This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Separable Spaces of Continuous Functions as Calkin Algebras

Pavlos Motakis P. Motakis, Department of Mathematics and Statistics, York University, 4700 Keele Street, Toronto, Ontario, M3J 1P3, Canada [email protected]
Abstract.

It is proved that for every compact metric space KK there exists a Banach space XX whose Calkin algebra (X)/𝒦(X)\mathcal{L}(X)/\mathcal{K}(X) is homomorphically isometric to C(K)C(K). This is achieved by appropriately modifying the Bourgain-Delbaen \mathscr{L}_{\infty}-space of Argyros and Haydon in such a manner that sufficiently many diagonal operators on this space are bounded.

2020 Mathematics Subject Classification:
46B07, 46B25, 46B28, 46J10.
The author was supported by NSERC Grant RGPIN-2021-03639.

1. Introduction

For a Banach space XX denote by (X)\mathcal{L}(X) the algebra of bounded linear operators on XX and by 𝒦(X)\mathcal{K}(X) the compact operator ideal in (X)\mathcal{L}(X). A Banach algebra 𝒜\mathcal{A} is said to be a Calkin algebra if there exists an underlying Banach space XX so that the Calkin algebra 𝒞𝒶𝓁(𝒳)=(𝒳)/𝒦(𝒳)\mathpzc{Cal}(X)=\mathcal{L}(X)/\mathcal{K}(X) of XX is isomorphic as a Banach algebra (not necessarily isometrically) to 𝒜\mathcal{A}. The question of what unital Banach algebras are Calkin algebras is very rudimentary. Calkin introduced this object for X=2X=\ell_{2} in 1941 ([11]). Since then 𝒞𝒶𝓁(2)\mathpzc{Cal}(\ell_{2}) has uninterruptedly been in the spotlight, partly owing to the fact that it has highlighted connections between operator algebras and other fields of mathematics, e.g., KK-theory (Brown, Douglas, and Fillmore, [10]), Set Theory (Phillips and Weaver, [39]), and Descriptive Set Theory (Farah, [14]). The systematic study of 𝒞𝒶𝓁(𝒳)\mathpzc{Cal}(X) for general Banach spaces XX dates back to Yood’s 1954 paper [45]. The term Calkin algebra in this precise context can be traced at least as far back as 1974 to Caradus, Pfaffenberger, and Yood’s book [12] who proposed the problem of specifying criteria on XX which would determine whether 𝒞𝒶𝓁(𝒳)\mathpzc{Cal}(X) is semi-simple. The advent of powerful modern construction techniques in Banach spaces, such as the Gowers-Maurey ([17]) and Argyros-Haydon ([4]) methods, made it finally possible to represent certain relatively simple Banach algebras as Calkin algebras. Notable examples include the complex field ([4]), the semigroup algebra of 0\mathbb{N}_{0} (Tarbard, [44]), and C(K)C(K) for a countable compact space KK (Puglisi, Zisimopoulou, and the author, [33]). Although this is an impressive fact, more complicated Banach algebras, such as C[0,1]C[0,1], have been entirely out of reach with past methods. The purpose of the current paper is to develop a new technique that bridges this gap. To overcome existing limitations, a radically new way of imposing a prescribed structure of operators on a Banach space XX is presented. As a result, for every compact metric space KK, C(K)C(K) admits a representation as a Calkin algebra.

Within the context of modern Banach space theory, there is a strong relation between the explicit description of quotient algebras of (X)\mathcal{L}(X) and the tight control of the structure of bounded linear operators on a Banach space XX. The most relevant examples to the present paper are the Gowers-Maurey space XGMX_{\mathrm{GM}} from 1993 ([17]) and the Argyros-Haydon space 𝔛AH\mathfrak{X}_{\mathrm{AH}} from 2011 ([4]). Each of these groundbreaking constructions solved numerous longstanding open problems and revolutionized the view of general Banach spaces. Both spaces have the tightest possible space of operators modulo a small ideal, which is different in each case. An operator between Banach spaces is called strictly singular if it does not preserve an isomorphic copy of any infinite dimensional subspace of its domain. Denote by 𝒮𝒮(X)\mathcal{SS}(X) the ideal of strictly singular operators on XX. Then, every T(XGM)T\in\mathcal{L}(X_{\mathrm{GM}}) is a scalar operator plus a strictly singular operator and every T(𝔛AH)T\in\mathcal{L}(\mathfrak{X}_{\mathrm{AH}}) is a scalar operator plus a compact one. In other words, (XGM)/𝒮𝒮(XGM)\mathcal{L}(X_{\mathrm{GM}})/\mathcal{SS}(X_{\mathrm{GM}}) and (𝔛AH)/𝒦(𝔛AH)\mathcal{L}(\mathfrak{X}_{\mathrm{AH}})/\mathcal{K}(\mathfrak{X}_{\mathrm{AH}}) are both one-dimensional. The space 𝔛AH\mathfrak{X}_{\mathrm{AH}} was constructed by combining the Bourgain-Delbaen method for defining \mathscr{L}_{\infty}-spaces from [8] with the Gowers-Maurey space XGMX_{\mathrm{GM}}. It is now understood that frequently phenomena that can be witnessed “modulo strictly singular operators” in Gowers-Maurey-type spaces can also be witnessed “modulo compact operators” in Argyros-Haydon-type spaces (see, e.g., [44], [29],[6]).

Gowers and Maurey sought in [18] the introduction of a prescribed structure on the space of operators (X)\mathcal{L}(X) of a Banach space XX. Building upon their aforementioned paper [17] they developed a general method for defining Banach spaces whose algebra of operators has a quotient algebra that is generated by what they called a proper family of spread operators. Among other examples, they defined a space XSX_{S} with a basis on which both the left shift LL and the right shift RR are bounded. In fact, any operator T(XS)T\in\mathcal{L}(X_{S}) can be written as T=λI+n=1anRn+n=1bnLn+ST=\lambda I+\sum_{n=1}^{\infty}a_{n}R^{n}+\sum_{n=1}^{\infty}b_{n}L^{n}+S, where S𝒮𝒮(XS)S\in\mathcal{SS}(X_{S}) and the scalar coefficients in both series are absolutely summable. As a consequence, the quotient algebra (XS)/𝒮𝒮(XS)\mathcal{L}(X_{S})/\mathcal{SS}(X_{S}) coincides with the convolution algebra 1()\ell_{1}(\mathbb{Z}) (also known as the Wiener algebra).

Tarbard adapted some of the techniques from [18] and defined an Argyros-Haydon-type \mathscr{L}_{\infty}-space 𝔛T\mathfrak{X}_{\mathrm{T}} in [44] on which a kind of right shift operator RR is bounded. On this space every bounded linear poperator TT can be writen as T=λI+n=1anRn+AT=\lambda I+\sum_{n=1}^{\infty}a_{n}R^{n}+A with A𝒦(𝔛T)A\in\mathcal{K}(\mathfrak{X}_{\mathrm{T}}) and the scalar coefficients in the series being absolutely summable. Thusly, the resulting Calkin algebra of 𝔛T\mathfrak{X}_{\mathrm{T}} is the convolution algebra 1(0)\ell_{1}(\mathbb{N}_{0}). With similar techniques, he had earlier constructed in [43] for each nn\in\mathbb{N} a space whose Calkin algebra is the algebra of n×nn\times n upper triangular Toeplitz matrices.

There is a natural three-step process for defining a space with a prescribed Calkin algebra 𝒜\mathcal{A}. The first step is to identify an appropriate class 𝒞\mathcal{C} of operators, acting on a classical Banach space X0X_{0} with a basis, that generates 𝒜\mathcal{A} in (X0)\mathcal{L}(X_{0}). For example, the left and right shift acting on X0=1()X_{0}=\ell_{1}(\mathbb{Z}) generate the Wiener algebra. The next step is to adapt the methods from [18] to define a Gowers-Maurey-type space XX on which this predetermined class of operators can be used to approximate all TT in (X)\mathcal{L}(X), modulo the strictly singular operators. The final step is to involve the Bourgain-Delbaen construction method to produce an Argyros-Haydon-type space whose Calkin algebra is explicitly 𝒜\mathcal{A}. How well this process works depends on the class 𝒞\mathcal{C} and the space X0X_{0}. Classes of spread operators on 1\ell_{1} have fitted within this framework nicely. Without elaborating too much on the reason of this success, the second step comes down to the fact that XGMX_{\mathrm{GM}}, being based on Schlumprecht space ([42]), has a lot of local 1\ell_{1}-structure and that shift operators don’t interfere too much with conditional structure. For the third step, it is important that bounded shift operators can already be found ([43, Theorem 3.7, page 742]) in a certain “simple” mixed-Tsirelson Bourgain-Delbaen \mathscr{L}_{\infty}-space of Argyros and Haydon ([4, Section 4]). The present paper follows this paradigm but, due to inherent limitations of all previous Bourgain-Delbaen constructions, it cannot be carried out without drastic conceptual modification.

There are also examples of explicit Calkin algebras that do not adhere to this specific three-step process. For example, in [23] it was observed by Kania and Laustsen that by combining finitely many carefully chosen Argyros-Haydon spaces one can obtain any finite dimensional semi-simple complex algebra as a Calkin algebra. An earlier, but more involved, instance of this type of construction is due to Puglisi, Zisimopoulou, and the author from [33]. The main idea is that by taking infinitely many Argyros-Haydon spaces and combining them with an Argyros-Haydon sum (introduced by Zisimopoulou in [46]) the resulting space has Calkin algebra C(ω)C(\omega). Iterating this process yields, for every countable compactum KK, a C(K)C(K) Calkin algebra. A similar method was used by Puglisi, Tolias, and the author in [32] to construct a variety of Calkin algebras, e.g., quasi-reflexive and hereditarily indecomposable ones. All underlying Banach spaces mentioned in this paragraph may be viewed as composite Argyros-Haydon spaces.

Although the statement of the main result of this paper is very similar to the aforementioned one from [33] the proof is very different. As it was pointed out in that paper, the iterative method is insufficient for uncountable KK. Here, the concept of introducing a prescribed structure of operators in a Bourgain-Delbaen space is built from the ground up to achieve the desired result. With regards to the first step of the three-step process towards constructing a C(K)C(K) Calkin algebra, normal operators on 2\ell_{2} are the natural candidates. Indeed, by the spectral theorem the CC^{*}-algebra generated by a normal operator TT is C(σ(T))C(\sigma(T)). This is in fact fairly straightforward whenever TT is diagonal with respect to some orthonormal basis and thus its norm is the supremum of its diagonal entries. Going through this exercise clarifies that on any space with an unconditional basis any family of diagonal operators generates a C(K)C(K) Banach algebra. This supports the conclusion that the class 𝒞\mathcal{C} in the first step needs to be one consisting of sufficiently many diagonal operators that capture the information of the compact space KK.

Theorem A.

Let KK be a compact metric space. There exists an Argyros-Haydon-type Bourgain-Delbaen \mathscr{L}_{\infty}-space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with a conditional Schauder basis (dγ)γΓ(d_{\gamma})_{\gamma\in\Gamma} that satisfies the following properties.

  1. (a)

    Every bounded linear operator T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} can be written in the form T=D+AT=D+A where DD is diagonal bounded linear operator and AA is a compact linear operator.

  2. (b)

    There exists a Banach algebra isomorphism Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}).

  3. (c)

    The space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} admits an equivalent norm with respect to which Ψ\Psi is an isometry. In particular, this space’s Calkin algebra is homomorphically isometric to C(K)C(K).

Statement (c) is separate to indicate the fact that the usual Bourgain-Delbaen norm does not have the isometric property.

As it has been mentioned repeatedly, the construction of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} cannot be an immediate application of the aforementioned three-step process. Although the first and second step would go through with reasonably standard modifications of classical methods, the third one meets a dead end; on all types of previously defined Bourgain-Delbaen spaces non-trivial diagonal operators are never bounded. To overcome this, it is necessary to model the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} on a different Gowers-Maurey-type space that relies heavily on a technique called saturation under constraints initiated by Odell and Schlumprecht ([34] and [35]) and extensively developed by Argyros, the author, and others ([5], [6], etc.). In the end, boundedness of non-scalar operators is achieved in a fundamentally different way compared to previous non-classical spaces (e.g., [15], [18]). Significant effort has been made to explain this new idea. To this end, Section 2 is entirely devoted to an exposition of concepts in a less intimidating mixed-Tsirelson stage. None of the results from that section are used directly in the proof of Theorem A. Instead, everything needs to be reframed in the setting of the Argyros-Haydon construction. However, the main point is that the final result is based on an accessible novel principle that is then thrusted by powerful existing technologies to fulfill its potential.

The paper is organized into nine sections. Arguably, the most important one is Section 2 in which the underlying principles behind Theorem A are clarified. This is done by defining and briefly studying a mixed-Tsirelson space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. Section 3 deals with the development of the Argyros-Haydon-type Bourgain-Delbaen incarnation 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} of the simpler space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. In Section 4 it is shown that sufficiently many diagonal operators on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} are bounded to define a homomorphic embedding Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}). In Section 5 the main Theorem is proved by showing that Ψ\Psi is onto. This argument is made modulo two black-box theorems that rely on the conditional structure of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, which is studied in Sections 6 and 7. In these two sections, techniques from the theory of hereditarily indecomposable (HI) spaces are implemented and a satisfactory first reading of the paper is possible by omitting them. Some fundamental and non-trivial parts of the Argyros-Haydon construction, such as estimations on rapidly increasing sequences, carry over verbatim to the current paper. To avoid inflating the contents of the paper, they have not been repeated. Section 8 outlines some additional structural properties of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. The final section is devoted to a detailed discussion of possible future directions in this line of research.

All vector spaces are over the complex field. This is to mirror standard practice in the study of operator algebras and it is not essential; all definitions and proofs work just as well over the real field. Denote by \mathbb{C}_{\mathbb{Q}} the field of complex numbers with rational real and imaginary parts and put 𝔻={λ:|λ|1}\mathbb{D}_{\mathbb{Q}}=\{\lambda\in\mathbb{C}_{\mathbb{Q}}:|\lambda|\leq 1\}. Let c00c_{00} denote the vector space of eventually zero complex sequences. For f=(bn)nf=(b_{n})_{n}, x=(an)nx=(a_{n})_{n} in c00c_{00} define f(x)=nbnanf(x)=\sum_{n}b_{n}a_{n}. The unit vector basis of c00c_{00} is denoted both by (en)n(e_{n}^{*})_{n} and by (en)n(e_{n})_{n} depending on whether it is seen as a sequence of functionals or one of vectors. For xc00x\in c_{00} denote supp(x)={n:en(x)0}\operatorname{supp}(x)=\{n\in\mathbb{N}:e_{n}^{*}(x)\neq 0\} and for EE\subset\mathbb{N} let Ex=nEen(x)enEx=\sum_{n\in E}e_{n}^{*}(x)e_{n}. For xx, yy in c00c_{00} write x<yx<y to mean that their supports are successive subsets of \mathbb{N}. A sequence of successive vectors in c00c_{00} is called a block sequence.

Let, for the entirety of this paper, (K,ρ)(K,\rho) be a fixed compact metric space.

2. Heuristic explanation on a mixed-Tsirelson space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

Schlumprecht space from [42] is one of the important evolutionary steps in the history of non-classical Banach spaces. It was an integral ingredient in the solution to the unconditional sequence problem by Gowers and Maurey who constructed in [17] the first hereditarily indecomposable (HI) space XGMX_{GM}. On XGMX_{GM}, every bounded linear operator is of the form T=λI+ST=\lambda I+S, with SS strictly singular. Mixed-Tsirelson spaces can be viewed as a discretization of Schlumprecht space. The first such space was defined by Argyros and Deliyanni in [1] and this section introduces a new variant.

With a novel approach, a Gowers-Maurey-like reflexive mixed-Tsirelson space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is defined on which a large class of diagonal operators are bounded. More precisely, this space has a conditional Schauder basis and the following properties.

  1. (a)

    Every bounded linear operator TT on XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} can be written as T=D+ST=D+S with DD diagonal and SS strictly singular.

  2. (b)

    The quotient algebra (XC(K))/𝒮𝒮(XC(K))\mathcal{L}(X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})/\mathcal{SS}(X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is isomorphic, as a Banach algebra, to C(K)C(K).

It is important to point out that not all bounded scalar sequences define bounded diagonal operators on XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, otherwise its basis would be unconditional. Modulo the strictly singular operators, the bounded diagonal operators describe the space C(K)C(K).

There already exists an approach that has been used to introduce a prescribed structure of operators in Gowers-Maurey-type spaces (see, e.g., [15] and [18]). In this method a space is defined by constructing a norming set, i.e., a subset WW of the ball BXB_{X^{*}} of the dual that defines the norm of the resulting space XX. To enforce the boundedness of a desired collection of operators, the standard norming set WGMW_{\mathrm{GM}} of XGMX_{\mathrm{GM}} is augmented to a larger set WW by making it stable under the action of TT^{*} (i.e., T(W)WT^{*}(W)\subset W), for TT in an appropriate class 𝒞\mathcal{C}. The present approach is entirely different. It is based on a saturation under constraints from [7] and instead of enriching the set WGMW_{\mathrm{GM}}, very strict conditions are imposed on what members are allowed to be used from it. The operators that will eventually be bounded never appear explicitly in the construction and the imposed constraints relate to weights of functions and the metric of the compact space under consideration. Rather unexpectedly, this impoverishment of the norming set results in the enrichment of the space of operators. This somewhat bizarre phenomenon is being discussed here with the sole purpose of introducing the ideas behind the Bourgain-Delbaen spaces that appear in the main result of the paper. The properties of XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} are not justified in full detail and the only proof presented here is that certain diagonal operators are bounded.

At this point it needs to be pointed out that there are more conventional paths that yield a space satisfying properties (a) and (b). These paths however do not translate well to the Bourgain-Delbaen setting. Also, in reality the current construction does not use the set WGMW_{\mathrm{GM}} as a starting point but a similar standard set WmTHIW_{\mathrm{mT}}^{\mathrm{HI}} introduced below.

2.1. Mixed-Tsirelson spaces

Fix, for the remainder of this paper, a double sequence of positive even integers (mj,nj)j(m_{j},n_{j})_{j} that satisfies the conditions

  1. (a)

    m18m_{1}\geq 8,

  2. (b)

    mj+1mj2m_{j+1}\geq m_{j}^{2},

  3. (c)

    n1m12n_{1}\geq m_{1}^{2}, and

  4. (d)

    nj+1(16nj)log2mj+1n_{j+1}\geq(16n_{j})^{\log_{2}m_{j+1}}.

Definition 2.1.

Let WW be a subset of the unit ball of (c00,)(c_{00},\|\cdot\|_{\infty}).

  1. (i)

    The set WW is called a norming set if it contains the unit vector basis and for every fWf\in W, λ𝔻\lambda\in\mathbb{D}_{\mathbb{Q}}, and interval EE of \mathbb{N}, λEfW\lambda Ef\in W.

  2. (ii)

    For jj\in\mathbb{N}, WW is said to be closed under the (mj1,𝒜nj)(m_{j}^{-1},\mathcal{A}_{n_{j}})-operation, if for every 1dnj1\leq d\leq n_{j} and f1<<fdWf_{1}<\cdots<f_{d}\in W,

    f=1mji=1dfiW.f=\frac{1}{m_{j}}\sum_{i=1}^{d}f_{i}\in W.

    Such an ff is said to be the outcome of an (mj1,𝒜nj)(m_{j}^{-1},\mathcal{A}_{n_{j}})-operation and it is called a weighted functional with weight(f)=mj1\operatorname{weight}(f)=m_{j}^{-1}. Note that this notion is dependent on WW and it is not necessarily uniquely determined. As a convention it will be assumed that for all nn\in\mathbb{N} and λ𝔻\lambda\in\mathbb{D}_{\mathbb{Q}}, weight(λen)=0\operatorname{weight}(\lambda e^{*}_{n})=0. This is important.

  3. (iii)

    For jj\in\mathbb{N} and a family \mathcal{F} of finite sequences of successive members of WW, the set WW is said to be closed under the (mj1,𝒜nj,)(m_{j}^{-1},\mathcal{A}_{n_{j}},\mathcal{F})-operation if for every 1dnj1\leq d\leq n_{j} and (f1,,fd)(f_{1},\ldots,f_{d}) in \mathcal{F},

    f=1mji=1dfiW.f=\frac{1}{m_{j}}\sum_{i=1}^{d}f_{i}\in W.

    Such an ff is said to be the outcome of an (mj1,𝒜nj,)(m_{j}^{-1},\mathcal{A}_{n_{j}},\mathcal{F})-operation.

Definition 2.2.

The fundamental mixed-Tsirelson norming set WmTW_{\mathrm{mT}} is the smallest subset of c00c_{00} that satisfies the following properties.

  1. (i)

    The set WmTW_{\mathrm{mT}} is a norming set.

  2. (ii)

    For every positive integer jj\in\mathbb{N} the set WmTW_{\mathrm{mT}} is closed under the (mj1,𝒜nj)(m_{j}^{-1},\mathcal{A}_{n_{j}})-operation.

The space XmTX_{\mathrm{mT}} induced by this norming set is the completion of c00c_{00} under the norm x=sup{|f(x)|:fWmT}\|x\|=\sup\{|f(x)|:f\in W_{\mathrm{mT}}\}. The experienced reader may have spotted that WmTW_{\mathrm{mT}} is not closed under rational convex combinations. This is intentional and it is a necessary prerequisite for the application of saturation under constraints with increasing weights in the style of [7]. This ingredient will be included in the recipe of the space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. The omission of convex combinations is a constant feature in this paper.

