilde{\Xi }}_4(\{a_{m,i,j}\})=1\), then
\(\textrm{Sp}(T(\{a_{m,i,j}\}_{m,i,j}))\) is countable and hence
\(\Xi _3^L(T(\{a_{m,i,j}\}_{m,i,j}))=1\). On the other hand, if
\({\tilde{\Xi }}_4(\{a_{m,i,j}\})=0\), then there exists
m with
\(\textrm{Sp}(A_m)=[-1,1]\), and hence,
\(I_m\subset \textrm{Sp}(T(\{a_{m,i,j}\}_{m,i,j}))\) so that
\(\Xi _3^L(T(\{a_{m,i,j}\}_{m,i,j}))=0\). It follows that
\(\{{{\tilde{\Gamma }}}_{n_3,n_2,n_1}\}\) provides a height three tower for
\(\{{\tilde{\Xi }}_4,{\tilde{\Omega }}\}\), a contradiction.
Step 4 \(\{\Xi _3^L,\Omega _{\textrm{B}},\Lambda _1\}\in \Pi _4^A\). Recall the tower of algorithms to solve
\(\{\Xi _1^L,\Omega _{\textrm{B}},\Lambda _1\}\) and denote it by
\({\widetilde{\Gamma }}\). Our strategy will be the same as in step 1 but with an additional successive limit. It is easy to show that
-
\({\widetilde{\Gamma }}_{n_3,n_2,n_1}(A)\) converges to \({\widetilde{\Gamma }}_{n_3,n_2}(A)\) from above as \(n_1\rightarrow \infty \).
-
\({\widetilde{\Gamma }}_{n_3,n_2}(A)\) converges to \({\widetilde{\Gamma }}_{n_3}(A)\) from below as \(n_2\rightarrow \infty \).
-
\({\widetilde{\Gamma }}_{n_3}(A)\) converges to \(\textrm{Leb}(\textrm{Sp}(A))\) from above as \(n_3\rightarrow \infty \).
Again, by taking successive maxima or minima where appropriate, we can assume that all of these are monotonic. Now let
$$\begin{aligned} \Gamma _{n_4,n_3,n_2,n_1}(A)=\chi _{[0,1/n_4]}({\widetilde{\Gamma }}_{n_3,n_2,n_1}(A)). \end{aligned}$$
Note that
\(\chi _{[0,1/n_4]}\) is left continuous on
\([0,\infty )\) with right limits. Hence by the assumed monotonicity and arguments as in step 1, it is easy to see that
$$\begin{aligned} \lim _{n_4\rightarrow \infty }\lim _{n_3\rightarrow \infty }\lim _{n_2\rightarrow \infty }\lim _{n_1\rightarrow \infty }\Gamma _{n_4,n_3,n_2,n_1}(A)=\Xi _3^L(A). \end{aligned}$$
It is also clear that the answer to the question is “No” if
\(\Gamma _{n_4}(A)=0\), which yields the
\(\Pi _4^A\) classification.
\(\square \)
9 Proofs Concerning Fractal Dimensions
We begin with the box-counting dimension. For the construction of towers of algorithms, it is useful to use a slightly different but equivalent [
71] definition of the upper and lower box-counting dimensions. Let
\(F\subset {\mathbb {R}}\) be bounded and
\(N'_\delta (F)\) denote the number of
\(\delta \)-mesh intervals that intersect
F. A
\(\delta \)-mesh interval is an interval of the form
\([m\delta ,(m+1)\delta ]\) for
\(m\in {\mathbb {Z}}\). Then
$$\begin{aligned} \overline{\textrm{dim}}_B(F)=\limsup _{\delta \downarrow {}0}\frac{\log (N'_{\delta }(F))}{\log (1/\delta )},\quad \underline{\textrm{dim}}_B(F)=\liminf _{\delta \downarrow {}0}\frac{\log (N'_{\delta }(F))}{\log (1/\delta )}. \end{aligned}$$
Proof of Theorem 3.18 for box-counting dimension
Since \(\Omega _{BD}^\textrm{D}\subset \Omega ^{BD}_f\subset \Omega ^{BD}_{\textrm{SA}}\), it is enough to prove that \(\{\Xi _B,\Omega ^{BD}_f,\Lambda _1\}\in \Pi _2^A\), \(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _2\}\in \Pi _2^A\), \(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _1\}\in \Pi _3^A\), \(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _1\}\not \in \Delta _3^A\) and \(\{\Xi _B,\Omega ^{BD}_\textrm{D},\Lambda _2\}\not \in \Delta _2^A\).