When it comes to HI-type constructions, the family of special sequences resulting from a Maurey-Rosenthal coding function ([30]) is a ubiquitous tool. Let 𝒬\mathcal{Q} denote the collection of all finite sequences of successive non-zero members of the unit ball of (c00,)(c_{00},\|\cdot\|_{\infty}) that have coefficients in 𝔻\mathbb{D}_{\mathbb{Q}}. Fix an injection σ:𝒬\sigma:\mathcal{Q}\to\mathbb{N} so that for every (f1,,fd)𝒬(f_{1},\ldots,f_{d})\in\mathcal{Q}, σ(f1,,fd)>fd12maxsupp(fd)\sigma(f_{1},\ldots,f_{d})>\|f_{d}\|_{\infty}^{-1}2^{\max\operatorname{supp}(f_{d})}.

Definition 2.3.

Let WW be a norming set. A sequence (f1,,fd)𝒬(f_{1},\ldots,f_{d})\in\mathcal{Q} of weighted functionals is WW is called a special sequence if the following hold:

  1. (i)

    for some j1j_{1}\in\mathbb{N}, weight(f1)=m4j121\operatorname{weight}(f_{1})=m_{4j_{1}-2}^{-1}, and

  2. (ii)

    for 1<id1<i\leq d weight(fi)=m4σ(f1,,fi1)1\operatorname{weight}(f_{i})=m_{4\sigma(f_{1},\ldots,f_{i-1})}^{-1}.

Denote the collection of all special sequences in WW by sp(W)\mathcal{F}_{\mathrm{sp}}(W).

The important feature of a special sequence is that the weight of the last member uniquely determines the sequence of its predecessors.

Definition 2.4.

The fundamental mixed-Tsirelson HI norming set WmThiW_{\mathrm{mT}}^{\mathrm{hi}} is the smallest subset of c00c_{00} that satisfies the following properties.

  1. (i)

    The set WmThiW_{\mathrm{mT}}^{\mathrm{hi}} is a norming set.

  2. (ii)

    For every even positive integer 2j2j\in\mathbb{N} the set WmThiW_{\mathrm{mT}}^{\mathrm{hi}} is closed under the (m2j1,𝒜n2j)(m_{2j}^{-1},\mathcal{A}_{n_{2j}})-operation.

  3. (iii)

    For every odd positive integer 2j12j-1\in\mathbb{N} the set WmThiW_{\mathrm{mT}}^{\mathrm{hi}} is closed under the (m2j11,𝒜n2j1,sp(WmThi))(m_{2j-1}^{-1},\mathcal{A}_{n_{2j-1}},\mathcal{F}_{\mathrm{sp}}(W_{\mathrm{mT}}^{\mathrm{hi}}))-operation.

The space XmThiX_{\mathrm{mT}}^{\mathrm{hi}} induced by this norming set is an HI space on which every bounded linear operator TT can be written as T=λI+ST=\lambda I+S with SS strictly singular.

2.2. Definition of XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

In [15] and [18] the space of operators is augmented by loosening the above condition (iii) and thus creating a larger norming set. Instead, in the current paper conditions (ii) and (iii) are tightened even further to result in a smaller norming set. As it turns out, this enriches the space of operators. The following constraint was first considered in [7]. It has its roots in papers [34] and [35] of Odell and Schlumprecht and it was further developed in a series of several papers by Argyros, the author, and others ([5], [6], etc.).

Definition 2.5.

Let WW be a norming set. A sequence of weighted functionals f1<f2<<fdf_{1}<f_{2}<\cdots<f_{d} in WW is said to be very fast growing if, for 1<id1<i\leq d, weight(fi)<2maxsupp(fi1)\operatorname{weight}(f_{i})<2^{-\max\operatorname{supp}(f_{i-1})}. Denote the collection of all very fast growing sequences in WW by vfg(W)\mathcal{F}_{\mathrm{vfg}}(W).

Note that any sequence of basis elements is very fast growing and that sp(W)vfg(W)\mathcal{F}_{\mathrm{sp}}(W)\subset\mathcal{F}_{\mathrm{vfg}}(W). The family vfg(W)\mathcal{F}_{\mathrm{vfg}}(W) yields a constraint by applying it to Definition 2.1 (iii). This type of constraint has been used to study the local and asymptotic structure of Banach spaces. Such structure has strong implications to spaces of operators, e.g., in [5] the first reflexive space with the invariant subspace property was constructed.

As it was already mentioned, an additional constraint needs to be introduced which comes from the metric space (K,ρ)(K,\rho). For the remainder of this section fix a sequence (κi)i=1(\kappa_{i})_{i=1}^{\infty} in KK so that for all nn\in\mathbb{N}, {κi:in}\{\kappa_{i}:i\geq n\} is dense in KK. Next, for a given norming set WW, one associates to some fWf\in W an element κ(f)\kappa(f) of KK.

Definition 2.6.

Let WW be a norming set. For every nn\in\mathbb{N} and λ𝔻\lambda\in\mathbb{D}_{\mathbb{Q}} define κ(λen)=κn\kappa(\lambda e^{*}_{n})=\kappa_{n}.

  1. (a)

    A sequence f1<<fdf_{1}<\cdots<f_{d} in vfg(W)\mathcal{F}_{\mathrm{vfg}}(W) is said to have essentially rapidly converging supports if, for 1id1\leq i\leq d, κ(fi)\kappa(f_{i}) is defined and there exists κ0K\kappa_{0}\in K such that for 1<id1<i\leq d, ρ(κ(fi),κ0)2maxsupp(fi1)\rho(\kappa(f_{i}),\kappa_{0})\leq 2^{-\max\operatorname{supp}(f_{i-1})}. Denote the collection of all such sequences in WW by ercs(W)\mathcal{F}_{\mathrm{ercs}}(W).

  2. (b)

    For an ff in WW that is the outcome of an (mj1,𝒜nj,ercs(W))(m_{j}^{-1},\mathcal{A}_{n_{j}},\mathcal{F}_{\mathrm{ercs}}(W))-operation applied to a sequence (fi)i=1d(f_{i})_{i=1}^{d} as above, define κ(f)=κ0\kappa(f)=\kappa_{0}.

Note that a sequence with essentially rapidly converging supports is always assumed to be very fast growing. This has notational advantages but it would also have been fine to disentangle the two notions. Similar to the weight function, the associated element κ(f)\kappa(f) is not unique and κ\kappa is a partially defined multi valued function from WW to KK. Also note that the definition of κ\kappa is implicit. This is formally sound and κ\kappa is, at the very least, defined on the basis elements. So if a norming set WW is closed under infinitely many (mj1,𝒜nj,ercs(W))(m_{j}^{-1},\mathcal{A}_{n_{j}},\mathcal{F}_{\mathrm{ercs}}(W))-operations then the compactness of KK yields a wealth of functionals ff for which κ(f)\kappa(f) is defined. Indeed, any very fast growing sequence (fn)n(f_{n})_{n}, for which all κ(fn)\kappa(f_{n}) are defined, has a subsequence with essentially rapidly converging supports on which the operations can be applied.

For a norming set WW denote ercssp(W)=ercs(W)sp(W)\mathcal{F}_{\mathrm{ercs}}^{\mathrm{sp}}(W)=\mathcal{F}_{\mathrm{ercs}}(W)\cap\mathcal{F}_{\mathrm{sp}}(W), i.e., the collection of special sequences with essentially rapidly converging supports. The time is ripe to define the norming set WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} of the space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

Definition 2.7.

Define WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} to be the smallest subset of c00c_{00} that satisfies the following properties.

  1. (i)

    The set WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is a norming set.

  2. (ii)

    For every even positive integer 2j2j\in\mathbb{N} the set WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is closed under the (m2j1,𝒜n2j,ercs(WC(K)))(m_{2j}^{-1},\mathcal{A}_{n_{2j}},\mathcal{F}_{\mathrm{ercs}}(W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}))-operation.

  3. (iii)

    For every odd positive integer 2j12j-1\in\mathbb{N} the set WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is closed under the (m2j11,𝒜n2j1,ercssp(WC(K)))(m_{2j-1}^{-1},\mathcal{A}_{n_{2j-1}},\mathcal{F}^{\mathrm{sp}}_{\mathrm{ercs}}(W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}))-operation.

It is useful to observe that the set WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} can be defined as an increasing union of sets WmW_{m}, m=0,1,m=0,1,\ldots where W0={λen:n and λ𝔻}W_{0}=\{\lambda e_{n}:n\in\mathbb{N}\text{ and }\lambda\in\mathbb{D}_{\mathbb{Q}}\} and if WmW_{m} has been defined then Wm+1W_{m+1} is the union of WmW_{m} with the collection of all λEf\lambda Ef, where λ𝔻\lambda\in\mathbb{D}_{\mathbb{Q}}, EE is an interval of \mathbb{N}, and ff is the outcome of an (m2j1,𝒜n2j,ercs(Wm))(m_{2j}^{-1},\mathcal{A}_{n_{2j}},\mathcal{F}_{\mathrm{ercs}}(W_{m}))-operation or an (m2j1,𝒜n2j,ercssp(Wm))(m_{2j}^{-1},\mathcal{A}_{n_{2j}},\mathcal{F}^{\mathrm{sp}}_{\mathrm{ercs}}(W_{m}))-operation. This in particular implies that for every fWC(K)f\in W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, κ(f)\kappa(f) is defined and that for every λ𝔻\lambda\in\mathbb{D}_{\mathbb{Q}} and interval EE of \mathbb{N} such that λEf0\lambda Ef\neq 0, κ(λEf)=κ(f)\kappa(\lambda Ef)=\kappa(f).

It is almost shocking that this norming set induces a space XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} that is not HI. Unless KK is a singleton, XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} contains decomposable subspaces (see Proposition 8.4 (ii)). For a continuous function ϕ:K\phi:K\to\mathbb{C} denote by ϕ^:c00c00\hat{\phi}:c_{00}\to c_{00} the linear operator with ϕ^(en)=ϕ(κn)en\hat{\phi}(e_{n})=\phi(\kappa_{n})e_{n}. Note that ϕ^\hat{\phi}, being a diagonal operator, is formally dual to itself, i.e., for every f,xc00f,x\in c_{00}, f(ϕ^x)=(ϕ^f)(x)f(\hat{\phi}x)=(\hat{\phi}f)(x).

Proposition 2.8.

Let ϕ:K\phi:K\to\mathbb{C} be a Lipschitz function. Then ϕ^\hat{\phi} extends to a bounded linear operator on XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

With regards, to the proof of Proposition 2.8, the fact that ϕ\phi is Lipschitz is not particularly important and by tweaking the metric (before defining XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) any continuous function may be assumed Lipschitz. The trueness of this result stems from the fact that every ff in WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} has a perturbation f~\tilde{f} so that the set {κn:nsupp(f~)}\{\kappa_{n}:n\in\operatorname{supp}(\tilde{f})\} has small diameter in KK. Therefore ϕ\phi is almost constant on this subset and thus ϕ^f\hat{\phi}f is close to a scalar multiple of ff. Crucially, there is no unique scalar that works for all ff. The following statement makes this more precise while simultaneously yielding Proposition 2.8.

Proposition 2.9.

Let ϕ:K\phi:K\to\mathbb{C} be a Lipschitz function. Then, there exists NN\in\mathbb{N} so that for every fWC(K)f\in W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with minsupp(f)N\min\operatorname{supp}(f)\geq N,

(1) ϕ^fϕ(κ(f))f3weight(f)ϕ.\Big{\|}\hat{\phi}f-\phi\big{(}\kappa(f)\big{)}f\Big{\|}\leq 3\operatorname{weight}(f)\|\phi\|_{\infty}.

Therefore, on XC(K)X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, ϕ^P[1,N)ϕ^2ϕ\|\hat{\phi}-P_{[1,N)}\hat{\phi}\|\leq 2\|\phi\|_{\infty}.

With this result at hand, and the Stone-Weierstrass theorem, it is not hard to see that C(K)C(K) embeds isomorphically, as a Banach algebra, into (XC(K))/𝒮𝒮(XC(K))\mathcal{L}(X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})/\mathcal{SS}(X_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) (here, it is necessary to use that for all nn\in\mathbb{N}, {κi:in}\{\kappa_{i}:i\geq n\} is dense in KK). Although Proposition 2.9 will be proved now, it will not be shown here that the aforementioned embedding is onto. The interested reader may be able to extrapolate this information from the Bourgain-Delbaen part of this paper.

Proof of Proposition 2.9.

Assume, without loss of generality, that ϕ=1\|\phi\|_{\infty}=1 and denote by LL the Lipschitz constant of ϕ\phi. Pick NN\in\mathbb{N} sufficiently large so that (3+L)/2N15/8(3+L)/2^{N-1}\leq 5/8. Statement (1) is proved by induction on m=0,1,m=0,1,\ldots for all fWmf\in W_{m} with minsupp(f)N\min\operatorname{supp}(f)\geq N. For m=0m=0 and f=λenW0f=\lambda e_{n}\in W_{0}, ϕ^f=ϕ(κn)en=ϕ(κ(f))f\hat{\phi}f=\phi(\kappa_{n})e_{n}=\phi(\kappa(f))f and thus the conclusion holds.

Assume next that (1) is true for all gWmg\in W_{m} with minsupp(g)N\min\operatorname{supp}(g)\geq N and let fWm+1f\in W_{m+1} with minsupp(f)N\min\operatorname{supp}(f)\geq N. Then, there exist jj\in\mathbb{N}, dnjd\leq n_{j}, and a sequence (fi)i=1d(f_{i})_{i=1}^{d} in WmW_{m} with essentially rapidly converging supports in [N,)[N,\infty) such that f=mj1i=1dfif=m_{j}^{-1}\sum_{i=1}^{d}f_{i}. Then,

ϕ^fϕ(κ(f))f\displaystyle\Big{\|}\hat{\phi}f-\phi(\kappa(f))f\Big{\|}=1mji=1d(ϕ^(fi)ϕ(κ(f)))fi\displaystyle=\frac{1}{m_{j}}\Big{\|}\sum_{i=1}^{d}\big{(}\hat{\phi}(f_{i})-\phi(\kappa(f))\big{)}f_{i}\Big{\|}
\displaystyle\leq 1mj(i=1d(ϕ^(fi)ϕ(κ(fi)))fi+i=1d|ϕ(κ(f))ϕ(κ(fi))|fi)\displaystyle\frac{1}{m_{j}}\Big{(}\sum_{i=1}^{d}\Big{\|}\big{(}\hat{\phi}(f_{i})-\phi(\kappa(f_{i}))\big{)}f_{i}\Big{\|}+\sum_{i=1}^{d}\Big{|}\phi(\kappa(f))-\phi(\kappa(f_{i}))\Big{|}\|f_{i}\|\Big{)}
\displaystyle\leq 1mj(3weight(f1)+i=2d3weight(fi)+2+i=2dLρ(κ(f),κ(fi)))\displaystyle\frac{1}{m_{j}}\Big{(}3\operatorname{weight}(f_{1})+\sum_{i=2}^{d}3\operatorname{weight}(f_{i})+2+\sum_{i=2}^{d}L\rho(\kappa(f),\kappa(f_{i}))\Big{)}
\displaystyle\leq 1mj(198+3i=2d2maxsupp(fi1)+Li=2d2maxsupp(fi1))\displaystyle\frac{1}{m_{j}}\Big{(}\frac{19}{8}+3\sum_{i=2}^{d}2^{-\max\operatorname{supp}(f_{i-1})}+L\sum_{i=2}^{d}2^{\max\operatorname{supp}(f_{i-1})}\Big{)}
\displaystyle\leq 1mj(19/8+5/8)=3weight(f).\displaystyle\frac{1}{m_{j}}\big{(}19/8+5/8\big{)}=3\operatorname{weight}(f).

The Bourgain-Delbaen construction that is about to follow is based on the same principles. In order to achieve an isometric result, some components of the definition are chosen more with more precision.

3. The Bourgain-Delbaen \mathscr{L}_{\infty}-space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

A separable Banach space XX is a ,C\mathscr{L}_{\infty,C}-space, where C1C\geq 1, if there exists an increasing sequence (Fn)n(F_{n})_{n} of finite dimensional subspaces of XX, the union of which is dense in XX and so that for all nn\in\mathbb{N}, FnF_{n} is CC-isomorphic to dim(Fn)\ell_{\infty}^{\mathrm{dim}(F_{n})}. Suppressing the constant CC, XX is called a \mathscr{L}_{\infty}-space. The class of p\mathscr{L}_{p}-spaces was introduced by Lindenstrauss and Pełczyński in [27]. Bourgain and Delbaen introduced in [8] a method for constructing non-classical separable \mathscr{L}_{\infty}-spaces. It is one of the essential components in the solution of the scalar-plus-compact problem by Argyros and Haydon in [4] (the other being a mixed-Tsirelson implementation of the hereditarily indecomposable Gowers-Maurey space). The purpose of the first part of this section is to recall a very general Bourgain-Delbaen scheme that is based on [3], where it was proved that every separable \mathscr{L}_{\infty}-space is isomorphic to a Bourgain-Delbaen space. Following this introduction (and following in the footsteps of [4]), a Bourgain-Delbaen space modeled after the space XmTX_{\mathrm{mT}} is introduced. Finally, the extra ingredients discussed in Section 2 are adjusted to this setting to define the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

3.1. General Bourgain-Delbaen \mathscr{L}_{\infty}-spaces

At the most basic level the idea behind the Bourgain-Delbaen construction method is very elegant. For two non-empty sets Γ1Γ\Gamma_{1}\subset\Gamma denote by rΓ1:(Γ)(Γ1)r_{{}_{\Gamma_{1}}}:\ell_{\infty}(\Gamma)\to\ell_{\infty}(\Gamma_{1}) the usual restriction operator. Any linear right inverse i:(Γ1)(Γ)i:\ell_{\infty}(\Gamma_{1})\to\ell_{\infty}(\Gamma) of rΓ1r_{{}_{\Gamma_{1}}} will be called an extension operator, i.e., ii is a linear operator so that for all x(Γ1)x\in\ell_{\infty}(\Gamma_{1}) and γΓ1\gamma\in\Gamma_{1}, i(x)(γ)=x(γ)i(x)(\gamma)=x(\gamma).

The Bourgain-Delbaen scheme is an infinite inductive process in which one defines finite sets Γ1Γ2Γ3\Gamma_{1}\subset\Gamma_{2}\subset\Gamma_{3}\subset\cdots and extension operators i1,2:(Γ1)(Γ2)i_{1,2}:\ell_{\infty}(\Gamma_{1})\to\ell_{\infty}(\Gamma_{2}), i2,3:(Γ2)(Γ3)i_{2,3}:\ell_{\infty}(\Gamma_{2})\to\ell_{\infty}(\Gamma_{3}),…. In the base step, one picks Γ1\Gamma_{1} and no extension operator. Having defined Γ1,,Γn\Gamma_{1},\ldots,\Gamma_{n} and i1,2,,in1,ni_{1,2},\ldots,i_{n-1,n}, denote Δ1=Γ1,Δ2=Γ2Γ1,,Δn=ΓnΓn1\Delta_{1}=\Gamma_{1},\Delta_{2}=\Gamma_{2}\setminus\Gamma_{1},\ldots,\Delta_{n}=\Gamma_{n}\setminus\Gamma_{n-1}. Note that Δ1,,Δn\Delta_{1},\ldots,\Delta_{n} are pairwise disjoint finite sets and Γi=j=1iΔj\Gamma_{i}=\cup_{j=1}^{i}\Delta_{j}, for 1in1\leq i\leq n. To perform the (n+1)(n+1)’th step, one chooses a finite set Δn+1\Delta_{n+1}, that is disjoint from all previously defined ones, and an extension operator in,n+1:(Γn)(Γn+1)i_{n,n+1}:\ell_{\infty}(\Gamma_{n})\to\ell_{\infty}(\Gamma_{n+1}), where Γn+1=ΓnΔn+1\Gamma_{n+1}=\Gamma_{n}\cup\Delta_{n+1}. Although this is the basic essence of the scheme, the inductive choice needs to be performed in a very special manner (described later) to achieve something of interest.

Note that for every m<nm<n this method yields an extension operator im,n=in1,nim+1,m+2im,m+1:(Γm)(Γn)i_{m,n}=i_{n-1,n}\circ\cdots\circ i_{m+1,m+2}\circ i_{m,m+1}:\ell_{\infty}(\Gamma_{m})\to\ell_{\infty}(\Gamma_{n}). Also, denote in,n=id:(Γn)(Γn)i_{n,n}=id:\ell_{\infty}(\Gamma_{n})\to\ell_{\infty}(\Gamma_{n}). The condition under which this construction yields a \mathscr{L}_{\infty}-space is the following.

Assumption 3.1.

There exists C1C\geq 1 so that for every mnm\leq n, im,nC\|i_{m,n}\|\leq C.

If this has been achieved, putting Γ=q=1Γq\Gamma=\cup_{q=1}^{\infty}\Gamma_{q}, for each mm\in\mathbb{N} define the extension operator im=limnim,n:(Γm)(Γ)i_{m}=\lim_{n}i_{m,n}:\ell_{\infty}(\Gamma_{m})\to\ell_{\infty}(\Gamma). Then, for all nn\in\mathbb{N}, inC\|i_{n}\|\leq C and, because ini_{n} is a right inverse of rΓnr_{{}_{\Gamma_{n}}}, the space Yn=in((Γn))Y_{n}=i_{n}(\ell_{\infty}(\Gamma_{n})) is CC-isomorphic to (Γn)\ell_{\infty}(\Gamma_{n}). It also follows that Y1Y2Y_{1}\subset Y_{2}\subset\cdots and therefore the space 𝔛(Γn,in)=n=1Yn¯(Γ)\mathfrak{X}_{(\Gamma_{n},i_{n})}=\overline{\cup_{n=1}^{\infty}Y_{n}}\subset\ell_{\infty}(\Gamma) is a ,C\mathscr{L}_{\infty,C}-space. Any space resulting from such a process is called a Bourgain-Delbaen \mathscr{L}_{\infty}-space.

On such a space 𝔛(Γn,in)\mathfrak{X}_{(\Gamma_{n},i_{n})}, the extension operators are used to define a finite dimensional decomposition (FDD). For each nn\in\mathbb{N} the map Pn=inrΓn:𝔛(Γn,in)YnP_{n}=i_{n}r_{{}_{\Gamma_{n}}}:\mathfrak{X}_{(\Gamma_{n},i_{n})}\to Y_{n} is a projection of norm at most CC. In fact PnPm=PnmP_{n}P_{m}=P_{n\wedge m} and thus the sequence of spaces Z1=P1(𝔛(Γn,in))=i1((Δ1))Z_{1}=P_{1}(\mathfrak{X}_{(\Gamma_{n},i_{n})})=i_{1}(\ell_{\infty}(\Delta_{1})), Zn=(PnPn1)(𝔛(Γn,in))=in((Δn))Z_{n}=(P_{n}-P_{n-1})(\mathfrak{X}_{(\Gamma_{n},i_{n})})=i_{n}(\ell_{\infty}(\Delta_{n})), n2n\geq 2, forms a FDD of 𝔛(Γn,in)\mathfrak{X}_{(\Gamma_{n},i_{n})}. Denote, for all m<nm<n\in\mathbb{N}, P(m,n]=PnPmP_{(m,n]}=P_{n}-P_{m} the associated projection onto the space Zm+1ZnZ_{m+1}\oplus\cdots\oplus Z_{n} by. For any interval II of \mathbb{N} define PIP_{I} analogously. It is also true that if, for all nn\in\mathbb{N} and γΔn\gamma\in\Delta_{n}, one defines dγ=in(eγ)d_{\gamma}=i_{n}(e_{\gamma}), then the sequence ((dγ)γΔn)n=1((d_{\gamma})_{\gamma\in\Delta_{n}})_{n=1}^{\infty} forms a Schauder basis of 𝔛(Γn,in)\mathfrak{X}_{(\Gamma_{n},i_{n})} (see [3, Remark 2.10, page 688]).

While carrying out the Bourgain-Delbaen construction, of particular importance are specific versions of the above projections that can be defined during the steps of the induction. For every mnm\leq n\in\mathbb{N}, and once the nn’th step is complete, put Pm(n)=im,nrΓm:(Γn)Ym(n)=im,n((Γm))P_{m}^{(n)}=i_{m,n}r_{{}_{\Gamma_{m}}}:\ell_{\infty}(\Gamma_{n})\to Y_{m}^{(n)}=i_{m,n}(\ell_{\infty}(\Gamma_{m})). For an interval II of {1,,n}\{1,\ldots,n\} define PI(n)P^{(n)}_{I} analogously.

3.2. Bourgain-Delbaen extension functionals

In the (n+1)(n+1)’th inductive step and having chosen Γ1,,Γn\Gamma_{1},\ldots,\Gamma_{n} and i1,2,,in1,ni_{1,2},\ldots,i_{n-1,n} one must define the set Δn+1\Delta_{n+1} and an extension operator in,n+1:(Γn)(Γn+1)i_{n,n+1}:\ell_{\infty}(\Gamma_{n})\to\ell_{\infty}(\Gamma_{n+1}). Presupposing that the index set Δn+1\Delta_{n+1} has been determined, defining in,n+1i_{n,n+1} is equivalent to finding linear functionals cγ:(Γn)c_{\gamma}^{*}:\ell_{\infty}(\Gamma_{n})\to\mathbb{C}, γΔn+1\gamma\in\Delta_{n+1}, so that for all x(Γn)x\in\ell_{\infty}(\Gamma_{n}) and γΔn+1\gamma\in\Delta_{n+1}, in,n+1(x)(γ)=cγ(x)i_{n,n+1}(x)(\gamma)=c_{\gamma}^{*}(x). Thus, one may shift their focus on defining (cγ)γΔn+1(c_{\gamma}^{*})_{\gamma\in\Delta_{n+1}} instead of in,n+1i_{n,n+1} directly. For obvious reasons, each such cγc_{\gamma}^{*} is called an extension functional.

Although, formally, for each γΓn+1\gamma\in\Gamma_{n+1}, cγc_{\gamma}^{*} is defined on (Γn)\ell_{\infty}(\Gamma_{n}) in the end it can also be viewed as functional on (Γ)\ell_{\infty}(\Gamma), and thus on 𝔛(Γn,in)\mathfrak{X}_{(\Gamma_{n},i_{n})} if Assumption 3.1 is satisfied. This is done by identifying cγc_{\gamma}^{*} with cγrΓnc_{\gamma}^{*}\circ r_{{}_{\Gamma_{n}}}. Make the convention that for every γΓ1\gamma\in\Gamma_{1}, cγ=0c_{\gamma}^{*}=0 (this is natural as these are not truly extension functionals). By setting, for each γΓ\gamma\in\Gamma, dγ=eγcγd_{\gamma}^{*}=e_{\gamma}^{*}-c_{\gamma}^{*} it turns out that dγ(dγ)=δγ,γd_{\gamma}^{*}(d_{\gamma^{\prime}})=\delta_{\gamma,\gamma^{\prime}}, i.e., (dγ,dγ)γΓ(d_{\gamma},d_{\gamma}^{*})_{\gamma\in\Gamma} forms a biorthogonal system in 𝔛(Γn,in)×𝔛(Γn,in)\mathfrak{X}_{(\Gamma_{n},i_{n})}\times\mathfrak{X}_{(\Gamma_{n},i_{n})}^{*} (see [3, Proposition 2.17 (i), page 690]).

Bourgain and Delbaen pointed out in [8, Lemma 4.1, page 161] that if the extension functionals cγc_{\gamma}^{*} are of a certain form, then Assumption 3.1 will be automatically satisfied. Here, a specific case of this form is borrowed from the Argyros-Haydon construction in [4]. The following notation will be used frequently. For nn\in\mathbb{N} and γΔn\gamma\in\Delta_{n}, write rank(γ)=n\operatorname{rank}(\gamma)=n.

Proposition 3.2.

Assume that for every nn\in\mathbb{N} and γΔn+1\gamma\in\Delta_{n+1} the functional cγc_{\gamma}^{*} is either zero or of one of the following forms.

  1. (a)

    There exist an interval I{1,,n}I\subset\{1,\ldots,n\}, ηΓn\eta\in\Gamma_{n}, λ𝔻\lambda\in\mathbb{D}, and jj\in\mathbb{N} such that

    cγ=1mjλeηPI(n).c_{\gamma}^{*}=\frac{1}{m_{j}}\lambda e^{*}_{\eta}\circ P^{(n)}_{I}.
  2. (b)

    There exist ξΓn1\xi\in\Gamma_{n-1}, and interval II of {rank(ξ)+1,,n}\{\operatorname{rank}(\xi)+1,\ldots,n\}, ηΓn\eta\in\Gamma_{n} with rank(η)>rank(ξ)\operatorname{rank}(\eta)>\operatorname{rank}(\xi), λ𝔻\lambda\in\mathbb{D}, and jj\in\mathbb{N} such that

    cγ=eξ+1mjλeηPI(n).c_{\gamma}^{*}=e^{*}_{\xi}+\frac{1}{m_{j}}\lambda e^{*}_{\eta}\circ P^{(n)}_{I}.

Then, for every mnm\leq n\in\mathbb{N}, im,nm1/(m12)\|i_{m,n}\|\leq m_{1}/(m_{1}-2) and thus Assumption 3.1 is satisfied.

Proof.

See [8, Lemma 4.1, page 161] or [4, Theorem 3.5, page 11]. ∎

Remark 3.3.

If the assumptions of Proposition 3.2 are satisfied then

  1. (i)

    for all mnm\leq n\in\mathbb{N}, P[1,n]2\|P_{[1,n]}\|\leq 2 and P[m,n]3\|P_{[m,n]}\|\ \leq 3 and

  2. (ii)

    for all γΓ\gamma\in\Gamma, dγ2\|d_{\gamma}\|\leq 2 and dγ3\|d_{\gamma}^{*}\|\leq 3.

Indeed, P[1,n]=inrΓnP_{[1,n]}=i_{n}r_{{}_{\Gamma_{n}}} and inm1/(m12)4/3\|i_{n}\|\ \leq m_{1}/(m_{1}-2)\leq 4/3. Also, dγ=irank(γ)(eγ)d_{\gamma}=i_{\operatorname{rank}(\gamma)}(e_{\gamma}) and dγ=eγP[rank(γ),)d_{\gamma}^{*}=e_{\gamma}^{*}\circ P_{[\operatorname{rank}(\gamma),\infty)}.

3.3. A Mixed-Tsirelson Bourgain-Delbaen \mathscr{L}_{\infty}-space 𝔛mT\mathfrak{X}_{\mathrm{mT}}

As it was explained in [4], a Bourgain-Delbaen \mathscr{L}_{\infty}-space can be constructed by specifying a set of instructions (i.e., an algorithm) that takes as input disjoint index sets Δ1,,Δn\Delta_{1},\ldots,\Delta_{n} (with perhaps additional information encoded in them via certain functions) and returns an index set Δn+1\Delta_{n+1} together with functionals cγc_{\gamma}^{*}, γΔn+1\gamma\in\Delta_{n+1}, that adhere to the assumptions of Proposition 3.2.

The following set of instructions defines a Bourgain-Delbaen space 𝔛mT=𝔛(Γ¯n,i¯n)\mathfrak{X}_{\mathrm{mT}}=\mathfrak{X}_{(\bar{\Gamma}_{n},\bar{i}_{n})} that is based on XmTX_{\mathrm{mT}} from Section 2. The bar-notation is used to differentiate the objects associated to 𝔛mT\mathfrak{X}_{\mathrm{mT}} from the ones associated to the final space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. To facilitate the forthcoming construction of the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, there will be an additional involved parameter, namely an (as of yet unspecified) increasing sequence (Kn)n(K_{n})_{n} of finite subsets of KK. For each nn\in\mathbb{N}, denote 𝔻n={(k1/2n)ei(k2/2n)2π:k1,k2=1,,2n}\mathbb{D}_{n}=\{(k_{1}/2^{n})e^{i(k_{2}/2^{n})2\pi}:k_{1},k_{2}=1,\ldots,2^{n}\}. Each constructed γ\gamma will have additional information encoded in it, namely a weight and an age.

Instruction 3.4.

Put

Γ¯1=Δ¯1={(1,κ):κK1}\bar{\Gamma}_{1}=\bar{\Delta}_{1}=\{(1,\kappa):\kappa\in K_{1}\}

and for γΔ¯1\gamma\in\bar{\Delta}_{1} define weight(γ)=0\operatorname{weight}(\gamma)=0 and leave age(γ)\operatorname{age}(\gamma) undefined.

Assume that Γ¯1,,Γ¯n\bar{\Gamma}_{1},\ldots,\bar{\Gamma}_{n} as well as i¯1,2,,i¯n1,n\bar{i}_{1,2},\ldots,\bar{i}_{n-1,n} have been defined. Also assume that a function weight:Γ¯n{mj1:j}{0}\operatorname{weight}:\bar{\Gamma}_{n}\to\{m_{j}^{-1}:j\in\mathbb{N}\}\cup\{0\} and a partially defined function age:Γ¯n\operatorname{age}:\bar{\Gamma}_{n}\to\mathbb{N} have been constructed.

The set Δ¯n+1\bar{\Delta}_{n+1} is defined as the disjoint union of sets Δ¯n+10\bar{\Delta}^{0}_{n+1}, Δ¯n+1(a)\bar{\Delta}^{\text{\ref{age zero}}}_{n+1}, and Δ¯n+1(b)\bar{\Delta}^{\text{\ref{age nonzero}}}_{n+1}. First, put

Δ¯n+10={(n+1,κ):κKn+1}\bar{\Delta}_{n+1}^{0}=\Big{\{}(n+1,\kappa):\kappa\in K_{n+1}\Big{\}}

and for γΔ¯n+1\gamma\in\bar{\Delta}_{n+1} define weight(γ)=0\operatorname{weight}(\gamma)=0, leave age(γ)\operatorname{age}(\gamma) undefined, and put c¯γ=0\bar{c}_{\gamma}^{*}=0.

Let

Δ¯n+1(a)={\displaystyle\bar{\Delta}_{n+1}^{\text{\ref{age zero}}}=\Big{\{} (n+1,I,η,λ,mj1,κ):I is an inteval of {1,,n},\displaystyle(n+1,I,\eta,\lambda,m_{j}^{-1},\kappa):\;I\text{ is an inteval of }\{1,\ldots,n\},
ηΓ¯n,λ𝔻n+1,j{1,,n+1}, and κKn+1}.\displaystyle\eta\in\bar{\Gamma}_{n},\;\lambda\in\mathbb{D}_{n+1},\;j\in\{1,\ldots,n+1\},\text{ and }\kappa\in K_{n+1}\Big{\}}.

For γ=(n+1,I,η,λ,mj1,κ)Δ¯n+1(a)\gamma=(n+1,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\bar{\Delta}^{\text{\ref{age zero}}}_{n+1} define weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1}, age(γ)=1\operatorname{age}(\gamma)=1, and

c¯γ=1mjλe¯ηP¯I(n).\bar{c}_{\gamma}^{*}=\frac{1}{m_{j}}\lambda\bar{e}^{*}_{\eta}\circ\bar{P}_{I}^{(n)}.

Note that for γΔ¯n+1(a)\gamma\in\bar{\Delta}_{n+1}^{\text{\ref{age zero}}}, c¯γ\bar{c}_{\gamma}^{*} is of type (a) from Proposition 3.2.

Next, let

Δ¯n+1(b)={\displaystyle\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}}=\Big{\{} (n+1,ξ,I,η,λ,mj1,κ):ξΓ¯n1 with weight(ξ)=mj1 and\displaystyle(n+1,\xi,I,\eta,\lambda,m_{j}^{-1},\kappa):\;\xi\in\bar{\Gamma}_{n-1}\text{ with }\operatorname{weight}(\xi)=m_{j}^{-1}\text{ and}
age(ξ)<nj,I is an inteval of {rank(ξ)+1,,n},\displaystyle\operatorname{age}(\xi)<n_{j},\;I\text{ is an inteval of }\{\operatorname{rank}(\xi)+1,\ldots,n\},
ηΓ¯n,λ𝔻¯n+1,j{1,,n+1}, and κKn+1}.\displaystyle\eta\in\bar{\Gamma}_{n},\;\lambda\in\bar{\mathbb{D}}_{n+1},\;j\in\{1,\ldots,n+1\},\text{ and }\kappa\in K_{n+1}\Big{\}}.

For (n+1,ξ,I,η,λ,mj1,κ)Δ¯n+1(b)(n+1,\xi,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}} define weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1}, age(γ)=age(ξ)+1\operatorname{age}(\gamma)=\operatorname{age}(\xi)+1, and

c¯γ=e¯ξ+1mjλe¯ηP¯I(n).\bar{c}_{\gamma}^{*}=\bar{e}_{\xi}^{*}+\frac{1}{m_{j}}\lambda\bar{e}_{\eta}^{*}\bar{P}_{I}^{(n)}.

Note that for γΔ¯n+1(b)\gamma\in\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}}, c¯γ\bar{c}_{\gamma}^{*} is of type (b) from Proposition 3.2.

Define Δ¯n+1=Δ¯n+10Δ¯n+1(a)Δ¯n+1(b)\bar{\Delta}_{n+1}=\bar{\Delta}_{n+1}^{0}\cup\bar{\Delta}_{n+1}^{\text{\ref{age zero}}}\cup\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}}.

This set of instructions defines the Bourgain-Delbaen \mathscr{L}_{\infty}-space 𝔛mT\mathfrak{X}_{\mathrm{mT}}. Note that the space 𝔛mT\mathfrak{X}_{\mathrm{mT}} is similar to the space 𝔅mT\mathfrak{B}_{\mathrm{mT}} from [4]. There are three differences. The first one is that here, for each nn\in\mathbb{N}, there are several copies of each extension functional (even the zero one) indexed over KnK_{n}. The second difference is that here no convex combinations of eηe_{\eta}^{*} are allowed. As already mentioned, this is necessary to perform a saturation under constraints with increasing weights in the flavour of [7]. The final difference is that in [4] there was no need to use the entire collection of intervals II used here. This full collection was also used in the construction in [6].

The claimed connection to the space XmTX_{\mathrm{mT}} is made a lot clearer by the evaluation analysis of each γΓ¯\gamma\in\bar{\Gamma}. This concept is from [4]. Note that every γΓ\gamma\in\Gamma with non-zero weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1} has an age(γ)=anj\operatorname{age}(\gamma)=a\leq n_{j}.

Proposition 3.5.

Let γΓ¯\gamma\in\bar{\Gamma} have weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1} and age(γ)=anj\operatorname{age}(\gamma)=a\leq n_{j}. Then, there exist

  1. (i)

    ξ1,,ξaΓ¯\xi_{1},\ldots,\xi_{a}\in\bar{\Gamma} with jrank(ξ1)<<rank(ξa)j\leq\operatorname{rank}(\xi_{1})<\cdots<\operatorname{rank}(\xi_{a}) and ξa=γ\xi_{a}=\gamma,

  2. (ii)

    intervals I1I_{1} of [1,rank(ξ1))[1,\operatorname{rank}(\xi_{1})) and IrI_{r} of (rank(ξr1),rank(ξr))(\operatorname{rank}(\xi_{r-1}),\operatorname{rank}(\xi_{r})), r=2,,ar=2,\ldots,a,

  3. (iii)

    ηrΓ¯\eta_{r}\in\bar{\Gamma} with rank(ηr)<rank(ξr)\operatorname{rank}(\eta_{r})<\operatorname{rank}(\xi_{r}), for r=1,,ar=1,\ldots,a, and

  4. (iv)

    λr𝔻rank(ξr)\lambda_{r}\in\mathbb{D}_{\operatorname{rank}(\xi_{r})}, for r=1,,ar=1,\ldots,a,

such that

e¯γ=r=1ad¯ξr+1mjr=1aλre¯ηrP¯Ir.\bar{e}_{\gamma}^{*}=\sum_{r=1}^{a}\bar{d}_{\xi_{r}}^{*}+\frac{1}{m_{j}}\sum_{r=1}^{a}\lambda_{r}\bar{e}_{\eta_{r}}^{*}\circ\bar{P}_{I_{r}}.
Proof.

This is proved by induction on age(γ)\operatorname{age}(\gamma), by writing e¯γ=d¯γ+c¯γ\bar{e}_{\gamma}^{*}=\bar{d}_{\gamma}^{*}+\bar{c}_{\gamma}^{*} and unravelling the definition of c¯γ\bar{c}_{\gamma}^{*}. See [4, Proposition 4.5, page 15] for further details. ∎

Intuitively, each e¯γ\bar{e}_{\gamma}^{*} can be obtained as the outcome of a certain kind of (mj1,𝒜nj)(m_{j}^{-1},\mathcal{A}_{n_{j}})-operation applied to other e¯η\bar{e}_{\eta}^{*}.

3.4. Definition of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

To define the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, an appropriate subset Γ\Gamma of Γ¯\bar{\Gamma} will be chosen. Recall that bar-notation is used for objects relevant to 𝔛mT\mathfrak{X}_{\mathrm{mT}} whereas objects without bars are related to 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. Start with the sets (Δ¯n)n(\bar{\Delta}_{n})_{n}. Put Δ1=Δ¯1\Delta_{1}=\bar{\Delta}_{1} and for each n2n\geq 2 choose a suitable ΔnΔ¯n\Delta_{n}\subset\bar{\Delta}_{n}. Among other specialized properties, the set Γ=nΔn\Gamma=\cup_{n}\Delta_{n} will be a self-determined subset of Γ¯=nΔ¯n\bar{\Gamma}=\cup_{n}\bar{\Delta}_{n}. This means that for every nn\in\mathbb{N} and γΔn+1\gamma\in\Delta_{n+1}, whichever ξ\xi or η\eta appear in the defining formula of c¯γ\bar{c}_{\gamma}^{*} must be in Γn\Gamma_{n}. For example, if γΔn+1Δ¯n+1(b)\gamma\in\Delta_{n+1}\cap\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}} and c¯γ=e¯ξ+mj1λe¯ηP¯I(n)\bar{c}_{\gamma}^{*}=\bar{e}_{\xi}^{*}+m_{j}^{-1}\lambda\bar{e}^{*}_{\eta}\circ\bar{P}^{(n)}_{I}, then ξ,ηΓn\xi,\eta\in\Gamma_{n}. Then the functional cγ=eξ+mj1λeηPI(n):(Γn)c_{\gamma}^{*}=e_{\xi}^{*}+m_{j}^{-1}\lambda e^{*}_{\eta}\circ P^{(n)}_{I}:\ell_{\infty}(\Gamma_{n})\to\mathbb{C} is well defined and is of type (b) from Proposition 3.2. A similar situation occurs for γΔn+1\gamma\in\Delta_{n+1} that are in Δ¯n+1(a)\bar{\Delta}_{n+1}^{\text{\ref{age zero}}} or Δ¯n+10\bar{\Delta}^{0}_{n+1}. Therefore, this process yields a Bourgain-Delbaen \mathscr{L}_{\infty}-space 𝔛(Γn,in)\mathfrak{X}_{(\Gamma_{n},i_{n})}, which will be the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. This technique was also used by Tarbard in [43] and [44], by Argyros and the author in [6], and by Manoussakis, Pelczar-Barwacz, and Świȩtek in [29]. It was formulated as a method for general Bourgain-Delbaen spaces by Argyros and the author in [6]. According to [6, Proposition 1.12, page 1893], the restriction map rΓr_{{}_{\Gamma}} on 𝔛mT\mathfrak{X}_{\mathrm{mT}} is a quotient map onto 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. This process is analogous to the fact that WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is built as a subset of WmTW_{\mathrm{mT}}.