Step 1 \(\{\Xi _B,\Omega ^{BD}_f,\Lambda _1\}\in \Pi _2^A\). Recall the existence of a height one tower,
\(\{{{\tilde{\Gamma }}}_n\}\), using
\(\Lambda _1\) for
\(\textrm{Sp}(A)\),
\(A\in \Omega ^{BD}_f\) from “Appendix A”. Furthermore,
\({{\tilde{\Gamma }}}_n(A)\) outputs a finite collection
\(\{z_{1,n},\ldots ,z_{k_n,n}\}\subset {\mathbb {Q}}\) such that
\(\textrm{dist}(z_{j,n},\textrm{Sp}(A))\le 2^{-n}\). Define the intervals
$$\begin{aligned} I_{j,n}=[z_{j,n}-2^{-n},z_{j,n}+2^{-n}] \end{aligned}$$
and let
\({\mathcal {I}}_{m}\) denote the collection of all
\(2^{-m}\)-mesh intervals. Let
\(\Upsilon _{m,n}(A)\) be any union of finitely many such mesh intervals with minimal length
\(\left| \Upsilon _{m,n}(A)\right| \) (“length” being the number of intervals
\(\in {\mathcal {I}}_{m}\) that make up
\(\Upsilon _{m,n}(A)\)) such that
$$\begin{aligned} \Upsilon _{m,n}(A)\cap I_{j,l}\ne \emptyset ,\quad \text {for }1\le l\le n,1\le j\le k_l. \end{aligned}$$
There may be more than one such collection, so we can gain a deterministic algorithm by enumerating each
\({\mathcal {I}}_{m}\) and choosing the first such collection in this enumeration. It is then clear that
\(\left| \Upsilon _{m,n}(A)\right| \) is increasing in
n. Furthermore, to determine
\(\Upsilon _{m,n}(A)\), there are only finitely many intervals in
\({\mathcal {I}}_{m}\) to consider, namely those that have non-empty intersection with at least one
\(I_{j,l}\) with
\(1\le l\le n,1\le j\le k_l\). It follows that
\(\Upsilon _{m,n}(A)\) and hence
\(\left| \Upsilon _{m,n}(A)\right| \) can be computed in finitely may arithmetic operations and comparisons using
\(\Lambda _1\).