The choice of Γ\Gamma requires the enforcement of certain constraints onto its member γ\gamma. The first one comes from a Maurey-Rosenthal coding function applied to the construction of cγc^{*}_{\gamma} when γ\gamma has odd weight. Therefore, fix an injection σ:Γ¯\sigma:\bar{\Gamma}\to\mathbb{N} such that for every γΓ¯\gamma\in\bar{\Gamma}, m4σ(γ)>2rank(γ)m_{4\sigma(\gamma)}>2^{\operatorname{rank}(\gamma)}.

The second constraint pertains to the weight of η\eta, which needs to be sufficiently small, whenever it appears in the definition of a given γ\gamma. For this, it is necessary to define an additional weight function. For every interval II of \mathbb{N} and γΓ¯\gamma\in\bar{\Gamma} define

(2) weight(I,γ)={0if rank(γ)min(I),weight(γ)otherwise.\operatorname{weight}(I,\gamma)=\left\{\begin{array}[]{ll}0&\mbox{if }\operatorname{rank}(\gamma)\leq\min(I),\\ \operatorname{weight}(\gamma)&\mbox{otherwise.}\end{array}\right.

The utility of this is that in the end, for every γΓ\gamma\in\Gamma, if rank(γ)=min(I)\operatorname{rank}(\gamma)=\min(I) then eγPI=dγe_{\gamma}^{*}\circ P_{I}=d_{\gamma}^{*}. This, being a basis element, models the behaviour of basis elements in WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, which by convention have zero weight. If rank(γ)<min(I)\operatorname{rank}(\gamma)<\min(I), then eγPI=0e_{\gamma}^{*}\circ P_{I}=0.

The final constraint relates to the metric space KK and is encoded in the sets (Kn)n(K_{n})_{n} which were used as a parameter in the construction of Γ¯\bar{\Gamma}. Recall that for every γΓ¯\gamma\in\bar{\Gamma}, with rank(γ)=n\operatorname{rank}(\gamma)=n, its last coordinate is some κKn\kappa\in K_{n}. Define the function κ:Γ¯nKnK\kappa:\bar{\Gamma}\to\cup_{n}K_{n}\subset K that retrieves from each γ\gamma its final coordinate κ=κ(γ)\kappa=\kappa(\gamma). This will be used as follows. For any γ\gamma and for whichever ξ\xi, η\eta appear in the defining formula of cγc_{\gamma}^{*}, the elements κ(γ)\kappa(\gamma), κ(ξ)\kappa(\xi), and κ(η)\kappa(\eta) must be close to one another in KK.

To specify the sequence of sets (Kn)n(K_{n})_{n}, fix a countable dense subset 𝒜\mathcal{A} of C(K)C(K) that is closed under \mathbb{C}_{\mathbb{Q}}-linear combinations. Enumerate the set ={ϕ𝒜:ϕ1}\mathcal{B}=\{\phi\in\mathcal{A}:\|\phi\|_{\infty}\leq 1\} as {ϕ1,ϕ2,}\{\phi_{1},\phi_{2},\ldots\}. For all nn\in\mathbb{N} define

n={θkϕi1ϕi2ϕik:k,0θnn+1, and 1i1,i2,,ikn}.\mathcal{B}_{n}=\Big{\{}\theta^{k}\phi_{i_{1}}\phi_{i_{2}}\cdots\phi_{i_{k}}:k\in\mathbb{N},0\leq\theta\leq\frac{n}{n+1},\text{ and }1\leq i_{1},i_{2},\ldots,i_{k}\leq n\Big{\}}.

Then, 12\mathcal{B}_{1}\subset\mathcal{B}_{2}\subset\cdots and each n\mathcal{B}_{n} is a compact multiplicative subsemigroup of the unit ball of C(K)C(K). Define the equivalent metric on KK

ϱ(κ,κ)=n=12nmax{|ϕ(κ)ϕ(κ)|:ϕn}.\varrho(\kappa,\kappa^{\prime})=\sum_{n=1}^{\infty}2^{-n}\max\Big{\{}\Big{|}\phi(\kappa)-\phi(\kappa^{\prime})\Big{|}:\phi\in\mathcal{B}_{n}\Big{\}}.

This change of metric will make it possible to renorm the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} to achieve an isometric result. What will eventually be required is that (n)n(\mathcal{B}_{n})_{n} is increasing and that for each nn\in\mathbb{N} every ϕn\phi\in\mathcal{B}_{n} is 2n2^{n}-Lipschitz and n\mathcal{B}_{n} is a multiplicative semigroup. Note that the equivalence of the metric follows from the compactness of the sets n\mathcal{B}_{n}, which will not be used again. Now, fix an increasing sequence of finite subsets (Kn)n(K_{n})_{n} of KK so that, for each nn\in\mathbb{N}, KnK_{n} is an mn12(n+1)m_{n}^{-1}2^{-(n+1)}-net of KK (with respect to ϱ\varrho).

It is time to define the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

Instruction 3.6.

Define Γ1=Δ1=Δ¯1\Gamma_{1}=\Delta_{1}=\bar{\Delta}_{1}. Assume that Γ1,,Γn\Gamma_{1},\ldots,\Gamma_{n} as well as i1,2,,in1,ni_{1,2},\ldots,i_{n-1,n} have been defined.

The set Δn+1\Delta_{n+1} is defined as the disjoint union of sets Δn+10\Delta^{0}_{n+1}, Δn+1(a),even\Delta^{\text{\ref{age zero},even}}_{n+1}, Δn+1(a),odd\Delta^{\text{\ref{age zero},odd}}_{n+1}, Δn+1(b),even\Delta^{\text{\ref{age nonzero},even}}_{n+1}, and Δn+1(b),odd\Delta^{\text{\ref{age nonzero},odd}}_{n+1}. First, put Δn+10=Δ¯n+10\Delta^{0}_{n+1}=\bar{\Delta}^{0}_{n+1}. Let

Δn+1(a),even={\displaystyle\Delta_{n+1}^{\text{\ref{age zero},even}}=\Big{\{} (n+1,I,η,λ,mj1,κ)Δ¯n+1(a):j(2) and ηΓn}\displaystyle(n+1,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\bar{\Delta}_{n+1}^{\text{\ref{age zero}}}:\;j\in(2\mathbb{N})\text{ and }\eta\in\Gamma_{n}\Big{\}}
Δn+1(a),odd={\displaystyle\Delta_{n+1}^{\text{\ref{age zero},odd}}=\Big{\{} (n+1,I,η,λ,mj1,κ)Δ¯n+1(a):j(21),\displaystyle(n+1,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\bar{\Delta}_{n+1}^{\text{\ref{age zero}}}:\;j\in(2\mathbb{N}-1),
ηΓn and weight(η)=m4i21, for some i}.\displaystyle\eta\in\Gamma_{n}\text{ and }\operatorname{weight}(\eta)=m^{-1}_{4i-2}\text{, for some }i\in\mathbb{N}\Big{\}}.

Put Δn+1(a)=Δn+1(a),evenΔn+1(a),odd\Delta^{\text{\ref{age zero}}}_{n+1}=\Delta^{\text{\ref{age zero},even}}_{n+1}\cup\Delta^{\text{\ref{age zero},odd}}_{n+1} and for γ=(n+1,I,η,λ,mj1,κ)Δn+1(a)\gamma=(n+1,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\Delta^{\text{\ref{age zero}}}_{n+1} define

cγ=1mjλeηPI(n).c_{\gamma}^{*}=\frac{1}{m_{j}}\lambda e^{*}_{\eta}\circ P_{I}^{(n)}.

Note that for γΔn+1(a)\gamma\in\Delta^{\text{\ref{age zero}}}_{n+1}, cγc_{\gamma}^{*} is of type (a) from Proposition 3.2.

Next, let

Δn+1(b),even={\displaystyle\Delta_{n+1}^{\text{\ref{age nonzero},even}}=\Big{\{} (n+1,ξ,I,η,λ,mj1,κ)Δ¯n+1(b):j(2),ξΓn1,ηΓn\displaystyle(n+1,\xi,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}}:\;j\in(2\mathbb{N}),\;\xi\in\Gamma_{n-1},\;\eta\in\Gamma_{n}
with weight(I,η)2rank(ξ),ϱ(κ(ξ),κ)mj12rank(ξ),\displaystyle\text{with }\operatorname{weight}(I,\eta)\leq 2^{-\operatorname{rank}(\xi)},\;\varrho(\kappa(\xi),\kappa)\leq m_{j}^{-1}2^{-\operatorname{rank}(\xi)},
and ϱ(κ(η),κ)2rank(ξ)}.\displaystyle\text{and }\varrho(\kappa(\eta),\kappa)\leq 2^{-\operatorname{rank}(\xi)}\Big{\}}.
Δn+1(b),odd={\displaystyle\Delta_{n+1}^{\text{\ref{age nonzero},odd}}=\Big{\{} (n+1,ξ,I,η,λ,mj1,κ)Δ¯n+1(b):j(21),ξΓn1,\displaystyle(n+1,\xi,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\bar{\Delta}_{n+1}^{\text{\ref{age nonzero}}}:\;j\in(2\mathbb{N}-1),\;\xi\in\Gamma_{n-1},
ηΓn with weight(η)=m4σ(ξ)1,ϱ(κ(ξ),κ)mj12rank(ξ),\displaystyle\eta\in\Gamma_{n}\text{ with }\operatorname{weight}(\eta)=m^{-1}_{4\sigma(\xi)},\;\varrho(\kappa(\xi),\kappa)\leq m_{j}^{-1}2^{-\operatorname{rank}(\xi)},
and ϱ(κ(η),κ)2rank(ξ)}.\displaystyle\text{and }\varrho(\kappa(\eta),\kappa)\leq 2^{-\operatorname{rank}(\xi)}\Big{\}}.

Put Δn+1(b)=Δn+1(b),evenΔn+1(b),odd\Delta_{n+1}^{\text{\ref{age nonzero}}}=\Delta_{n+1}^{\text{\ref{age nonzero},even}}\cup\Delta_{n+1}^{\text{\ref{age nonzero},odd}} and for (n+1,ξ,I,η,λ,mj1,κ)Δn+1(b)(n+1,\xi,I,\eta,\lambda,m_{j}^{-1},\kappa)\in\Delta_{n+1}^{\text{\ref{age nonzero}}} define

cγ=eξ+1mjλeηPI(n).c_{\gamma}^{*}=e_{\xi}^{*}+\frac{1}{m_{j}}\lambda e_{\eta}^{*}P_{I}^{(n)}.

Note that for γΔn+1(b)\gamma\in\Delta_{n+1}^{\text{\ref{age nonzero}}}, cγc_{\gamma}^{*} is of type (b) from Proposition 3.2.

Define Δn+1=Δn+10Δn+1(a)Δn+1(b)\Delta_{n+1}=\Delta_{n+1}^{0}\cup\Delta_{n+1}^{\text{\ref{age zero}}}\cup\Delta_{n+1}^{\text{\ref{age nonzero}}} and denote 𝔛C(K)=𝔛(Γn,in)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}=\mathfrak{X}_{(\Gamma_{n},i_{n})}.

This time, the evaluation analysis of a γΓ\gamma\in\Gamma can be supplemented with information about weights and metric distance. This information resembles the conditions imposed on WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

Proposition 3.7.

Let γΓ\gamma\in\Gamma have weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1} and age(γ)=anj\operatorname{age}(\gamma)=a\leq n_{j}. Then, there exist

  1. (i)

    ξ1,,ξaΓ\xi_{1},\ldots,\xi_{a}\in\Gamma with jrank(ξ1)<<rank(ξa)j\leq\operatorname{rank}(\xi_{1})<\cdots<\operatorname{rank}(\xi_{a}) and ξa=γ\xi_{a}=\gamma,

  2. (ii)

    intervals I1I_{1} of [1,rank(ξ1))[1,\operatorname{rank}(\xi_{1})) and IrI_{r} of (rank(ξr1),rank(ξr))(\operatorname{rank}(\xi_{r-1}),\operatorname{rank}(\xi_{r})), r=2,,ar=2,\ldots,a,

  3. (iii)

    ηrΓ\eta_{r}\in\Gamma with rank(ηr)<rank(ξr)\operatorname{rank}(\eta_{r})<\operatorname{rank}(\xi_{r}), for 1=2,,a1=2,\ldots,a, and

  4. (iv)

    λr𝔻rank(ξr)\lambda_{r}\in\mathbb{D}_{\operatorname{rank}(\xi_{r})}, r=1,,ar=1,\ldots,a,

such that

(3) eγ=r=1adξr+1mjr=1aλreηrPIre_{\gamma}^{*}=\sum_{r=1}^{a}d_{\xi_{r}}^{*}+\frac{1}{m_{j}}\sum_{r=1}^{a}\lambda_{r}e_{\eta_{r}}^{*}\circ P_{I_{r}}

and for r=2,,ar=2,\ldots,a,

(4) weight(Ir,ηr)\displaystyle\operatorname{weight}(I_{r},\eta_{r}) 2rank(ξr1),\displaystyle\leq 2^{-\operatorname{rank}(\xi_{r-1})},
(5) ϱ(κ(ξr1),κ(ξr))\displaystyle\varrho\big{(}\kappa(\xi_{r-1}),\kappa(\xi_{r})\big{)} mj12rank(ξr1), and\displaystyle\leq m_{j}^{-1}2^{-\operatorname{rank}(\xi_{r-1})},\text{ and}
(6) ϱ(κ(ηr),κ(ξr))\displaystyle\varrho\big{(}\kappa(\eta_{r}),\kappa(\xi_{r})\big{)} 2rank(ξr1)\displaystyle\leq 2^{-\operatorname{rank}(\xi_{r-1})}

The representation (4) is called the evaluation analysis of γ\gamma.

Proof.

This is proved by induction on age(γ)=a\operatorname{age}(\gamma)=a. If age(γ)=1\operatorname{age}(\gamma)=1, put ξ1=γ\xi_{1}=\gamma and write cγ=mj1λ1eη1PI1c_{\gamma}^{*}=m_{j}^{-1}\lambda_{1}e_{\eta_{1}}^{*}\circ P_{I_{1}}. Therefore one has eγ=dγ+cγ=dξ1+mj1λ1eη1PI1e_{\gamma}^{*}=d_{\gamma}^{*}+c_{\gamma}^{*}=d_{\xi_{1}}^{*}+m_{j}^{-1}\lambda_{1}e_{\eta_{1}}^{*}\circ P_{I_{1}}. If the statement holds for all γΓ\gamma\in\Gamma with age(γ)=a\operatorname{age}(\gamma)=a, let γΓ\gamma\in\Gamma with age(γ)=a+1\operatorname{age}(\gamma)=a+1. Put ξa+1=γ\xi_{a+1}=\gamma and write cγ=eξa+mj1λa+1eηa+1c_{\gamma}^{*}=e^{*}_{\xi_{a}}+m_{j}^{-1}\lambda_{a+1}e_{\eta_{a+1}}^{*}. By the definition of Δn+1(b)\Delta_{n+1}^{\text{\ref{age nonzero}}},

weight(Ia+1,γa+1)\displaystyle\operatorname{weight}(I_{a+1},\gamma_{a+1}) 2rank(ξa),\displaystyle\leq 2^{-\operatorname{rank}(\xi_{a})},
ϱ(κ(ξa),κ(ξa+1))\displaystyle\varrho(\kappa(\xi_{a}),\kappa(\xi_{a+1})) mj12rank(ξa), and\displaystyle\leq m_{j}^{-1}2^{\operatorname{rank}(\xi_{a})},\text{ and}
ϱ(κ(ηa+1),κ(ξa+1))\displaystyle\varrho\big{(}\kappa(\eta_{a+1}),\kappa(\xi_{a+1})\big{)} 2rank(ξa).\displaystyle\leq 2^{-\operatorname{rank}(\xi_{a})}.

Write eγ=dξa+1+cξa+1=dξa+1+mj1λa+1eηa+1+eξae_{\gamma}^{*}=d_{\xi_{a+1}}^{*}+c_{\xi_{a+1}}^{*}=d_{\xi_{a+1}}^{*}+m_{j}^{-1}\lambda_{a+1}e_{\eta_{a+1}}^{*}+e^{*}_{\xi_{a}} and recover the remaining information from the inductive hypothesis applied to eξae^{*}_{\xi_{a}}. ∎

Similar to WC(K)W_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, the set Γ\Gamma is in a certain sense closed under the (m2j1,𝒜n2j)(m_{2j}^{-1},\mathcal{A}_{n_{2j}})-operations applied to sequences (λreηrPIr)r(\lambda_{r}e^{*}_{\eta_{r}}\circ P_{I_{r}})_{r} that satisfy a condition similar to having essentially rapidly converging supports.

Proposition 3.8.

Let jj\in\mathbb{N}, (qr)r=1n2j(q_{r})_{r=1}^{n_{2j}} be a strictly sequence of natural numbers, (Ir)r=1n2j(I_{r})_{r=1}^{n_{2j}} be finite intervals of \mathbb{N}, ηrΓ\eta_{r}\in\Gamma and λr𝔻qr\lambda_{r}\in\mathbb{D}_{q_{r}}, for r=1,,n2jr=1,\ldots,n_{2j}, and κ0K\kappa_{0}\in K. Assume that the following are satisfied.

  1. (i)

    2jq12j\leq q_{1}, I1[1,q1)I_{1}\subset[1,q_{1}), and Ir(qr1,qr)I_{r}\subset(q_{r-1},q_{r}) , for r=2,,n2jr=2,\ldots,n_{2j}.

  2. (ii)

    rank(ηr)<qr\operatorname{rank}(\eta_{r})<q_{r}, for r=1,,n2jr=1,\ldots,n_{2j},

  3. (iii)

    weight(Ir,ηr)2qr1\operatorname{weight}(I_{r},\eta_{r})\leq 2^{-q_{r-1}}, for r=2,,n2jr=2,\ldots,n_{2j}.

  4. (iv)

    ϱ(κ(ηk),κ0)2(qr1+1)\varrho(\kappa(\eta_{k}),\kappa_{0})\leq 2^{-(q_{r-1}+1)}, for r=2,,n2jr=2,\ldots,n_{2j}.

Then, there exists γΓ\gamma\in\Gamma with weight(γ)=m2j1\operatorname{weight}(\gamma)=m_{2j}^{-1}, rank(γ)=qn2j\operatorname{rank}(\gamma)=q_{n_{2j}}, and ϱ(κ(γ),κ0)2qn2j\varrho(\kappa(\gamma),\kappa_{0})\leq 2^{-q_{n_{2j}}} such that eγe_{\gamma}^{*} has an evaluation analysis

eγ=r=1n2jdξr+1m2jr=1n2jλreηrPIr,e_{\gamma}^{*}=\sum_{r=1}^{n_{2j}}d_{\xi_{r}}^{*}+\frac{1}{m_{2j}}\sum_{r=1}^{n_{2j}}\lambda_{r}e_{\eta_{r}}^{*}\circ P_{I_{r}},

where rank(ξr)=qr\operatorname{rank}(\xi_{r})=q_{r}, for r=1,,n2jr=1,\ldots,n_{2j}.

Proof.

Recall that KqrK_{q_{r}} is a mqr12(qr+1)m_{q_{r}}^{-1}2^{-(q_{r}+1)}-net of KK and thus it is possible to pick κrKqr\kappa_{r}\in K_{q_{r}} with ϱ(κr,κ0)mqr12(qr+1)\varrho(\kappa_{r},\kappa_{0})\leq m_{q_{r}}^{-1}2^{-(q_{r}+1)}. In particular, for r=2,,n2jr=2,\ldots,n_{2j}, ϱ(κr,κr1)m2j12qr1\varrho(\kappa_{r},\kappa_{r-1})\leq m_{2j}^{-1}2^{-q_{r-1}}. Next, choose inductively ξ1,,ξn2j=γ\xi_{1},\ldots,\xi_{n_{2j}}=\gamma that satisfy the following properties.

  1. (a)

    rank(ξr)=qr\operatorname{rank}(\xi_{r})=q_{r}, for r=1,,n2j1r=1,\ldots,n_{2j-1}

  2. (b)

    κ(ξr)=κr\kappa(\xi_{r})=\kappa_{r}, for r=1,,n2j1r=1,\ldots,n_{2j-1}

  3. (c)

    cξ1=m2j1λ1eγ1PI1c_{\xi_{1}}^{*}=m_{2j}^{-1}\lambda_{1}e_{\gamma_{1}}^{*}\circ P_{I_{1}} and cξr=eξr1+m2j1λreηrPIrc_{\xi_{r}}^{*}=e_{\xi_{r-1}}^{*}+m_{2j}^{-1}\lambda_{r}e_{\eta_{r}}\circ P_{I_{r}}^{*}, for r=2,,n2j1r=2,\ldots,n_{2j-1}.

Note that ξ1=(q1,I1,η1,λ1,m2j1,κ1)\xi_{1}=(q_{1},I_{1},\eta_{1},\lambda_{1},m_{2j}^{-1},\kappa_{1}) meets all the required conditions to be in Δq1(a),even\Delta_{q_{1}}^{\text{\ref{age zero},even}}. If, for some 2n2j12\leq n_{2j-1}, ξr1\xi_{r-1} has been chosen then, ϱ(κ(ηr),κr)ϱ(κ(ηr),κ0)+ϱ(κr,κ0)2qr1\varrho(\kappa(\eta_{r}),\kappa_{r})\leq\varrho(\kappa(\eta_{r}),\kappa_{0})+\varrho(\kappa_{r},\kappa_{0})\leq 2^{-q_{r-1}}. Therefore, ξr=(qr,ξr1,Ir,ηr,λr,m2j1,κr)\xi_{r}=(q_{r},\xi_{r-1},I_{r},\eta_{r},\lambda_{r},m_{2j}^{-1},\kappa_{r}) is in Δqr(b),even\Delta_{q_{r}}^{\text{\ref{age nonzero},even}}. It is the straightforward to check that γ=ξn2j\gamma=\xi_{n_{2j}} has the desired weight and proximity to κ0\kappa_{0}. The fact that it also has the desired evaluation analysis follows from the proof of Proposition 3.7. ∎

Remark 3.9.

A similar process can be carried out to construct γ\gamma of odd weight m2j1m_{2j}^{-1}. The difference is that in each step the resulting ξr\xi_{r} determines, via the coding function σ\sigma, the weight m4σ(ξr)1m^{-1}_{4\sigma(\xi_{r})} that ηr+1{\eta_{r+1}} is allowed to have.

4. Diagonal operators on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} and their image in 𝒞𝒶𝓁(𝔛𝒞(𝒦))\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})

For every continuous function ϕ:K\phi:K\to\mathbb{C} denote by ϕ^\hat{\phi} the linear operator on {dγ:γΓ}\langle\{d_{\gamma}:\gamma\in\Gamma\}\rangle given by ϕ^(dγ)=ϕ(κ(γ))dγ\hat{\phi}(d_{\gamma})=\phi(\kappa(\gamma))d_{\gamma}. In this section it will be proved that whenever ϕ\phi is Lipschitz then ϕ^\hat{\phi} extends to a bounded linear operator on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. The derived estimates yield a natural Banach algebra embedding Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}). At the end of this section it will be shown that 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} admits an equivalent norm that turns this into a homomorphic isometric embedding. All the results of this section deal with the “unconditional” structure of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, i.e., the special properties of odd-weight functionals are not used. These will be necessary further down the road when it will be established that Ψ\Psi is onto as well.

For ϕ\phi as above also define ϕ^\hat{\phi} on {dγ:γΓ}𝔛C(K)\langle\{d_{\gamma}^{*}:\gamma\in\Gamma\}\rangle\subset\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{*} using the same formula, i.e., for all γΓ\gamma\in\Gamma, ϕ^(dγ)=ϕ(κ(γ))dγ\hat{\phi}(d^{*}_{\gamma})=\phi(\kappa(\gamma))d^{*}_{\gamma}. Under this notation, for all x{dγ:γΓ}x\in\langle\{d_{\gamma}:\gamma\in\Gamma\}\rangle and f{dγ:γΓ}f\in\langle\{d_{\gamma}^{*}:\gamma\in\Gamma\}\rangle, f(ϕ^x)=(ϕ^f)(x)f(\hat{\phi}x)=(\hat{\phi}f)(x). That is, ϕ^\hat{\phi} is its own dual. Note that for every γΓ\gamma\in\Gamma and interval II of \mathbb{N}, eγPIe_{\gamma}^{*}\circ P_{I} is in the linear span of {dγ:γΓ}\{d_{\gamma^{\prime}}^{*}:\gamma^{\prime}\in\Gamma\} (see, e.g., [3, Proposition 2.17 (ii), page 690]). Therefore, ϕ^(eγPI)\hat{\phi}(e_{\gamma}^{*}\circ P_{I}) is always well defined.

4.1. Boundedness of diagonal operators

The following statement is analogous to Proposition 2.9. The additional precision will be required later in this section to define a renorming of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

Proposition 4.1.

For every L,M0L,M\geq 0 there exists NN\in\mathbb{N} with the following property. For every ϕ:K\phi:K\to\mathbb{C} that is LL-Lipschitz with ϕM\|\phi\|_{\infty}\leq M, for every γΓ\gamma\in\Gamma, and for every interval II of \mathbb{N} with min(I)N\min(I)\geq N,

(7) ϕ^(eγPI)ϕ(κ(γ))eγPI7weight(I,γ)M.\Big{\|}\hat{\phi}\big{(}e^{*}_{\gamma}\circ P_{I}\big{)}-\phi\big{(}\kappa(\gamma)\big{)}e_{\gamma}^{*}\circ P_{I}\Big{\|}\leq 7\operatorname{weight}(I,\gamma)M.

In particular, ϕ^:𝔛C(K)𝔛C(K)\hat{\phi}:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is bounded and ϕ^P[1,N)ϕ^4M\big{\|}\hat{\phi}-P_{[1,N)}\hat{\phi}\big{\|}\leq 4M.

The proof comes down to taking the evaluation analysis of a given γ\gamma and applying an inductive hypothesis to its components, just like in the proof of Proposition 2.9. For the sake of tidiness, some computations have been isolated and gathered in the following.

Lemma 4.2.

Let γΓ\gamma\in\Gamma have weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1} and age(γ)=a>1\operatorname{age}(\gamma)=a>1. Following the notation of Proposition 3.7, for every 1r0<a1\leq r_{0}<a,

(8) r=r0+1aweight(Ir,ηr)\displaystyle\sum_{r=r_{0}+1}^{a}\operatorname{weight}(I_{r},\eta_{r}) 22rank(ξr0),\displaystyle\leq 2\cdot 2^{-\operatorname{rank}(\xi_{r_{0}})},
(9) r=r0aϱ(κ(ξr),κ(γ))\displaystyle\sum_{r=r_{0}}^{a}\varrho\big{(}\kappa(\xi_{r}),\kappa(\gamma)\big{)} 4mj2rank(ξr0), and\displaystyle\leq\frac{4}{m_{j}}2^{-\operatorname{rank}(\xi_{r_{0}})},\text{ and}
(10) r=r0+1aϱ(κ(ηr),κ(γ))\displaystyle\sum_{r=r_{0}+1}^{a}\varrho\big{(}\kappa(\eta_{r}),\kappa(\gamma)\big{)} 32rank(ξr0).\displaystyle\leq 3\cdot 2^{-\operatorname{rank}(\xi_{r_{0}})}.
Proof.

For (8), use (4).

r=r0+1aweight(Ir,ηr)r=r0+1a2rank(ξr1)s=02(rank(ξr0)+s)=22rank(ξr0).\sum_{r=r_{0}+1}^{a}\operatorname{weight}(I_{r},\eta_{r})\leq\sum_{r=r_{0}+1}^{a}2^{-\operatorname{rank}(\xi_{r-1})}\leq\sum_{s=0}^{\infty}2^{-(\operatorname{rank}(\xi_{r_{0}})+s)}=2\cdot 2^{-\operatorname{rank}(\xi_{r_{0}})}.

Next, to prove (9), (5) is used.

r=r0aϱ(κ(ξr),κ(γ))\displaystyle\sum_{r=r_{0}}^{a}\varrho\big{(}\kappa(\xi_{r}),\kappa(\gamma)\big{)} =r=r0a1ϱ(κ(ξr),κ(γ))r=r0a1s=r+1aϱ(κ(ξs1),κ(ξs))\displaystyle=\sum_{r=r_{0}}^{a-1}\varrho\big{(}\kappa(\xi_{r}),\kappa(\gamma)\big{)}\leq\sum_{r=r_{0}}^{a-1}\sum_{s=r+1}^{a}\varrho\big{(}\kappa(\xi_{s-1}),\kappa(\xi_{s})\big{)}
(5)1mjr=r0a1s=r+1a2rank(ξs1)1mjr=r0a1s=02(rank(ξr)+s)\displaystyle\stackrel{{\scriptstyle\eqref{X_K evaluation analysis2}}}{{\leq}}\frac{1}{m_{j}}\sum_{r=r_{0}}^{a-1}\sum_{s=r+1}^{a}2^{-\operatorname{rank}(\xi_{s-1})}{\leq}\frac{1}{m_{j}}\sum_{r=r_{0}}^{a-1}\sum_{s=0}^{\infty}2^{-(\operatorname{rank}(\xi_{r})+s)}
2mjr=r0a12rank(ξr)4mj2rank(ξr0).\displaystyle\leq\frac{2}{m_{j}}\sum_{r=r_{0}}^{a-1}2^{-\operatorname{rank}(\xi_{r})}\leq\frac{4}{m_{j}}\cdot 2^{-\operatorname{rank}(\xi_{r_{0}})}.

Finally, to obtain (10), use (6).

r=r0+1aϱ(κ(ηr),κ(γ))\displaystyle\sum_{r=r_{0}+1}^{a}\varrho\big{(}\kappa(\eta_{r}),\kappa(\gamma)\big{)} r=r0+1aϱ(κ(ηr),κ(ξr))+r=r0+1aϱ(κ(ξr),κ(γ))\displaystyle\leq\sum_{r=r_{0}+1}^{a}\varrho\big{(}\kappa(\eta_{r}),\kappa(\xi_{r})\big{)}+\sum_{r=r_{0}+1}^{a}\varrho\big{(}\kappa(\xi_{r}),\kappa(\gamma)\big{)}
(6)&(9)r=r0+1a2rank(ξr1)+4mj2rank(ξr0+1)\displaystyle\!\!\!\!\!\!\stackrel{{\scriptstyle\eqref{X_K evaluation analysis3}\text{\&}\eqref{telescoping2}}}{{\leq}}\sum_{r=r_{0}+1}^{a}2^{-\operatorname{rank}(\xi_{r-1})}+\frac{4}{m_{j}}\cdot 2^{-\operatorname{rank}(\xi_{r_{0}+1})}
22rank(ξr0)+462rank(ξr0).\displaystyle\leq 2\cdot 2^{-\operatorname{rank}(\xi_{r_{0}})}+\frac{4}{6}\cdot 2^{-\operatorname{rank}(\xi_{r_{0}})}.

Proof of Proposition 4.1.

The second part follows from the first one and the fact that for any x𝔛C(K)x\in\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with x1\|x\|\leq 1,

P[N,)ϕ^(x)\displaystyle\|P_{[N,\infty)}\hat{\phi}(x)\| =supγΓ|(ϕ^(eγP[N,)))(x)|\displaystyle=\sup_{\gamma\in\Gamma}|(\hat{\phi}(e_{\gamma}^{*}\circ P_{[N,\infty)}))(x)|
supγΓlimm(|ϕ(κ(γ))|eγ(P[N,m)x)+7weight([N,m),γ)M)\displaystyle\leq\sup_{\gamma\in\Gamma}\lim_{m}\Big{(}\big{|}\phi\big{(}\kappa(\gamma)\big{)}\big{|}\Big{\|}e_{\gamma}^{*}\Big{(}P_{[N,m)}x\Big{)}\Big{\|}+7\operatorname{weight}([N,m),\gamma)M\Big{)}
supγΓ(MeγP[N,)+(7/8)M)4M.\displaystyle\leq\sup_{\gamma\in\Gamma}\Big{(}M\Big{\|}e_{\gamma}^{*}\circ P_{[N,\infty)}\Big{\|}+(7/8)M\Big{)}\leq 4M.

Choose NN\in\mathbb{N} such that (21(L/M)+14)/2N1/8(21(L/M)+14)/2^{N}\leq 1/8 (if M=0M=0 then there is nothing to prove). Observe that whenever weight(I,γ)=0\operatorname{weight}(I,\gamma)=0 then either eγPI=dγe_{\gamma}^{*}\circ P_{I}=d_{\gamma}^{*} or eγPI=0e_{\gamma}^{*}\circ P_{I}=0. In either case, ϕ^(eγPI)=ϕ(κ(γ))eγPI\hat{\phi}(e_{\gamma}^{*}\circ P_{I})=\phi(\kappa(\gamma))e_{\gamma}^{*}\circ P_{I} and in particular (7) holds. For the remaining cases, (7) is proved by induction on rank(γ)\operatorname{rank}(\gamma). The case rank(γ)=1\operatorname{rank}(\gamma)=1 is covered by the fact that all such γ\gamma have zero weight.

Assume now that (7) holds for all γ\gamma with rank(γ)n\operatorname{rank}(\gamma)\leq n. Let γΓ\gamma\in\Gamma with rank(γ)=n+1\operatorname{rank}(\gamma)=n+1 and let II be an interval of \mathbb{N}. If weight(I,γ)=0\operatorname{weight}(I,\gamma)=0 the desired conclusion holds. Assume therefore that weight(I,γ)=weight(γ)=mj1\operatorname{weight}(I,\gamma)=\operatorname{weight}(\gamma)=m_{j}^{-1}. Apply Proposition 3.7 and write

eγPI\displaystyle e_{\gamma}^{*}\circ P_{I} =r=1adξrPI+1mjr=1aλreηrPIrPI.\displaystyle=\sum_{r=1}^{a}d_{\xi_{r}}^{*}\circ P_{I}+\frac{1}{m_{j}}\sum_{r=1}^{a}\lambda_{r}e_{\eta_{r}}^{*}\circ P_{I_{r}}\circ P_{I}.
=rank(ξr)Idξr+1mjr=r0a0λreηrPIrI,\displaystyle=\sum_{\operatorname{rank}(\xi_{r})\in I}d_{\xi_{r}}^{*}+\frac{1}{m_{j}}\sum_{r=r_{0}}^{a_{0}}\lambda_{r}e_{\eta_{r}}^{*}\circ P_{I_{r}\cap I},

where r0=min{r:rank(ξr)min(I)}r_{0}=\min\{r:\operatorname{rank}(\xi_{r})\geq\min(I)\} (which is well defined because weight(I,γ)>0\operatorname{weight}(I,\gamma)>0) and a0=min{r:rank(ξr)=a or rank(ξr)max(I)}a_{0}=\min\{r:\operatorname{rank}(\xi_{r})=a\text{ or }\operatorname{rank}(\xi_{r})\geq\max(I)\}. Therefore,

(11) ϕ^(eγPI)ϕ(κ(γ))eγPI=rank(ξr)I(ϕ(κ(ξr))ϕ(κ(γ))dξr\displaystyle\Big{\|}\hat{\phi}\big{(}e^{*}_{\gamma}\circ P_{I}\big{)}-\phi\big{(}\kappa(\gamma)\big{)}e_{\gamma}^{*}\circ P_{I}\Big{\|}=\Big{\|}\sum_{\operatorname{rank}(\xi_{r})\in I}\Big{(}\phi\big{(}\kappa(\xi_{r})\big{)}-\phi\big{(}\kappa(\gamma\big{)}\Big{)}d_{\xi_{r}}^{*}
+1mjr=r0b0λr(ϕ^(eηrPIrI)ϕ(κ(γ))eηrPIrI)\displaystyle\phantom{=}+\frac{1}{m_{j}}\sum_{r=r_{0}}^{b_{0}}\lambda_{r}\Big{(}\hat{\phi}\big{(}e_{\eta_{r}}^{*}\circ P_{I_{r}\cap I}\big{)}-\phi\big{(}\kappa(\gamma)\big{)}e_{\eta_{r}}^{*}\circ P_{I_{r}\cap I}\Big{)}\Big{\|}
(12) Lrank(ξr)Iϱ(κ(ξr),κ(γ))dξr\displaystyle\leq L\sum_{\operatorname{rank}(\xi_{r})\in I}\varrho\big{(}\kappa(\xi_{r}),\kappa(\gamma)\big{)}\|d_{\xi_{r}}^{*}\|
(13) +1mjr=r0a0ϕ^(eηrPIrI)ϕ(κ(ηr))eηrPIrI\displaystyle\phantom{=}+\frac{1}{m_{j}}\sum_{r=r_{0}}^{a_{0}}\Big{\|}\hat{\phi}\big{(}e_{\eta_{r}}^{*}\circ P_{I_{r}\cap I}\big{)}-\phi\big{(}\kappa(\eta_{r})\big{)}e_{\eta_{r}}^{*}\circ P_{I_{r}\cap I}\Big{\|}
(14) +1mjr=r0a0|ϕ(κ(ηr))ϕ(κ(γ))|eηrPIrI.\displaystyle\phantom{=}+\frac{1}{m_{j}}\sum_{r=r_{0}}^{a_{0}}\Big{|}\phi\big{(}\kappa(\eta_{r})\big{)}-\phi\big{(}\kappa(\gamma)\big{)}\Big{|}\big{\|}e_{\eta_{r}}^{*}\circ P_{I_{r}\cap I}\big{\|}.

It is better to treat the three above terms separately. By (9),

(12)3L4mj2rank(ξr0)1mj12(L/M)2NM.\eqref{mess1}\leq 3L\frac{4}{m_{j}}2^{-\operatorname{rank}(\xi_{r_{0}})}\leq\frac{1}{m_{j}}\frac{12(L/M)}{2^{N}}M.

For the next term, use the inductive hypothesis to obtain

(13)\displaystyle\eqref{mess2} 1mjr=r0b07weight(IrI,ηr)M\displaystyle\leq\frac{1}{m_{j}}\sum_{r=r_{0}}^{b_{0}}7\operatorname{weight}(I_{r}\cap I,\eta_{r})M
(8)1mj7(weight(Ir0I,ηr0)+22rank(ξr0))M\displaystyle\stackrel{{\scriptstyle\eqref{telescoping1}}}{{\leq}}\frac{1}{m_{j}}7\Big{(}\operatorname{weight}(I_{r_{0}}\cap I,\eta_{r_{0}})+2\cdot 2^{\operatorname{rank}(\xi_{r_{0}})}\Big{)}M
1mj(78+142N)M.\displaystyle\leq\frac{1}{m_{j}}\Big{(}\frac{7}{8}+\frac{14}{2^{N}}\Big{)}M.

Evaluate the third term.

(14)\displaystyle\eqref{mess3} 1mj3(2M+Lr=r0+1a0ϱ(κ(ηr),κ(γ)))\displaystyle\leq\frac{1}{m_{j}}3\Big{(}2M+L\sum_{r=r_{0}+1}^{a_{0}}\varrho\big{(}\kappa(\eta_{r}),\kappa(\gamma)\big{)}\Big{)}
(10)1mj(6+9(L/M)2N)M\displaystyle\stackrel{{\scriptstyle\eqref{telescoping3}}}{{\leq}}\frac{1}{m_{j}}\Big{(}6+\frac{9(L/M)}{2^{N}}\Big{)}M

Combining the estimates for (12), (13), and (14) one obtains

(11)1mj(78+6+21(L/M)+142N)M7weight(γ)M.\eqref{want to bound}\leq\frac{1}{m_{j}}\Big{(}\frac{7}{8}+6+\frac{21(L/M)+14}{2^{N}}\Big{)}M\leq 7\operatorname{weight}(\gamma)M.

4.2. The embedding Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})

Denote by Lip(K)\mathrm{Lip}(K) the algebra of all Lipschitz function ϕ:K\phi:K\to\mathbb{C}. By the Stone-Weierstrass theorem, Lip(K)\mathrm{Lip}(K) is dense in C(K)C(K). Proposition 4.1 yields that the map ^:Lip(K)(𝔛C(K))\widehat{\,\cdot\,}:\mathrm{Lip}(K)\to\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is a well defined (but unbounded) linear homomorphism. Denote by []:(𝔛C(K))(𝔛C(K))/𝒦(𝔛C(K))=𝒞𝒶𝓁(𝔛𝒞(𝒦))[\,\cdot\,]:\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})\to\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})/\mathcal{K}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})=\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) the quotient map. The map Ψ:Lip(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:\mathrm{Lip}(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) given by Ψ(ϕ)=[ϕ^]\Psi(\phi)=[\hat{\phi}] is a bounded homomorphism and it is shown here that it is also an embedding. It is then shown that, by renorming 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, Ψ\Psi can be turned into a homomorphic isometric embedding.

Proposition 4.3.

The map Ψ:Lip(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:\mathrm{Lip}(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) extends to a homomorphic embedding Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}).

In fact, for every equivalent norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, the map Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦),||||||)\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}},{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}) is noncontractive.

Proof.

By Proposition 4.1, the map Ψ\Psi is well defined and Ψ4\|\Psi\|\leq 4. To verify the last statement, fix an equivalent norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} and ϕLip(K)\phi\in\mathrm{Lip}(K). Take κnKn\kappa\in\cup_{n}K_{n} and let Yκ={dγ:κ(γ)=κ}¯Y_{\kappa}=\overline{\langle\{d_{\gamma}:\kappa(\gamma)=\kappa\}\rangle}. This space is infinite dimensional; for nn sufficiently large γ=(n,κ)Δn0\gamma=(n,\kappa)\in\Delta^{0}_{n} and κ(γ)=κ\kappa(\gamma)=\kappa. By definition, ϕ^(x)=ϕ(κ)x\hat{\phi}(x)=\phi(\kappa)x, for all xYκx\in Y_{\kappa}. Therefore, for all A𝒦(𝔛C(K))A\in\mathcal{K}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}), |||(ϕ^A)|Yκ||||ϕ(κ)|{|\kern-1.07639pt|\kern-1.07639pt|(\hat{\phi}-A)|_{Y_{\kappa}}|\kern-1.07639pt|\kern-1.07639pt|}\geq|\phi(\kappa)|. Because nKn\cup_{n}K_{n} is dense in KK, |ϕ^A|ϕ{|\kern-1.07639pt|\kern-1.07639pt|\hat{\phi}-A|\kern-1.07639pt|\kern-1.07639pt|}\geq\|\phi\|_{\infty}. ∎

To turn Ψ\Psi into an isometric embedding the semigroups (n)n(\mathcal{B}_{n})_{n} are used.

Proposition 4.4.

There exists an equivalent norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} such that the map Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦),||||||)\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}},{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}) is a homomorphic isometry.