Suppose that
\(I=[a,b]\in {\mathcal {I}}_{m}\) has
\((a,b)\cap \textrm{Sp}(A)\ne \emptyset \). Then for large
n there exists
\(z_{j,n}\in I\) such that
\(I_{j,n}\subset I\) and hence
\(I\subset \Upsilon _{m,n}(A)\) for large
n. If
\(z\in \textrm{Sp}(A)\cap 2^{-m}{\mathbb {Z}}\), then a similar argument shows that
\(z\subset \Upsilon _{m,n}(A)\) for large
n. Since
\(\textrm{Sp}(A)\) is bounded and
\(\textrm{Sp}(A)\cap 2^{-m}{\mathbb {Z}}\) finite, it follows that
\(\textrm{Sp}(A)\subset \Upsilon _{m,n}(A)\) for large
n and hence
$$\begin{aligned} N_{2^{-m}}(\textrm{Sp}(A))\le \liminf _{n\rightarrow \infty }\left| \Upsilon _{m,n}(A)\right| . \end{aligned}$$
Let
\(W_m(A)\) be the union of all intervals in
\({\mathcal {I}}_m\) that intersect
\(\textrm{Sp}(A)\). It is clear that
\(W_m(A)\cap I_{j,l}\ne \emptyset \) for
\(1\le l\le n,1\le j\le k_l\), and hence,
\(\left| \Upsilon _{m,n}(A)\right| \le N'_{2^{-m}}(\textrm{Sp}(A))\). It follows that
\(\lim _{n\rightarrow \infty }\left| \Upsilon _{m,n}(A)\right| =\delta _{m}(A)\) exists with
$$\begin{aligned} N_{2^{-m}}(\textrm{Sp}(A))\le \delta _{m}(A)\le N'_{2^{-m}}(\textrm{Sp}(A)). \end{aligned}$$
(9.1)
For
\(n_2>n_1\) set
\(\Gamma _{n_2,n_1}(A)=0\), otherwise set
$$\begin{aligned} \Gamma _{n_2,n_1}(A)=\max _{n_2\le k \le n_1} \max _{1\le j\le n_1}\frac{\log (\left| \Upsilon _{k,j}(A)\right| )}{k\log (2)}. \end{aligned}$$
The above monotone convergence and (
9.1) shows that
$$\begin{aligned} \lim _{n_1\rightarrow \infty }\Gamma _{n_2,n_1}(A)&=\Gamma _{n_2}(A)=\sup _{k\ge n_2}\frac{\log (\delta _k(A))}{k\log (2)}\ge \limsup _{k\rightarrow \infty }\frac{\log (\delta _k(A))}{k\log (2)},\\ \lim _{n_2\rightarrow \infty }\Gamma _{n_2}(A)&=\limsup _{k\rightarrow \infty }\frac{\log (\delta _k(A))}{k\log (2)}. \end{aligned}$$
Hence, by the assumption that the box-counting dimension exists, we have constructed a
\(\Pi _2^A\) tower.
Step 2 \(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _2\}\in \Pi _2^A\) and
\(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _1\}\in \Pi _3^A\). The first of these is exactly as in step 1, using
\(\Lambda _2\) to construct the relevant
\(\Sigma _1^A\) tower for the spectrum. The proof that
\(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _1\}\in \Pi _3^A\) uses a height two tower,
\(\{{{\tilde{\Gamma }}}_{n_2,n_1}\}\), using
\(\Lambda _1\) for
\(\textrm{Sp}(A)\),
\(A\in \Omega _{\textrm{SA}}^{BD}\) (or any self-adjoint
A) constructed in [
18]. This tower has the property that each
\({{\tilde{\Gamma }}}_{n_2,n_1}(A)\) is a finite subset of
\({\mathbb {Q}}\) and, for fixed
\(n_2\), is constant for large
\(n_1\). Moreover, if
\(z\in \lim _{n_1\rightarrow \infty }{{\tilde{\Gamma }}}_{n_2,n_1}(A)\), then
\(\textrm{dist}(z,\textrm{Sp}(A))\le 2^{-n_2}\). It follows that we can use the same construction as step 1 with an additional limit at the start to reach the finite set
\(\lim _{n_1\rightarrow \infty }{{\tilde{\Gamma }}}_{n_2,n_1}(A)\).
Step 3 \(\{\Xi _B,\Omega ^{BD}_\textrm{D},\Lambda _2\}\not \in \Delta _2^A\). This is exactly the same argument as step 2 of the proof of Theorem
3.13 with Lebesgue measure replaced by box-counting dimension.
Step 4 \(\{\Xi _B,\Omega _{\textrm{SA}}^{BD},\Lambda _1\}\not \in \Delta _3^A\). This is exactly the same argument as step 4 of the proof of Theorem
3.13 with Lebesgue measure replaced by box-counting dimension.
\(\square \)
We now turn to the Hausdorff dimension. Recall Lemma
3.21 on the problem of determining whether
\(\textrm{Sp}(A)\cap (a,b)\ne \emptyset \).