Proof.

By Proposition 4.3, it suffices to find a norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} that makes Ψ\Psi nonexpansive. Recall that, for nn\in\mathbb{N}, each ϕn\phi\in\mathcal{B}_{n} has norm at most one and is 2n2^{n}-Lipschitz. Let N(n)N(n) denote the minimum NN given by Proposition 4.1, for L=2nL=2^{n} and M=1M=1. This means that for each mnm\leq n\in\mathbb{N} and ϕm\phi\in\mathcal{B}_{m}, P[N(n),)ϕ^4\|P_{[N(n),\infty)}\hat{\phi}\|\leq 4. For each xx\in\mathbb{N} define x0=x\|x\|_{0}=\|x\|,

xn=sup{P[N(n),)ϕ^(x):ϕn}, for n,\|x\|_{n}=\sup\Big{\{}\big{\|}P_{[N(n),\infty)}\hat{\phi}(x)\big{\|}:\phi\in\mathcal{B}_{n}\Big{\}},\text{ for }n\in\mathbb{N},

and |x|=supn0xn{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=\sup_{n\geq 0}\|x\|_{n}. Then, x|x|4x\|x\|\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\leq 4\|x\|. Recall that 𝒜\mathcal{A} is a dense \mathbb{C}_{\mathbb{Q}}-linear subspace of C(K)C(K) and that ={ϕ𝒜:ϕ1}={ϕ1,ϕ2,}\mathcal{B}=\{\phi\in\mathcal{A}:\|\phi\|_{\infty}\leq 1\}=\{\phi_{1},\phi_{2},\ldots\}. To complete the proof, it suffices to fix ϕ\phi\in\mathcal{B} and show infA𝒦(𝔛C(K))|ϕ^A|1\inf_{A\in\mathcal{K}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})}{|\kern-1.07639pt|\kern-1.07639pt|\hat{\phi}-A|\kern-1.07639pt|\kern-1.07639pt|}\leq 1.

Let n0n_{0}\in\mathbb{N} such that ϕ=ϕn0\phi=\phi_{n_{0}}. For nn0n\geq n_{0} it will be now shown that |ϕ^P[1,N(n))ϕ^|n+1n{|\kern-1.07639pt|\kern-1.07639pt|\hat{\phi}-P_{[1,N(n))}\hat{\phi}|\kern-1.07639pt|\kern-1.07639pt|}\leq\frac{n+1}{n}. To that end, let x𝔛C(K)x\in\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with |x|1{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\leq 1. First, to compute ϕ^P[N(n),)x0\|\hat{\phi}P_{[N(n),\infty)}x\|_{0} note that nn+1ϕn\frac{n}{n+1}\phi\in\mathcal{B}_{n}. Therefore,

P[N(n),)ϕ^x0=n+1nP[N(n),)(nn+1ϕ)^xn+1nxnn+1n|x|.\|P_{[N(n),\infty)}\hat{\phi}x\|_{0}=\frac{n+1}{n}\Big{\|}P_{[N(n),\infty)}\widehat{\big{(}\textstyle{\frac{n}{n+1}}\phi\big{)}}x\Big{\|}\leq\frac{n+1}{n}\|x\|_{n}\leq\frac{n+1}{n}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}.

Next, take m{1,,n}m\in\{1,\ldots,n\} and ψm\psi\in\mathcal{B}_{m}. Then, ϕ^\hat{\phi}, ψ^\hat{\psi}, P[N(m),)P_{[N(m),\infty)}, P[N(m),)P_{[N(m),\infty)} all commute and nn+1ϕψn\frac{n}{n+1}\phi\psi\in\mathcal{B}_{n}. This yields

P[N(m),)ψ^(P[N(n),)ϕ^x)=n+1nP[N(n),)(nn+1ϕψ)^xn+1nxn.\Big{\|}P_{[N(m),\infty)}\hat{\psi}\Big{(}P_{[N(n),\infty)}\hat{\phi}x\Big{)}\Big{\|}=\frac{n+1}{n}\Big{\|}P_{[N(n),\infty)}\widehat{\big{(}\textstyle{\frac{n}{n+1}}\phi\psi\big{)}}x\Big{\|}\leq\frac{n+1}{n}\|x\|_{n}.

Therefore, P[N(n),)ϕ^xmn+1n|x|\|P_{[N(n),\infty)}\hat{\phi}x\|_{m}\leq\frac{n+1}{n}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}. Finally, take m>nm>n and ψm\psi\in\mathcal{B}_{m}. Then, nn+1ϕψm\frac{n}{n+1}\phi\psi\in\mathcal{B}_{m}. This yields,

P[N(m),)ψ^P[N(n),)ϕ^x=n+1nP[N(m),)(nn+1ϕψ)^xn+1nxm,\Big{\|}P_{[N(m),\infty)}\hat{\psi}P_{[N(n),\infty)}\hat{\phi}x\Big{\|}=\frac{n+1}{n}\Big{\|}P_{[N(m),\infty)}\widehat{\big{(}\textstyle{\frac{n}{n+1}}\phi\psi\big{)}}x\Big{\|}\leq\frac{n+1}{n}\|x\|_{m},

i.e., P[N(n),)ϕ^xmn+1n|x|\|P_{[N(n),\infty)}\hat{\phi}x\|_{m}\leq\frac{n+1}{n}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|x\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}. ∎

Remark 4.5.

Unless KK is finite, it is not true that for every continuous function ϕ:K\phi:K\to\mathbb{C} the linear map ϕ^:{dγ:γΓ}{dγ:γΓ}\hat{\phi}:\langle\{d_{\gamma}:\gamma\in\Gamma\}\rangle\to\langle\{d_{\gamma}:\gamma\in\Gamma\}\rangle extends to a bounded linear operator on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. If this were the case then, by the closed graph theorem, ^:C(K)(𝔛C(K))\widehat{\cdot}:C(K)\to\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) would be a homomorphic embedding. By Proposition 8.6 (i) this is impossible.

5. The impact of the conditional structure of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

This relatively brief section discusses the outcomes of the conditional structure of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} (the properties derived from the definition of odd-weight members γ\gamma of Γ\Gamma). These outcomes are presented in the form of two black-box Theorems that can be used to directly prove that the embedding Ψ:C(K)𝒞𝒶𝓁(𝔛𝒞(𝒦))\Psi:C(K)\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is onto. The proofs of these Theorems are based on HI techniques and are included in Sections 6 and 7.

Theorem 5.1.

Let T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} be a bounded linear operator. Then, for every ε>0\varepsilon>0 there exist nn\in\mathbb{N} and δ>0\delta>0 such that for all γ\gamma,γΓ\gamma^{\prime}\in\Gamma with min{rank(γ),rank(γ)}n\min\{\operatorname{rank}(\gamma),\operatorname{rank}(\gamma^{\prime})\}\geq n and ϱ(κ(γ),κ(γ))<δ\varrho(\kappa(\gamma),\kappa(\gamma^{\prime}))<\delta,

(15) |dγ(Tdγ)dγ(Tdγ)|<ε.\big{|}d_{\gamma}^{*}\big{(}Td_{\gamma}\big{)}-d_{\gamma^{\prime}}^{*}\big{(}Td_{\gamma^{\prime}}\big{)}\big{|}<\varepsilon.

Therefore, the function φT:K\varphi_{T}:K\to\mathbb{C} given by

φT(κ)=limrank(γ)κ(γ)κdγ(Tdγ)\varphi_{T}(\kappa)=\lim_{\begin{subarray}{c}\operatorname{rank}(\gamma)\to\infty\\ \kappa(\gamma)\to\kappa\end{subarray}}d_{\gamma}^{*}\big{(}Td_{\gamma}\big{)}

is well defined and continuous.

Property (15) can be seen as an eventual continuity of the diagonal entries (dγ(Tdγ))γΓ(d_{\gamma}^{*}(Td_{\gamma}))_{\gamma\in\Gamma} of TT. Its proof goes deeply into the details of HI techniques. The fact that φT\varphi_{T} is a well defined continuous function is a straightforward consequence of the density of nKn\cup_{n}K_{n} and some elementary real analysis. It then immediately follows that the linear map Φ~:(𝔛C(K))C(K)\widetilde{\Phi}:\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})\to C(K), given by Φ~(T)=φT\widetilde{\Phi}(T)=\varphi_{T} is a bounded linear operator. Indeed, for any T(𝔛C(K))T\in\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) and γΓ\gamma\in\Gamma, |dγ(Tdγ)|dγdγT6T|d_{\gamma}^{*}(Td_{\gamma})|\leq\|d_{\gamma}^{*}\|\|d_{\gamma}\|\|T\|\leq 6\|T\|. The next statement yields the remaining necessary information to complete the proof of Theorem A.

Theorem 5.2.

A bounded linear operator T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is compact if and only if φT=0\varphi_{T}=0.

An immediate consequence is that the map Φ:𝒞𝒶𝓁(𝔛𝒞(𝒦))𝒞(𝒦)\Phi:\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})\to C(K), defined via the formula Φ([T])=φT\Phi([T])=\varphi_{T}, is a bounded linear injection. Indeed, Φ=Φ~/kerΦ~\Phi=\widetilde{\Phi}/\mathrm{ker}\widetilde{\Phi}. Now, an almost straightforward computation (together with Proposition 4.4) yields the main result of this paper.

Corollary 5.3 (Theorem A (b) & (c)).

The map Φ:𝒞𝒶𝓁(𝔛𝒞(𝒦))𝒞(𝒦)\Phi:\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})\to C(K) is the inverse of Ψ:C(K)𝒞𝒶𝓁(𝒳)\Psi:C(K)\to\mathpzc{Cal}(X). Therefore, with an appropriate equivalent norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}, the Calkin algebra of (𝔛C(K),||||||)(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}},{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}) is homomorphically isometric to C(K)C(K).

Proof.

Since both Φ\Phi and Ψ\Psi are injections, proving that ΦΨ:C(K)C(K)\Phi\circ\Psi:C(K)\to C(K) is the identity map yields the conclusion. By density, it is sufficient to show that for ϕLip(K)\phi\in\mathrm{Lip}(K), Φ(Ψϕ)=ϕ\Phi(\Psi\phi)=\phi. Recall that ϕ^:𝔛C(K)𝔛C(K)\hat{\phi}:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is a well defined bounded linear operator and for all γΓ\gamma\in\Gamma,

φϕ^(κ)=limrank(γ)κ(γ)κdγ(ϕ^(dγ))=limrank(γ)κ(γ)κϕ(κ(γ))=ϕ(κ).\varphi_{\widehat{\phi}}(\kappa)=\lim_{\begin{subarray}{c}\operatorname{rank}(\gamma)\to\infty\\ \kappa(\gamma)\to\kappa\end{subarray}}d_{\gamma}^{*}\big{(}\hat{\phi}(d_{\gamma})\big{)}=\lim_{\begin{subarray}{c}\operatorname{rank}(\gamma)\to\infty\\ \kappa(\gamma)\to\kappa\end{subarray}}\phi\big{(}\kappa(\gamma)\big{)}=\phi(\kappa).

In other words, Φ(Ψϕ)=Φ([ϕ^])=Φ~(ϕ^)=φϕ^=ϕ\Phi(\Psi\phi)=\Phi([\hat{\phi}])=\widetilde{\Phi}(\hat{\phi})=\varphi_{\widehat{\phi}}=\phi. ∎

An interesting corollary is that the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} has the diagonal-plus-compact property.

Corollary 5.4 (Theorem A (a)).

Every bounded linear operator T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is of the form D+AD+A, where DD is diagonal bounded linear operator and AA is a compact linear operator.

Proof.

Let T(𝔛C(K))T\in\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) and pick an absolutely summable sequence (in the \|\cdot\|_{\infty}-norm) of Lipschitz functions ϕn:K\phi_{n}:K\to\mathbb{C} with n=1ϕn=φT\sum_{n=1}^{\infty}\phi_{n}=\varphi_{T}. Using Proposition 4.1, for each nn\in\mathbb{N}, pick N(n)N(n) such that P[N(n),)ϕ^n4ϕn\|P_{[N(n),\infty)}\hat{\phi}_{n}\|\leq 4\|\phi_{n}\|_{\infty}. Then, D=n=1P[N(n),)ϕ^nD=\sum_{n=1}^{\infty}P_{[N(n),\infty)}\hat{\phi}_{n} is a well defined diagonal bounded linear operator and

φD=Φ~(D)\displaystyle\varphi_{D}=\widetilde{\Phi}(D) =n=1Φ~(P[N(n),)ϕ^n)=n=1Φ~(ϕ^n)=n=1ϕn=φT.\displaystyle=\sum_{n=1}^{\infty}\widetilde{\Phi}\big{(}P_{[N(n),\infty)}\hat{\phi}_{n}\big{)}=\sum_{n=1}^{\infty}\widetilde{\Phi}\big{(}\hat{\phi}_{n}\big{)}=\sum_{n=1}^{\infty}\phi_{n}=\varphi_{T}.

This means φTD=φTφD=0\varphi_{T-D}=\varphi_{T}-\varphi_{D}=0, i.e., A=TDA=T-D is compact. ∎

Remark 5.5.

The expert reader will find the following fact interesting. Given a T(𝔛C(K))T\in\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}), Theorem 5.1 provides an explicit diagonal operator DD so that TDT-D is compact. Therefore, Kakutani’s fixed point theorem ([21]) is not required to prove non-constructively that such a DD must exist. This is a crucial fact in this paper as it allows the present method to work despite omitting convex combinations in the definition of Γ\Gamma. This fixed point theorem was used by Gowers and Maurey in [18] and by Tarbard in [44].

6. Common concepts from HI methods

This section goes through the ubiquitous notions of rapidly increasing sequences, exact pairs, and dependent sequences. These are specialized vectors and sequences of vectors that have been used in almost all HI and related constructions. In particular, they were also used in [4]. Estimates of the norms of linear combinations of such objects are fundamental in the study of the geometry of these spaces and of their bounded linear operators. These estimates are dependent on notions such as an auxiliary space and a basic inequality, originating in [1] but found in [4] as well. The versions of the proofs found in [4] work in the current setting as well. Rather than repeating pages of identical arguments, it was chosen to refer to that paper while only highlighting the minor differences.

Henceforth, for a vector x𝔛C(K)x\in\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} the support of xx is the set supp(x)={n:P{n}x0}\operatorname{supp}(x)=\{n\in\mathbb{N}:P_{\{n\}}x\neq 0\}. Similarly, the range of xx is the smallest interval range(x)\operatorname{range}(x) of \mathbb{N} containing supp(x)\operatorname{supp}(x). A bock sequence (xk)(x_{k}) in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is one for which maxsupp(xk)<minsupp(xk+1)\max\operatorname{supp}(x_{k})<\min\operatorname{supp}(x_{k+1}), for all kk.

6.1. Rapidly increasing sequences

These sequences are used, among other things, to identify “small” operators in HI-type constructions. Here, as in [4], they are used to characterize compact operators.

Definition 6.1.

Let C>0C>0. A block sequence (xk)kI(x_{k})_{k\in I} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, indexed over an interval II of \mathbb{N}, will be called a CC-rapidly increasing sequence (or CC-RIS) if there exists an increasing sequence (jk)kI(j_{k})_{k\in I} in \mathbb{N} such that the following hold for kIk\in I:

  1. (i)

    xkC\|x_{k}\|\leq C,

  2. (ii)

    if k>min(I)k>\min(I), jk1<minsupp(xk)j_{k-1}<\min\operatorname{supp}(x_{k}), and

  3. (iii)

    for γΓ\gamma\in\Gamma with weight(γ)=mi1>mjk1\operatorname{weight}(\gamma)=m_{i}^{-1}>m_{j_{k}}^{-1}, |eγ(xk)|Cmi1|e_{\gamma}^{*}(x_{k})|\leq Cm^{-1}_{i}.

Suppressing the constant CC, (xk)kI(x_{k})_{k\in I} will be called a RIS.

Example 6.2.

For every sequence (γk)k(\gamma_{k})_{k} in Γ\Gamma with rank(γk)\operatorname{rank}(\gamma_{k})\to\infty, (dγk)k(d_{\gamma_{k}})_{k} has a subsequence that is s RIS.

Indeed, for any γ\gamma, γΓ\gamma^{\prime}\in\Gamma with weight(γ)=mi1weight(γ)\operatorname{weight}(\gamma^{\prime})=m_{i}^{-1}\neq\operatorname{weight}(\gamma), Proposition 3.7 yields that, unless eγ(dγ)=0e_{\gamma^{\prime}}^{*}(d_{\gamma})=0, there exist λ𝔻\lambda\in\mathbb{D} and ηΓ\eta\in\Gamma such that eγ(dγ)=mi1λeη(dγ)e_{\gamma^{\prime}}^{*}(d_{\gamma})=m_{i}^{-1}\lambda e_{\eta}^{*}(d_{\gamma}), i.e., |eγ(dγ)|2mi1|e_{\gamma^{\prime}}^{*}(d_{\gamma})|\leq 2m_{i}^{-1}. If limkweight(γk)=0\lim_{k}\operatorname{weight}(\gamma_{k})=0, then it is not difficult to pick a subsequence of (dγk)k(d_{\gamma_{k}})_{k} that is a 22-RIS. Otherwise, if there is i0i_{0} so that for infinitely many kk\in\mathbb{N}, weight(γk)=mi01\operatorname{weight}(\gamma_{k})=m_{i_{0}}^{-1}, then (dγk)k(d_{\gamma_{k}})_{k} has subsequence that is a (2mi0)(2m_{i_{0}})-RIS.

Remark 6.3.

If (xk)k(x_{k})_{k} and (yk)k(y_{k})_{k} are RISs and (λk)k(\lambda_{k})_{k}, (μk)k(\mu_{k})_{k} are bounded sequences of scalars, then (λkxk+μkyk)k(\lambda_{k}x_{k}+\mu_{k}y_{k})_{k} has a subsequence that is a RIS.

Proposition 6.4.

Let YY be a Banach space and T:𝔛C(K)YT:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to Y be a bounded linear operator. If limkTxk=0\lim_{k}\|Tx_{k}\|=0 for every RIS (xk)k(x_{k})_{k} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} then limkTxk=0\lim_{k}\|Tx_{k}\|=0 for any bounded block sequence in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

Proposition 6.5.

The basis ((dγ)γΔn)n=1((d_{\gamma})_{\gamma\in\Delta_{n}})_{n=1}^{\infty} of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is shrinking and in particular 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{*} is isomorphic to 1\ell_{1}.

Comment on proof.

These are [4, Proposition 5.11 and Proposition 5.12, page 27]. The proofs are fundamental to the Argyros-Haydon construction and highly non-trivial. That being said, they translate almost verbatim to this paper. The unconvinced reader may retrace [4, pages 19-28] (which also use [4, Lemma 2.4, page 5]) with having only two things in mind.

In [4] the evaluation analysis of each γ\gamma contains components of the form brP(s,)b_{r}^{*}\circ P_{(s,\infty)}, where brb_{r}^{*} is a convex combination of certain eηe_{\eta}^{*}. Here, in the same place these components are of the form eηPIre_{\eta}^{*}\circ P_{I_{r}}, where IrI_{r} is a bounded interval of \mathbb{N}. This has the consequence that in several places intervals of the form (s,)(s,\infty) need to be replaced with bounded intervals II. This does not cause any change in constants, because in [4], P(s,)3\|P_{(s,\infty)}\|\leq 3 whereas here PI3\|P_{I}\|\leq 3 (see Remark 3.3). This is due to the present choice m18m_{1}\geq 8.

The second detail, is that in [4] there exists a unique γ0\gamma_{0} of weight zero, namely the member of Δ1\Delta_{1}. Here, there are infinitely many γ\gamma in Γ\Gamma of weight zero. By their nature, these can be treated exactly as γ0\gamma_{0} of [4]. This comes up in the proof of [4, Proposition 5.4 (Basic Inequality), page 20]. ∎

Recall that in a space with a shrinking basis, all bounded block sequences are weakly null and thus a compact operator maps them to norm-null sequences. Therefore, the above two propositions immediately imply the next.

Corollary 6.6.

A bounded linear operator T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is compact if and only if for every RIS (xk)k(x_{k})_{k} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, limkTxk=0\lim_{k}\|Tx_{k}\|=0.

6.2. Exact pairs and dependent sequences

These are highly specialized sequences of vectors that use RISs as their building blocks. Here, as in [4], their main purpose is to extract the compact part of a bounded linear operator. In this section the definition of these objects is recalled and it is reminded how they can be constructed from RISs. Estimates of the norm of their linear combinations are also given.

Definition 6.7.

Let C>0C>0, jj\in\mathbb{N}, and ε{0,1}\varepsilon\in\{0,1\}. A pair (x,η)𝔛C(K)×Γ(x,\eta)\in\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\times\Gamma is called a (C,j,ε)(C,j,\varepsilon)-exact pair if

  1. (i)

    for all ξΓ\xi\in\Gamma, |dξ(x)|Cmj1|d_{\xi}^{*}(x)|\leq Cm^{-1}_{j},

  2. (ii)

    weight(η)=mj1\operatorname{weight}(\eta)=m_{j}^{-1},

  3. (iii)

    xC\|x\|\leq C and eη(x)=εe_{\eta}^{*}(x)=\varepsilon, and

  4. (iv)

    for every ηΓ\eta^{\prime}\in\Gamma with weight(η)=mi1mj1\operatorname{weight}(\eta^{\prime})=m_{i}^{-1}\neq m_{j}^{-1},

    |eη(x)|{Cmi1if mi1>mj1,Cmj1if mi1<mj1.|e_{\eta^{\prime}}^{*}(x)|\leq\left\{\begin{array}[]{ll}Cm_{i}^{-1}&\mbox{if }m_{i}^{-1}>m_{j}^{-1},\\ Cm_{j}^{-1}&\mbox{if }m_{i}^{-1}<m_{j}^{-1}.\end{array}\right.

The next lemma explains how to construct (C,0)(C,0)-exact pairs in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, which are necessary in the study of operators. On contrast, (C,1)(C,1)-exact pairs are used to study the geometry of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, e.g., to prove that 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} contains no unconditional sequences (see Section 8). The construction of (C,0)(C,0)-exact pairs is done very similarly as in [4] and by applying Proposition 3.8.

A skipped block sequence (xk)k(x_{k})_{k} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is one for which maxsupp(xk)+1<minsupp(xk+1)\max\operatorname{supp}(x_{k})+1<\min\operatorname{supp}(x_{k+1}), for all kk.

Lemma 6.8.

Let jj\in\mathbb{N}, C>0C>0, (xk)k=1n2j(x_{k})_{k=1}^{n_{2j}} be a skipped block CC-RIS, (qk)k=1n2j(q_{k})_{k=1}^{n_{2j}} be a sequence of natural numbers, (Ik)k=1n2j(I_{k})_{k=1}^{n_{2j}} be finite intervals of \mathbb{N}, ηkΓ\eta_{k}\in\Gamma and λk𝔻qk\lambda_{k}\in\mathbb{D}_{q_{k}}, for k=1,,n2jk=1,\ldots,n_{2j}, and κ0K\kappa_{0}\in K. Assume that the following are satisfied.

  1. (i)

    2jq12j\leq q_{1}, supp(x1)I1[1,q1)\operatorname{supp}(x_{1})\cup I_{1}\subset[1,q_{1}), and supp(xk)Ik(qk1,qk)\operatorname{supp}(x_{k})\cup I_{k}\subset(q_{k-1},q_{k}) , for k=2,,n2jk=2,\ldots,n_{2j}.

  2. (ii)

    rank(ηk)<qk\operatorname{rank}(\eta_{k})<q_{k}, for k=1,,n2jk=1,\ldots,n_{2j},

  3. (iii)

    weight(Ik,ηk)2qk1\operatorname{weight}(I_{k},\eta_{k})\leq 2^{-q_{k-1}}, for k=2,,n2jk=2,\ldots,n_{2j}.

  4. (iv)

    ϱ(κ(ηk),κ0)2(qk1+1)\varrho(\kappa(\eta_{k}),\kappa_{0})\leq 2^{-(q_{k-1}+1)}, for k=2,,n2jk=2,\ldots,n_{2j}.

  5. (v)

    eηk(PIkxk)=0e_{\eta_{k}}^{*}(P_{I_{k}}x_{k})=0, for k=1,,n2jk=1,\ldots,n_{2j}.

Denote

z=m2jn2jk=1n2jxk.z=\frac{m_{2j}}{n_{2j}}\sum_{k=1}^{n_{2j}}x_{k}.

Then, there exists γΓ\gamma\in\Gamma with rank(γ)=qn2j\operatorname{rank}(\gamma)=q_{n_{2j}}, weight(γ)=m2j1\operatorname{weight}(\gamma)=m_{2j}^{-1}, and ϱ(κ(γ),κ0)2qn2j\varrho(\kappa(\gamma),\kappa_{0})\leq 2^{-q_{n_{2j}}} such that (z,γ)(z,\gamma) is a (16C,2j,0)(16C,2j,0)-exact pair. Furthermore, there exist ξkΔqk\xi_{k}\in\Delta_{q_{k}}, for k=1,,n2jk=1,\ldots,n_{2j}, such that eγe_{\gamma}^{*} has an evaluation analysis

eγ=k=1n2jdξk+1m2jk=1n2jλkeηkPIk.e_{\gamma}^{*}=\sum_{k=1}^{n_{2j}}d_{\xi_{k}}^{*}+\frac{1}{m_{2j}}\sum_{k=1}^{n_{2j}}\lambda_{k}e_{\eta_{k}}^{*}\circ P_{I_{k}}.
Proof.

Apply Proposition 3.8 to find γ\gamma with the desired weight, proximity to κ0\kappa_{0}, and evaluation analysis. The proof that (z,γ)(z,\gamma) is a (16C,2j,0)(16C,2j,0)-exact pair is identical to the proof of [4, Lemma 6.2, page 29]. ∎

Dependent sequences comprise exact pairs. They are chosen inductively with the help of the coding function σ\sigma.

Definition 6.9.

Let C>0C>0, j0j_{0}\in\mathbb{N}, and ε{0,1}\varepsilon\in\{0,1\}. A sequence (xi)i=1n2j01(x_{i})_{i=1}^{n_{2j_{0}}-1} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is called a (C,2j01,ε)(C,2j_{0}-1,\varepsilon)-dependent sequence if there exists a γΓ\gamma\in\Gamma with weight(γ)=m2j01\operatorname{weight}(\gamma)=m_{2j_{0}}^{-1} and evaluation analysis

eγ=i=1n2j01dξi+1m2j01i=1n2j01eηiPIie_{\gamma}^{*}=\sum_{i=1}^{n_{2j_{0}-1}}d_{\xi_{i}}^{*}+\frac{1}{m_{2j_{0}-1}}\sum_{i=1}^{n_{2j_{0}-1}}e_{\eta_{i}}^{*}\circ P_{I_{i}}

such that range(xi)Ii\operatorname{range}(x_{i})\subset I_{i}, for i=1,,n2j01i=1,\ldots,n_{2j_{0}-1}, denoting weight(η1)=m4j12\operatorname{weight}(\eta_{1})=m_{4j_{1}-2} then (x1,η1)(x_{1},\eta_{1}) is a (C,4j12,ε)(C,4j_{1}-2,\varepsilon)-exact pair, and denoting weight(ηi)=m4ji\operatorname{weight}(\eta_{i})=m_{4j_{i}}, for i=2,,n2j01i=2,\ldots,n_{2j_{0}-1} then (xi,γi)(x_{i},\gamma_{i}) is a (C,4ji,ε)(C,4j_{i},\varepsilon)-exact pair, for i=2,,n2j01i=2,\ldots,n_{2j_{0}-1}.

Note that in the construction of γ\gamma, after each step ii, ηi+1{\eta_{i+1}} is only allowed to have the weight m4σ(ξi)1m^{-1}_{4\sigma(\xi_{i})}. Therefore, the exact pair (xi+1,ηi+1)(x_{i+1},\eta_{i+1}) needs to be built after ξi\xi_{i}.

It is straightforward to check that the average of the terms of a (C,2j01,1)(C,2j_{0}-1,1)-dependent sequence have norm at least m2j011m_{2j_{0}-1}^{-1}. However, for (C,2j01,0)(C,2j_{0}-1,0)-dependent sequence the outcome is much smaller.

Proposition 6.10.

Let C>0C>0, j0j_{0}\in\mathbb{N}, and (xi)i=1n2j01(x_{i})_{i=1}^{n_{2j_{0}}-1} be a (C,2j01,0)(C,2j_{0}-1,0)-dependent sequence in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. Then,

n2j011i=1n2j01xi30Cm2j012.\Big{\|}n_{2j_{0}-1}^{-1}\sum_{i=1}^{n_{2j_{0}-1}}x_{i}\Big{\|}\leq 30Cm_{2j_{0}-1}^{-2}.
Proof.

This is proved identically to [4, Proposition 6.6, pages 30-31]. ∎

7. Bounded linear operators on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

In this section the common HI concepts are combined with the weight and metric constraints to prove Theorem 5.1 and Theorem 5.2.

7.1. Non-vanishing estimates of very fast growing sequences

The following states that any block sequence whose norm is bounded from below admits non-vanishing estimates by functions of the form eγPIe^{*}_{\gamma}\circ P_{I} with weight(I,γ)\operatorname{weight}(I,\gamma) tending to zero. This is necessary to be able to construct zero dependent sequences. This process has its roots in [7].

Proposition 7.1.

Any block sequence (xk)k(x_{k})_{k} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with lim infkxk>0\liminf_{k}\|x_{k}\|>0 has a subseqeuence, again denoted by (xk)k(x_{k})_{k}, with the following property. For each kk\in\mathbb{N} there exist an interval IkI_{k} or range(xk)\operatorname{range}(x_{k}) and ηkΓ\eta_{k}\in\Gamma such that

limkweight(Ik,ηk)=0 and lim infk|eηk(PIkxk)|>0.\lim_{k}\operatorname{weight}(I_{k},\eta_{k})=0\text{ and }\liminf_{k}\big{|}e^{*}_{\eta_{k}}\big{(}P_{I_{k}}x_{k}\big{)}\big{|}>0.

The following lemma is the main quantitative argument required in the proof of the above proposition.

Lemma 7.2.

Let x𝔛C(K)x\in\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} and γΓ\gamma\in\Gamma with weight(γ)=mj1\operatorname{weight}(\gamma)=m_{j}^{-1} such that |eγ(x)|(3/4)x|e_{\gamma}^{*}(x)|\geq(3/4)\|x\|. Then there exist ηΓ\eta\in\Gamma and a finite interval II of range(x)\operatorname{range}(x) with

weight(I,η)2minsupp(x) and |eη(PIx)|18njx.\operatorname{weight}(I,\eta)\leq 2^{-\min\operatorname{supp}(x)}\text{ and }\big{|}e_{\eta}^{*}(P_{I}x)\big{|}\geq\frac{1}{8n_{j}}\|x\|.
Proof.

Apply Proposition 3.7 to write

eγ\displaystyle e_{\gamma}^{*} =r=1adξr+1mjr=1aλreηrPIr.\displaystyle=\sum_{r=1}^{a}d_{\xi_{r}}^{*}+\frac{1}{m_{j}}\sum_{r=1}^{a}\lambda_{r}e_{\eta_{r}}^{*}\circ P_{I_{r}}.

Put J=range(x)J=\operatorname{range}(x), r0=min{r:rank(ξr)min(J)}r_{0}=\min\{r:\operatorname{rank}(\xi_{r})\geq\min(J)\} and for r=r0,,ar=r_{0},\ldots,a, Ir=IrJI_{r}^{\prime}=I_{r}\cap J. Recall that anja\leq n_{j}. Assuming that the conclusion is false,

34x\displaystyle\frac{3}{4}\|x\| |eγ(x)|=|r=1aeξr(P{rank(ξr)}x)+1mjr=r0aeηr(PIrx)|\displaystyle\leq|e_{\gamma}^{*}(x)|=\Big{|}\sum_{r=1}^{a}e_{\xi_{r}}^{*}(P_{\{\operatorname{rank}(\xi_{r})\}}x)+\frac{1}{m_{j}}\sum_{r=r_{0}}^{a}e_{\eta_{r}}^{*}\big{(}P_{I^{\prime}_{r}}x\big{)}\Big{|}
1mj|eηr0(PIr0x)|+r=1a|eξr(P{rank(ξr)}x)|+1mjr=r0+1a|eηr(PIrx)|.\displaystyle\leq\frac{1}{m_{j}}\big{|}e_{\eta_{r_{0}}}^{*}\big{(}P_{I^{\prime}_{r_{0}}}x\big{)}\big{|}+\sum_{r=1}^{a}\big{|}e_{\xi_{r}}^{*}(P_{\{\operatorname{rank}(\xi_{r})\}}x)\big{|}+\frac{1}{m_{j}}\sum_{r=r_{0}+1}^{a}\big{|}e_{\eta_{r}}^{*}\big{(}P_{I^{\prime}_{r}}x\big{)}\big{|}.

By definition, for r=r0,,ar=r_{0},\ldots,a, weight({rank(ξr)},ξr)=0\operatorname{weight}(\{\operatorname{rank}(\xi_{r})\},\xi_{r})=0. Furthermore, the growth condition on weights in the sets Δn(b)\Delta_{n}^{\text{\ref{age nonzero}}} yields that for r=r0+1,,ar=r_{0}+1,\ldots,a, weight(Ir,ηr)2rankξr02minsupp(x)\operatorname{weight}(I_{r},\eta_{r})\leq 2^{-\operatorname{rank}{\xi_{r_{0}}}}\leq 2^{-\min\operatorname{supp}(x)}. If ones supposes that the conclusion fails, it would follow that

34x<4xmj+nj18njx+njmj18njx(41m1+14)x,\frac{3}{4}\|x\|<\frac{4\|x\|}{m_{j}}+n_{j}\frac{1}{8n_{j}}\|x\|+\frac{n_{j}}{m_{j}}\frac{1}{8n_{j}}\|x\|\leq\Big{(}4\frac{1}{m_{1}}+\frac{1}{4}\Big{)}\|x\|,

i.e., m1<8m_{1}<8 which is absurd. ∎

Proof of Proposition 7.1.

For each kk\in\mathbb{N}, pick γkΓ\gamma_{k}\in\Gamma with |eγk(xk)|(3/4)xk|e_{\gamma_{k}}^{*}(x_{k})|\geq(3/4)\|x_{k}\|. Distinguish two cases. If limkweight(range(xk),γk)=0\lim_{k}\operatorname{weight}(\operatorname{range}(x_{k}),\gamma_{k})=0 then the proof is complete. Note that this also includes the case in which, for all kk\in\mathbb{N}, minrange(xk)=rank(γk)\min\operatorname{range}(x_{k})=\operatorname{rank}(\gamma_{k}) and thus weight(range(xk),γk)=0\operatorname{weight}(\operatorname{range}(x_{k}),\gamma_{k})=0. Otherwise, by passing to an infinite subsequence and relabeling, there exists j0j_{0}\in\mathbb{N} so that for all kk\in\mathbb{N}, weight(γk)=mj01\operatorname{weight}(\gamma_{k})=m^{-1}_{j_{0}}. Apply Lemma 7.2 to find for each kk\in\mathbb{N} a finite interval IkI_{k} of range(xk)\operatorname{range}(x_{k}) and ηkΓ\eta_{k}\in\Gamma with weight(Ik,ηk)2minsupp(xk)\operatorname{weight}(I_{k},\eta_{k})\leq 2^{-\min\operatorname{supp}(x_{k})} and |eηk(PIkxk)|1/(8nj0)|e_{\eta_{k}}^{*}(P_{I_{k}}x_{k})|\geq 1/(8n_{j_{0}}). ∎

7.2. Linear transformations of rapidly increasing sequences

This section deals with controlling the action of bounded linear operators on RISs. In [4] it was shown that for a bounded linear operator TT and a RIS (xk)k(x_{k})_{k}, dist(Txk,xk)0\mathrm{dist}(Tx_{k},\mathbb{C}x_{k})\to 0. Here, this is not true (unless KK is a singleton). If it were, 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} would have the scalar-plus-compact property. Instead, it is shown that TT is in a certain weak sense close to a diagonal operator. Recall that eventually it can be shown that TT is a compact perturbation of such an operator (Corollary 5.4).

Proposition 7.3.

Let T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} be a linear operator, (xk)k(x_{k})_{k} be a RIS in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, and (Ik,ηk)k(I_{k},\eta_{k})_{k} be a sequence, whose each term is a pair of an interval IkI_{k} of \mathbb{N} with an ηk\eta_{k} in Γ\Gamma, such that limkweight(Ik,ηk)=0\lim_{k}\operatorname{weight}(I_{k},\eta_{k})=0. If

eηk(PIkxk)=0, for all k, and lim infk|eηk(PIkTxk)|>0,e_{\eta_{k}}^{*}\big{(}P_{I_{k}}x_{k}\big{)}=0,\text{ for all }k\in\mathbb{N},\text{ and }\liminf_{k}\big{|}e_{\eta_{k}}^{*}\big{(}P_{I_{k}}Tx_{k}\big{)}\big{|}>0,

then TT is unbounded.

Proof.

Let (xk)k(x_{k})_{k} be a CC-RIS and assume that TT is bounded. Recall that the basis of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is shrinking and therefore (Txk)k(Tx_{k})_{k} is weakly null. By applying a compact perturbation to TT and passing to a subsequence, it may be assumed that (Txk)k(Tx_{k})_{k} is a block sequence and that for some ε>0\varepsilon>0 and each kk\in\mathbb{N}, eγk(PIkTxk)=|eγk(PIkTxk)|εe_{\gamma_{k}}^{*}(P_{I_{k}}Tx_{k})=|e_{\gamma_{k}}^{*}(P_{I_{k}}Tx_{k})|\geq\varepsilon (after perhaps applying a complex rotation to each xkx_{k}). By restricting each IkI_{k} it may be assumed that it is a subset of the smallest interval JkJ_{k} containing range(xk)range(Txk)\operatorname{range}(x_{k})\cup\operatorname{range}(Tx_{k}). Using the compactness of KK, it may also be assumed that there is κ0K\kappa_{0}\in K with limkϱ(κ(ηk),κ0)=0\lim_{k}\varrho(\kappa(\eta_{k}),\kappa_{0})=0. With this information, in two steps it is possible to construct a vector that “blows up” the norm of TT.

Step 1: For each j,qj,q\in\mathbb{N} and δ>0\delta>0, there exists a (16C,2j,0)(16C,2j,0)-exact pair (z,ζ)(z,\zeta) such that minsupp(z)>q\min\operatorname{supp}(z)>q, ϱ(κ(η),κ0)δ\varrho(\kappa(\eta),\kappa_{0})\leq\delta, and eζ(Tz)=|eζ(Tz)|εe_{\zeta}^{*}(Tz)=|e_{\zeta}^{*}(Tz)|\geq\varepsilon.

This is achieved with the help of Lemma 6.8. Choose members of the sequence xk1,,xkn2jx_{k_{1}},\ldots,x_{k_{n_{2}j}}, starting after qq, and q1<<qn2jq_{1}<\cdots<q_{n_{2j}} such that the assumptions of that Lemma are satisfied, while at the same time I1J1[1,q1)I_{1}\subset J_{1}\subset[1,q_{1}) and IkJk(qr1,qr)I_{k}\subset J_{k}\subset(q_{r-1},q_{r}), for r=2,,n2jr=2,\ldots,n_{2j}. Provided that qn2jq_{n_{2j}} is sufficiently large and using λ1==λn2j=1\lambda_{1}=\cdots=\lambda_{n_{2j}}=1, the resulting pair (z,ζ)(z,\zeta) has the desired properties.

Step 2: For each j0j_{0}\in\mathbb{N} there exists a vector ww in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with w480Cmj02\|w\|\leq 480Cm_{j_{0}}^{-2} and Twε/m2j01\|Tw\|\geq\varepsilon/m_{2j_{0}-1}. In particular, Tmj0ε/480C\|T\|\geq m_{j_{0}}\varepsilon/{480C}.

This vector ww is found by performing a similar construction process as in the proof of Proposition 3.8 to cosntruct a (16C,2j0,0)(16C,2j_{0},0)-dependent sequence (zi)i=1n2j01(z_{i})_{i=1}^{n_{2j_{0}-1}} and a γΓ\gamma\in\Gamma “close” to κ0\kappa_{0} with weight(γ)=m2j011\operatorname{weight}(\gamma)=m_{2j_{0}-1}^{-1}, κ(γ)\kappa(\gamma) and evaluation analysis

eγ=i=1n2j01dξi+1m2j01i=1n2j01eζiPJi,e_{\gamma}^{*}=\sum_{i=1}^{n_{2j_{0}-1}}d^{*}_{\xi_{i}}+\frac{1}{m_{2j_{0}-1}}\sum_{i=1}^{n_{2j_{0}-1}}e^{*}_{\zeta_{i}}\circ P_{J_{i}},

where JiJ_{i} is an interval of \mathbb{N} containing range(zi)\operatorname{range}(z_{i}) and range(Tzi)\operatorname{range}(Tz_{i}). In each step, if (zi)i=1i0(z_{i})_{i=1}^{i_{0}} and ξi0\xi_{i_{0}} have been chosen, use Step 1 to find a (16C,4σ(ξi0),0)(16C,4\sigma(\xi_{i_{0}}),0)-exact pair (zi0+1,ζi0+1)(z_{i_{0}+1},\zeta_{i_{0+1}}) with minsupp(zi0+1)>rank(ξi0)\min\operatorname{supp}(z_{i_{0}+1})>\operatorname{rank}(\xi_{i_{0}}), ϱ(κ(ζi0+1),κ0)2(rank(ξi0)+1)\varrho(\kappa(\zeta_{i_{0}+1}),\kappa_{0})\leq 2^{-(\operatorname{rank}(\xi_{i_{0}})+1)}, and eηi0+1(Tzi0+1)=|eηi0+1(Tzi0+1)|εe_{\eta_{i_{0}+1}}^{*}(Tz_{i_{0}+1})=|e_{\eta_{i_{0}+1}}^{*}(Tz_{i_{0}+1})|\geq\varepsilon. Pick qi0+1>max(range(zi0+1)range(Tzi0+1)q_{i_{0}+1}>\max(\operatorname{range}(z_{i_{0}+1})\cup\operatorname{range}(Tz_{i_{0}+1}) and choose an appropriate ξi0+1Δqi0+1(b),odd\xi_{i_{0}+1}\in\Delta_{q_{i_{0}+1}}^{\text{\ref{age nonzero},odd}}.

Then, put w=n2j011i=1n2j01ziw=n_{2j_{0}-1}^{-1}\sum_{i=1}^{n_{2j_{0}-1}}z_{i}. It follows that eγ(Tz)=|eγ(Tz)|ε/m2j01e_{\gamma}^{*}(Tz)=|e_{\gamma}^{*}(Tz)|\geq\varepsilon/m_{2j_{0}-1} and, by Proposition 6.10, z480Cm2j012\|z\|\leq 480Cm^{-2}_{2j_{0}-1}.