To build our algorithm for the Hausdorff dimension, we use an alternative, equivalent definition for compact sets. We consider the case of subsets of
\({\mathbb {R}}\). Let
\(\rho _k\) denote the set of all closed binary intervals of the form
\( [2^{-k}m,2^{-k}(m+1)], m\in {\mathbb {Z}}. \) Set
$$\begin{aligned} {\mathcal {A}}_k(F)=\left\{ \{U_i\}_{i\in I}:I\text { is finite },F\subset \cup _{i\in I}U_i,U_i\in \cup _{l\ge k}\rho _l\right\} \end{aligned}$$
and define
$$\begin{aligned} \tilde{{\mathcal {H}}}^{d}_{k}(F)=\inf \left\{ \sum _i\textrm{diam}(U_i)^d:\{U_i\}_{i\in I}\in {\mathcal {A}}_{k}(F)\right\} ,\quad \tilde{{\mathcal {H}}}^{d}(F)=\lim _{k\rightarrow \infty }\tilde{{\mathcal {H}}}^{d}_{k}(F). \end{aligned}$$
The following can be found in [
81] (Theorem 3.13):
Denoting the dyadic rationals by
\({\mathbb {D}}\), we shall compute
\(\textrm{dim}_H(\textrm{Sp}(A))\) via approximating the above applied to
\(F=\textrm{Sp}(A)\cap {\mathbb {D}}^c\) and using Lemma
3.21.
Proof of Theorem 3.18 for Hausdorff dimension
It is enough to prove the lower bounds \(\{\Xi _H,\Omega _\textrm{D},\Lambda _2\}\notin \Delta ^G_3\), \(\{\Xi _H,\Omega _\textrm{SA},\Lambda _1\}\notin \Delta ^G_4\) and construct the towers of algorithms for the inclusions \(\{\Xi _H,\Omega _f\cap \Omega _\textrm{SA},\Lambda _1\}\in \Sigma ^A_3\), \(\{\Xi _H,\Omega _\textrm{SA},\Lambda _1\}\in \Sigma _4^A\) and \(\{\Xi _H,\Omega _\textrm{SA},\Lambda _2\}\in \Sigma _3^A\).
Step 1 \(\{\Xi _H,\Omega _\textrm{D},\Lambda _2\}\notin \Delta ^G_3\). Suppose for a contradiction that a height two tower,
\(\{\Gamma _{n_2,n_1}\}\), exists for
\(\{\Xi _H,\Omega _\textrm{D}\}\) (taking values in [0, 1] without loss of generality). We repeat the argument in the proof of Theorem
3.16. Consider the same problem
$$\begin{aligned} {{\tilde{\Xi }}}_1(\{a_{i,j}\}):\text { Does }\{a_{i,j}\}\text { have a column containing infinitely many nonzero entries?} \end{aligned}$$
However, now we consider the above mapping to [0, 1] with the usual metric. We consider the same operator
\( A=\textrm{diag}(c_{\phi (1)},c_{\phi (2)},c_{\phi (3)},\ldots ) \) with
$$\begin{aligned} \textrm{Sp}(A)=\left( \bigcup _{j:\sum _{i}a_{i,j}=\infty }I_j\right) \cup Q, \end{aligned}$$
where
Q is at most countable. We use the fact that the Hausdorff dimension satisfies
$$\begin{aligned} \textrm{dim}_H(\cup _{j=1}^\infty X_j)=\sup _{j\in {\mathbb {N}}}\textrm{dim}_H(X_j) \end{aligned}$$
and that
\(\textrm{dim}_H(Q)=0\) for any countable
Q to note that
\(\Xi _H(A)={{\tilde{\Xi }}}_1(\{a_{i,j}\})\). We set
\({\tilde{\Gamma }}_{n_2,n_1}(\{a_{i,j}\}_{i,j})=\Gamma _{n_2,n_1}(A)\) to provide a height two tower for
\({{\tilde{\Xi }}}_1\). But this contradicts Theorem
5.19.
Step 2 \(\{\Xi _H,\Omega _\textrm{SA},\Lambda _1\}\notin \Delta ^G_4\). Suppose for a contradiction that
\(\{\Gamma _{n_3,n_2,n_1}\}\) is a height three tower of general algorithms for the problem
\(\{\Xi _H,\Omega _\textrm{SA},\Lambda _1\}\) (taking values in [0, 1] without loss of generality). Let
\(({\mathcal {M}},d)\) be the space [0, 1] with the usual metric, let
\({\tilde{\Omega }}\) denote the collection of all infinite arrays
\(\{a_{m,i,j}\}_{m,i,j\in {\mathbb {N}}}\) with entries
\(a_{m,i,j}\in \{0,1\}\) and consider the problem function
$$\begin{aligned} \begin{aligned}&{\tilde{\Xi }}_4(\{a_{m,i,j}\}):\text { For every { m}, does }\{a_{m,i,j}\}_{i,j}\text { have (only) finitely many columns}\\&\quad \text {with (only) finitely many 1's?} \end{aligned} \end{aligned}$$
Recall that it is shown in Sect.
5 that
\(\textrm{SCI}({\tilde{\Xi }}_4,{\tilde{\Omega }})_{G} = 4\). We will gain a contradiction by using the supposed height three tower to solve
\(\{{\tilde{\Xi }}_4,{\tilde{\Omega }}\}\). We use the same construction as in step 3 of the proof of Theorem
3.16. If
\({\tilde{\Xi }}_4(\{a_{m,i,j}\})=1\), then
\(\textrm{Sp}(T(\{a_{m,i,j}\}_{m,i,j}))\) is countable, and hence,
\(\Xi _H(T(\{a_{m,i,j}\}_{m,i,j}))=0\). On the other hand, if
\({\tilde{\Xi }}_4(\{a_{m,i,j}\})=0\), then there exists
m with
\(\textrm{Sp}(A_m)=[-1,1]\) and hence
\(I_m\subset \textrm{Sp}(T(\{a_{m,i,j}\}_{m,i,j}))\) so that
\(\Xi _H(T(\{a_{m,i,j}\}_{m,i,j}))=1\). It follows that
\({{\tilde{\Gamma }}}_{n_3,n_2,n_1}(\{a_{m,i,j}\}_{m,i,j})=1-\Gamma _{n_3,n_2,n_1}(T(\{a_{m,i,j}\}_{m,i,j}))\) provides a height three tower for
\(\{{\tilde{\Xi }}_4,{\tilde{\Omega }}\}\), a contradiction.
Step 3 \(\{\Xi _H,\Omega _f\cap \Omega _\textrm{SA},\Lambda _1\}\in \Sigma _3^A\). To construct a height three tower for
\(A\in \Omega _f\cap \Omega _\textrm{SA}\), if
\(n_2<n_3\) set
\(\Gamma _{n_3,n_2,n_1}(A)=0\). Otherwise, consider the set
$$\begin{aligned} {\mathcal {A}}_{n_3,n_2,n_1}(A)=\left\{ \{U_i\}_{i\in I}:I\text { is finite },S_{n_1,n_2}(A)\subset \cup _{i\in I}U_i,U_i\in \cup _{n_3\le l\le n_2}\rho _l\right\} \end{aligned}$$
where
\(S_{n_1,n_2}(A)\) is the union of all
\(S\in \rho _{n_2}\) with
\(S\subset [-n_1,n_1]\) and such that the algorithm discussed in Lemma
3.21 outputs “Yes” for the interior of
S and input parameter
\(n_1\). We then define
$$\begin{aligned} h_{n_3,n_2,n_1}(A,d)=\inf \left\{ \sum _i\textrm{diam}(U_i)^d:\{U_i\}\in {\mathcal {A}}_{n_3,n_2,n_1}(A)\right\} . \end{aligned}$$
If
\(S_{n_1,n_2}(A)\) is empty, then we interpret the infinum as 0. There are only finitely many sets to check and hence the infinum is a minimisation problem over finitely many coverings (see § B.4 for a discussion of efficient implementation). It follows that
\(h_{n_3,n_2,n_1}(A,d)\) defines a general algorithm computable in finitely many arithmetic operations and comparisons. Furthermore, it is easy to see that
$$\begin{aligned} \lim _{n_1\rightarrow \infty }h_{n_3,n_2,n_1}(A,d)=\inf \left\{ \sum _i\textrm{diam}(U_i)^d:\{U_i\}\in {\mathcal {C}}_{n_3,n_2}(A)\right\} =:h_{n_3,n_2}(A,d) \end{aligned}$$
from below (since we are covering larger sets as
\(n_1\) increases). Here,
$$\begin{aligned} {\mathcal {C}}_{n_3,n_2}(A)=\left\{ \{U_i\}_{i\in I}:I\text { is finite },\textrm{Sp}(A)\cap {\mathbb {D}}_{n_2}^c\subset \cup _{i\in I}U_i,U_i\in \cup _{n_3\le l\le n_2}\rho _l\right\} \end{aligned}$$
and
\({\mathbb {D}}_k:=1/{2^k}\cdot {\mathbb {Z}}\) denotes the dyadic rationals of resolution
k. We now use the property that
\({\mathcal {A}}_k(F)\) consists of collections of finite coverings. As
\(n_2\rightarrow \infty \),
\(h_{n_3,n_2}(A,d)\) is non-increasing (since we take infinum over a larger class of coverings and the sets
\(\textrm{Sp}(A)\cap {\mathbb {D}}_{n_2}^c\) decrease) and hence converges to some number. Clearly
$$\begin{aligned} \lim _{n_2\rightarrow \infty }h_{n_3,n_2}(A,d)=:h_{n_3}(A,d)\ge \tilde{{\mathcal {H}}}^d_{n_3}(\textrm{Sp}(A)\cap {\mathbb {D}}^c). \end{aligned}$$
For
\(\epsilon >0\), let
\(l\in {\mathbb {N}}\) and
\(\{U_i\}\in {\mathcal {A}}_{n_3}(\textrm{Sp}(A)\cap {\mathbb {D}}_l^c)\}\) with
$$\begin{aligned} \sum _{i}\textrm{diam}(U_i)^d\le \epsilon +\tilde{{\mathcal {H}}}^d_{n_3}(\textrm{Sp}(A)\cap {\mathbb {D}}_l^c). \end{aligned}$$
For large enough
\(n_2\),
\(\{U_i\}\in {\mathcal {C}}_{n_3,n_2}(A)\) and hence since
\(\epsilon >0\) was arbitrary,
$$\begin{aligned} h_{n_3}(A,d)\le \tilde{{\mathcal {H}}}^d_{n_3}(\textrm{Sp}(A)\cap {\mathbb {D}}_l^c) \end{aligned}$$
for all
l. For a fixed
A and
d,
\(h_{n_3}(A,d)\) is non-decreasing in
\(n_3\) and hence converges to a function of
d,
h(
A,
d) (possibly taking infinite values). Furthermore,
$$\begin{aligned} \tilde{{\mathcal {H}}}^{d}(\textrm{Sp}(A)\cap {\mathbb {D}}^c)\le h(A,d) \le \tilde{{\mathcal {H}}}^{d}(\textrm{Sp}(A)\cap {\mathbb {D}}_l^c). \end{aligned}$$
Since the set
\(\textrm{Sp}(A)\cap {\mathbb {D}}\) is countable, its Hausdorff dimension is zero. Using sub-additivity of Hausdorff dimension and Theorem
9.1,
$$\begin{aligned} \textrm{dim}_H(\textrm{Sp}(A))&\le \textrm{dim}_H(\textrm{Sp}(A)\cap {\mathbb {D}}^c)\\&\le \textrm{dim}_H(\overline{\textrm{Sp}(A)\cap {\mathbb {D}}^c})=\textrm{dim}_{H'}(\textrm{Sp}(A)\cap {\mathbb {D}}^c)\\&\le \textrm{dim}_H(\overline{\textrm{Sp}(A)\cap {\mathbb {D}}^c_l})=\textrm{dim}_{H'}(\textrm{Sp}(A)\cap {\mathbb {D}}^c_l)\\&\le \textrm{dim}_H(\textrm{Sp}(A)). \end{aligned}$$
It follows that
\(h(A,d)=0\) if
\(d>\textrm{dim}_H(\textrm{Sp}(A))\) and that
\(h(A,d)=\infty \) if
\(d<\textrm{dim}_H(\textrm{Sp}(A))\). Define
$$\begin{aligned} \Gamma _{n_3,n_2,n_1}(A)=\sup _{j=1,\ldots ,2^{n_3}}\left\{ \frac{j}{2^{n_3}}:h_{n_3,n_2,n_1}(A,k/{2^{n_3}})+\frac{1}{n_2}>\frac{1}{2}\text { for }k=1,\ldots ,j\right\} , \end{aligned}$$
where in this case we define the maximum over the empty set to be 0.
Consider
\(n_2\ge n_3\). Since
\(h_{n_3,n_2,n_1}(A,d)\uparrow h_{n_3,n_2}(A,d)\), it is clear that
$$\begin{aligned} \lim _{n_1\rightarrow \infty }\Gamma _{n_3,n_2,n_1}(A)= & {} \sup _{j=1,\ldots ,2^{n_3}}\left\{ \frac{j}{2^{n_3}}:h_{n_3,n_2}(A,k/2^{n_3})+\frac{1}{n_2}>\frac{1}{2}\text { for }k=1,\ldots ,j\right\} \\=: & {} \Gamma _{n_3,n_2}(A). \end{aligned}$$
If
\(h_{n_3}(A,d)\ge 1/2\), then
\(h_{n_3,n_2}(A,d)+1/{n_2}>1/2\) for all
\(n_2\) otherwise
\(h_{n_3,n_2}(A,d)+1/{n_2}<1/2\) eventually. Hence,
$$\begin{aligned} \lim _{n_2\rightarrow \infty }\Gamma _{n_3,n_2}(A)=\sup _{j=1,\ldots ,2^{n_3}}\left\{ \frac{j}{2^{n_3}}:h_{n_3}(A,k/{2^{n_3}})\ge \frac{1}{2}\text { for }k=1,\ldots ,j\right\} =:\Gamma _{n_3}(A). \end{aligned}$$
Using the monotonicity of
\(h_{n_3}(A,d)\) in
d and the proven properties of the limit function
h, it follows that
$$\begin{aligned} \lim _{n_3\rightarrow \infty }\Gamma _{n_3}(A)=\textrm{dim}_H(\textrm{Sp}(A)). \end{aligned}$$
The fact that
\(h_{n_3}\) is non-decreasing in
\(n_3\), the set
\(\{1/2^{n_3},2/2^{n_3},\ldots ,1\}\) refines itself, and the stated monotonicity collectively shows that convergence is monotonic from below, and hence, we get the
\(\Sigma _3^A\) classification.
Step 4 \(\{\Xi _H,\Omega _\textrm{SA},\Lambda _1\}\in \Sigma _4^A\) and
\(\{\Xi _H,\Omega _\textrm{SA},\Lambda _2\}\in \Sigma _3^A\). The first of these can be proven as in step 3 by replacing
\((n_1,n_2,n_3)\) by
\((n_2,n_3,n_4)\) and the set
\(S_{n_2,n_1}(A)\) by the set
\(S_{n_3,n_2,n_1}(A)\) given by the union of all
\(S\in \rho _{n_3}\) with
\(S\subset [-n_2,n_2]\) and such that the
\(\Sigma _2^A\) tower of algorithms discussed in Lemma
3.21 outputs “Yes” for the interior of
S and input parameters
\((n_2,n_1)\). To prove
\(\{\Xi _H,\Omega _\textrm{SA},\Lambda _2\}\in \Sigma _3^A\), we use exactly the same construction as in step 3 now using the
\(\Sigma _1^A\) algorithm (which uses
\(\Lambda _2\)) given by Lemma
3.21.
\(\square \)