Because j0j_{0} was arbitrary, T\|T\| cannot be finite and thus TT is unbounded. ∎

The next result may be viewed as first step towards the fact that a bounded TT must be close to a diagonal operator (Corollary 5.4).

Proposition 7.4.

Let T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} be bounded linear operator. Then,

  1. (i)

    for any RIS (xk)k(x_{k})_{k} in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}, limkTxkPrange(xk)Txk=0\lim_{k}\|Tx_{k}-P_{\operatorname{range}(x_{k})}Tx_{k}\|=0 and

  2. (ii)

    limrank(γ)dγ(Tdγ)dγTdγ=0\lim_{\operatorname{rank}(\gamma)\to\infty}\big{\|}d_{\gamma}^{*}(Td_{\gamma})d_{\gamma}-Td_{\gamma}\big{\|}=0.

Proof.

To prove (i), for each kk\in\mathbb{N} write range(xk)=[pk,qk]\operatorname{range}(x_{k})=[p_{k},q_{k}] and set Ik=[1,pk)I_{k}=[1,p_{k}), Ik=(qk,)I^{\prime}_{k}=(q_{k},\infty). If (i) is false, either lim infkPIkTxk>0\liminf_{k}\|P_{I_{k}}Tx_{k}\|>0, or lim infkPIkTxk>0\liminf_{k}\|P_{I_{k}^{\prime}}Tx_{k}\|>0. These cases are treated in the same way so assume the first one holds. It may also be assumed that (Txk)k(Tx_{k})_{k} is a block sequence. Therefore, (PIkTxk)k(P_{I_{k}}Tx_{k})_{k} is also a block sequence with lim infkPIkTxk>0\liminf_{k}\|P_{I_{k}}Tx_{k}\|>0. Apply Proposition 7.1 and pass to a subsequence to find, for each kk\in\mathbb{N}, an interval JkJ_{k} of range(PIkTxk)Ik\operatorname{range}(P_{I_{k}}Tx_{k})\subset I_{k} and ηk\eta_{k} in Γ\Gamma such that limkweight(Jk,ηk)=0 and lim infk|eηk(PJkxk)|>0\lim_{k}\operatorname{weight}(J_{k},\eta_{k})=0\text{ and }\liminf_{k}\big{|}e^{*}_{\eta_{k}}\big{(}P_{J_{k}}x_{k}\big{)}\big{|}>0. But then, because range(xk)Ik=\operatorname{range}(x_{k})\cap I_{k}=\emptyset, it follows that eηk(PJkxk)=0e_{\eta_{k}}^{*}(P_{J_{k}}x_{k})=0, for all kk\in\mathbb{N}. By Proposition 7.3, TT is unbounded.

To proceed with the proof of (ii), recall that for each nn\in\mathbb{N}, (dγ)γΔn(d_{\gamma})_{\gamma\in\Delta_{n}} is 2-equivalent to the unit vector basis of (Δn)\ell_{\infty}(\Delta_{n}). Therefore, for each γ0Δn\gamma_{0}\in\Delta_{n},

P{rank(γ0)}Tdγ0dγ0(Tdγ0)dγ0\displaystyle\Big{\|}P_{\{\operatorname{rank}(\gamma_{0})\}}Td_{\gamma_{0}}-d_{\gamma_{0}}^{*}\big{(}Td_{\gamma_{0}}\big{)}d_{\gamma_{0}}\Big{\|} =γΔn{γ0}dγ(Tdγ0)dγ\displaystyle=\Big{\|}\sum_{\gamma^{\prime}\in\Delta_{n}\setminus\{\gamma_{0}\}}d_{\gamma^{\prime}}^{*}\big{(}Td_{\gamma_{0}}\big{)}d_{\gamma^{\prime}}\Big{\|}
2maxγΔn{γ0}|dγ(Tdγ0)|.\displaystyle\leq 2\max_{\gamma^{\prime}\in\Delta_{n}\setminus\{\gamma_{0}\}}\big{|}d_{\gamma^{\prime}}^{*}\big{(}Td_{\gamma_{0}}\big{)}\big{|}.

Therefore, it suffices to check that

limnmaxγγΔn|dγ(Tdγ)|=0.\lim_{n}\max_{\gamma\neq\gamma^{\prime}\in\Delta_{n}}\Big{|}d^{*}_{\gamma^{\prime}}\big{(}Td_{\gamma}\big{)}\Big{|}=0.

If this is false then there exist ε>0\varepsilon>0, a strictly increasing sequence (qk)k(q_{k})_{k} in \mathbb{N}, and for each kk\in\mathbb{N}, γkγkΔqk\gamma_{k}\neq\gamma^{\prime}_{k}\in\Delta_{q_{k}}, such that |eγk(P{qk}Tdγk)|=|dγk(Tdγk)|ε|e_{\gamma^{\prime}_{k}}^{*}(P_{\{q_{k}\}}Td_{\gamma_{k}})|=|d_{\gamma_{k}^{\prime}}(Td_{\gamma_{k}})|\geq\varepsilon. Observe that weight({qk},γk)=0\operatorname{weight}(\{q_{k}\},\gamma_{k}^{\prime})=0 and eγk(P{qk}dγk)=dγk(dγk)=0e_{\gamma_{k}^{\prime}}(P_{\{q_{k}\}}d_{\gamma_{k}})=d_{\gamma_{k}^{\prime}}^{*}(d_{\gamma_{k}})=0, for all kk\in\mathbb{N}. By passing to a subsequence, as in example 6.2, (dγk)k(d_{\gamma_{k}})_{k} is a RIS. Proposition 7.3 yields that TT is unbounded. ∎

7.3. Eventual continuity of diagonal entries

In this section the two black-box theorems of Section 5 are proved. For convenience, their statements are repeated as propositions.

Proposition 7.5 (Theorem 5.1).

Let T:𝔛C(K)𝔛C(K)T:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} be a bounded linear operator. Then, for every ε>0\varepsilon>0 there exist nn\in\mathbb{N} and δ>0\delta>0 such that for all γ\gamma,γΓ\gamma^{\prime}\in\Gamma with min{rank(γ),rank(γ)}n\min\{\operatorname{rank}(\gamma),\operatorname{rank}(\gamma^{\prime})\}\geq n and ϱ(κ(γ),κ(γ))<δ\varrho(\kappa(\gamma),\kappa(\gamma^{\prime}))<\delta,

(16) |dγ(Tdγ)dγ(Tdγ)|<ε.\big{|}d_{\gamma}^{*}\big{(}Td_{\gamma}\big{)}-d_{\gamma^{\prime}}^{*}\big{(}Td_{\gamma^{\prime}}\big{)}\big{|}<\varepsilon.

Therefore, the function φT:K\varphi_{T}:K\to\mathbb{C} given by

φT(κ)=limrank(γ)κ(γ)κdγ(Tdγ)\varphi_{T}(\kappa)=\lim_{\begin{subarray}{c}\operatorname{rank}(\gamma)\to\infty\\ \kappa(\gamma)\to\kappa\end{subarray}}d_{\gamma}^{*}\big{(}Td_{\gamma}\big{)}

is well defined and continuous.

Proof.

If the conclusion is false, there exist sequences (ηk)(\eta_{k}), (ζk)(\zeta_{k}) and κ0K\kappa_{0}\in K such that ϱ(κ(ηk),κ0)0\varrho(\kappa(\eta_{k}),\kappa_{0})\to 0 and ϱ(κ(ζk),κ0)=0\varrho(\kappa(\zeta_{k}),\kappa_{0})=0 yet limkdηk(Tdηk)=λμ=limkdζ(Tdζ)\lim_{k}d_{\eta_{k}}^{*}(Td_{\eta_{k}})=\lambda\neq\mu=\lim_{k}d_{\zeta}^{*}(Td_{\zeta}). By virtue of Proposition 7.4 (ii), and by perturbing TT by a compact operator, it may be assumed that for all kk\in\mathbb{N}, Tdηk=λdηkTd_{\eta_{k}}=\lambda d_{\eta_{k}} and Tdζk=μdζkTd_{\zeta_{k}}=\mu d_{\zeta_{k}}. This will make it possible to perform a process similar to that in the proof of Proposition 7.3 to “blow up” the norm of TT.

Note that for some C>0C>0, it is possible to pass to subsequences of (ηk)(\eta_{k}) and (ζk)(\zeta_{k}) so that if

(θk)k=(η1,ζ1,η2,ζ2,,ηk,ζk,)(\theta_{k})_{k}=({\eta_{1}},{\zeta_{1}},{\eta_{2}},{\zeta_{2}},\ldots,{\eta_{k}},{\zeta_{k}},\ldots)

then the sequence ((1)kdθk)((-1)^{k}d_{\theta_{k}}) is a skipped block CC-RIS. This argument is similar to Example 6.2. If limkweight(ηk)=limkweight(ζk)=0\lim_{k}\operatorname{weight}(\eta_{k})=\lim_{k}\operatorname{weight}(\zeta_{k})=0 then one can choose this sequence to be a 22-RIS. If instead there exists j0j_{0}\in\mathbb{N} such that lim supkmax(weight(ηk),weight(ζk))=mj01\limsup_{k}\max(\operatorname{weight}(\eta_{k}),\operatorname{weight}(\zeta_{k}))=m_{j_{0}}^{-1}, then ((1)kdθk)((-1)^{k}d_{\theta_{k}}) may be chosen to be 2mj02m_{j_{0}}-RIS. Furthermore, while choosing (θk)k(\theta_{k})_{k}, it is possible to choose a sequence of natural numbers (qk)k(q_{k})_{k} with rank(θ1)<q1<rank(θ2)<q2<\operatorname{rank}(\theta_{1})<q_{1}<\operatorname{rank}(\theta_{2})<q_{2}<\cdots such that, for each k2k\geq 2, ϱ(κ(θk),κ0)2(qk1+1)\varrho(\kappa(\theta_{k}),\kappa_{0})\leq 2^{-(q_{k-1}+1)}. Also recall that for each kk\in\mathbb{N}, dθk=eθkP{rank(θk)}d_{\theta_{k}}^{*}=e_{\theta_{k}}^{*}\circ P_{\{\operatorname{rank}(\theta_{k})\}} and weight({rank(θk)},θk)=0\operatorname{weight}(\{\operatorname{rank}(\theta_{k})\},\theta_{k})=0.

Step 1: For each j,qj,q\in\mathbb{N} and δ>0\delta>0, there exists a (16C,2j,0)(16C,2j,0)-exact pair (z,ζ)(z,\zeta) such that minsupp(z)>q\min\operatorname{supp}(z)>q, ϱ(κ(η),κ0)δ\varrho(\kappa(\eta),\kappa_{0})\leq\delta, and |eζ(Tz)|ε=|λμ|/2|e_{\zeta}^{*}(Tz)|\geq\varepsilon=|\lambda-\mu|/2.

This is a repetition of the argument in Lemma 6.8. Omitting a few initial terms of the sequence ((1)kdθk)k((-1)^{k}d_{\theta_{k}})_{k}, there exists a ζΓ\zeta\in\Gamma with evaluation analysis

eζ=k=1n2jdξk+1m2jk=1n2jdθk,e_{\zeta}^{*}=\sum_{k=1}^{n_{2j}}d_{\xi_{k}}^{*}+\frac{1}{m_{2j}}\sum_{k=1}^{n_{2j}}d^{*}_{\theta_{k}},

where rank(ξk)=qk\operatorname{rank}(\xi_{k})=q_{k}, for k=1,,n2jk=1,\ldots,n_{2j}. Put z=m2jn2j1k=1n2j(1)kdθkz=m_{2j}n_{2j}^{-1}\sum_{k=1}^{n_{2j}}(-1)^{k}d_{\theta_{k}}. Then, (z,ζ)(z,\zeta) is a (16C,2j,0)(16C,2j,0)-exact pair (recall that n2jn_{2j} is even) and eζ(Tz)=(λμ)/2e^{*}_{\zeta}(Tz)=(\lambda-\mu)/2.

Step 2: For each j0j_{0}\in\mathbb{N} there exists a vector ww in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} with w480Cmj02\|w\|\leq 480Cm_{j_{0}}^{-2} and Twε/m2j01\|Tw\|\geq\varepsilon/m_{2j_{0}-1}. In particular, Tmj0ε/480C\|T\|\geq m_{j_{0}}\varepsilon/{480C}.

This is identical to the proof of the second step in Proposition 7.3 and, in conclusion, TT is unbounded. ∎

Proposition 7.6 (Theorem 5.2).

A bounded linear operator T:(𝔛C(K))(𝔛C(K))T:\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})\to\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is compact if and only if φT=0\varphi_{T}=0.

Proof.

If TT is compact, then it sends weakly null sequences to norm-null ones. Because (dγ)γΓ(d_{\gamma})_{\gamma\in\Gamma} is shrinking, it follows that φT=0\varphi_{T}=0.

Assume now that TT is a bounded linear operator with φT=0\varphi_{T}=0, i.e., limrank(γ)dγ(Tdγ)=0\lim_{\operatorname{rank}(\gamma)\to\infty}d_{\gamma}^{*}(Td_{\gamma})=0. Proposition 7.4 (ii) yields

(17) limrank(γ)Tdγ=0.\lim_{\operatorname{rank}(\gamma)\to\infty}\big{\|}Td_{\gamma}\big{\|}=0.

Towards contradiction, assume that TT is not compact. Then, there exists a RIS (xk)k(x_{k})_{k} with lim supkTxk>0\limsup_{k}\|Tx_{k}\|>0. As usual, assume that (Txk)k(Tx_{k})_{k} is a block sequence for. By Proposition 7.1 one may pass to a further subsequence and find ε>0\varepsilon>0 such that, for each kk\in\mathbb{N}, there exist an interval IkI_{k} of range(Txk)\operatorname{range}(Tx_{k}) and ηkΓ\eta_{k}\in\Gamma such that |eηk(PIkTxk)|ε|e_{\eta_{k}}^{*}(P_{I_{k}}Tx_{k})|\geq\varepsilon and limkweight(Ik,ηk)=0\lim_{k}\operatorname{weight}(I_{k},\eta_{k})=0.

Assume for the moment that rank(ηk)Ik\operatorname{rank}(\eta_{k})\in I_{k}, for all kk\in\mathbb{N}. Although this might not be true, later the general case will be reduced to this one. For each kk\in\mathbb{N}, consider the sequence yk=xkeηk(PIkxk)dηky_{k}=x_{k}-e_{\eta_{k}}^{*}(P_{I_{k}}x_{k})d_{\eta_{k}}. By passing to a subsequence, this is a RIS and because it was assumed that rank(ηk)Ik\operatorname{rank}(\eta_{k})\in I_{k}, for all kk\in\mathbb{N}, it follows that

eηk(PIkyk)=eηk(PIkxk)eetak(PIkxk)eηk(PIkdηk)=0.e_{\eta_{k}}^{*}\big{(}P_{I_{k}}y_{k}\big{)}=e_{\eta_{k}}^{*}\big{(}P_{I_{k}}x_{k}\big{)}-e_{\\ eta_{k}}^{*}\big{(}P_{I_{k}}x_{k}\big{)}e_{\eta_{k}}^{*}\big{(}P_{I_{k}}d_{\eta_{k}}\big{)}=0.

By (17), lim inf|eηk(PIkTyk)|=lim inf|eηk(PIkTxk)|ε\liminf|e_{\eta_{k}}^{*}(P_{I_{k}}Ty_{k})|=\liminf|e_{\eta_{k}}^{*}(P_{I_{k}}Tx_{k})|\geq\varepsilon, By Proposition 7.3, TT is unbounded.

To treat the remaining case, assume that rank(ηk)Ik\operatorname{rank}(\eta_{k})\notin I_{k}, for all kk\in\mathbb{N}. This in particular implies that rank(ηk)>max(Ik)\operatorname{rank}(\eta_{k})>\max(I_{k}) because otherwise eηkPIk=0e_{\eta_{k}}^{*}\circ P_{I_{k}}=0. In particular, weight(Ik,ηk)=weight(ηk)\operatorname{weight}(I_{k},\eta_{k})=\operatorname{weight}(\eta_{k}), for all kk\in\mathbb{N}. Let IkI_{k}^{\prime} be the smallest interval of \mathbb{N} containing IkI_{k} and rank(ηk)\operatorname{rank}(\eta_{k}). Write Ik=IkIk′′I_{k}^{\prime}=I_{k}\cup I_{k}^{\prime\prime}, where Ik′′={max(Ik)+1,,rank(ηk)}I_{k}^{\prime\prime}=\{\max(I_{k})+1,\ldots,\operatorname{rank}(\eta_{k})\}. Then, eηkPIk=eηkPIkeηkPIk′′e_{\eta_{k}}^{*}\circ P_{I_{k}}=e_{\eta_{k}}^{*}\circ P_{I^{\prime}_{k}}-e_{\eta_{k}}^{*}\circ P_{I_{k}^{\prime\prime}}. Therefore,

either |eηk(PIkxk)|ε/2 or |eηk(PIkxk)|ε/2.\text{either }\big{|}e_{\eta_{k}}^{*}\big{(}P_{I^{\prime}_{k}}x_{k}\big{)}\big{|}\geq\varepsilon/2\text{ or }\big{|}e_{\eta_{k}}^{*}\big{(}P_{I^{\prime}_{k}}x_{k}\big{)}\big{|}\geq\varepsilon/2.

Let Jk=IkJ_{k}=I_{k}^{\prime} or Jk=Ik′′J_{k}=I_{k}^{\prime\prime} accordingly. In either case, rank(ηk)Jk\operatorname{rank}(\eta_{k})\in J_{k} and weight(Jk,ηk)weight(ηk)=weight(Ik,ηk)0\operatorname{weight}(J_{k},\eta_{k})\leq\operatorname{weight}(\eta_{k})=\operatorname{weight}(I_{k},\eta_{k})\to 0. Note that it might no longer be true that Jkrange(xk)J_{k}\subset\operatorname{range}(x_{k}), but this was not required in the proof of the special case above. ∎

8. Additional properties of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

This section outlines some additional properties of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} without giving detailed proofs. For example, the complemented subspaces of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} are classified and it is mentioned that 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} does not contain unconditional basic sequences.

8.1. Complemented subspaces of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

Denote by Clopen(K)\mathrm{Clopen}(K) the collection of clopen subsets of KK. The wealth of projections on 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is directly linked to this collection. For each FClopen(K)F\in\mathrm{Clopen}(K), the characteristic χF:K\chi_{F}:K\to\mathbb{C} is Lipschitz and therefore the diagonal operator χ^K:𝔛C(K)𝔛C(K)\hat{\chi}_{K}:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is a bounded 0-1 valued diagonal operator, in particular it is the canonical projection onto the subspace WF={dγ:κ(γ)F}¯W_{F}=\overline{\langle\{d_{\gamma}:\kappa(\gamma)\in F\}\rangle}. Observe that, because FF is open, whenever FF\neq\emptyset then WFW_{F} is infinite dimensional. This is because the sequence (Kn)n(K_{n})_{n} is increasing with dense union and for each nn\in\mathbb{N}, {κ(γ):γΔn}=Kn\{\kappa(\gamma):\gamma\in\Delta_{n}\}=K_{n}. For each KClopen(K)K\in\mathrm{Clopen}(K), denote PF=χ^KP_{F}=\hat{\chi}_{K}.

Proposition 8.1.

The map FPFF\mapsto P_{F} is an injection from Clopen(K)\mathrm{Clopen}(K) into the collection of bounded canonical basis projections such that PFP_{F} is of infinite rank if and only if FF\neq\emptyset. Furthermore the following hold.

  1. (i)

    For every projection P:𝔛C(K)𝔛C(K)P:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} there exists a unique FClopen(K)F\in\mathrm{Clopen}(K) such that PPFP-P_{F} is compact.

  2. (ii)

    For every complemented subspace WW of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} there exist FClopen(K)F\in\mathrm{Clopen}(K) and n{0}n\in\mathbb{N}\cup\{0\} such that either WWFnW\simeq W_{F}\oplus\mathbb{C}^{n} or WWF/nW\simeq W_{F}/\mathbb{C}^{n}.

  3. (iii)

    For each FClopen(K)F\in\mathrm{Clopen}(K), WFW_{F} has the diagonal-plus-compact property.

Sketch of proof.

For (i), note that ϕ=Ψ(P)\phi=\Psi(P) is an idempotent in C(K)C(K) and therefore the characteristic of a clopen set FF. By Theorem 5.2, PPFP-P_{F} is compact. The uniqueness comes from the fact that for FGF\neq G, PFPGP_{F}-P_{G} is non-compact.

For (ii), let PP be a projection onto WW and write P=PF+AP=P_{F}+A, with AA compact. By Proposition 4.1, there exists NN\in\mathbb{N} such that for all nNn\geq N, PFP[n,)4\|P_{F}P_{[n,\infty)}\|\leq 4. For 0<ε<10<\varepsilon<1 and nn sufficiently large, AP[n,)ε\|AP_{[n,\infty)}\|\leq\varepsilon. Some classical functional analysis gymnastics yield that the space Z=P(PFP[n,)𝔛C(K))Z=P(P_{F}P_{[n,\infty)}\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is 4(1+ε)/(1ε)4(1+\varepsilon)/(1-\varepsilon)-complemented in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. Also, ZPFP[n,)𝔛C(K)Z\simeq P_{F}P_{[n,\infty)}\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} and dim(W/Z)<n\dim(W/Z)<n.

Item (iii) follows from extending an operator on WFW_{F} to 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} and using Corollary 5.4. Note that in particular the Calkin algebra of WFW_{F} is C(F)C(F). ∎

The next statement discusses the decomposability of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}.

Proposition 8.2.
  1. (i)

    The space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is indecomposable if and only if KK is connected.

  2. (ii)

    If KK is totally disconnected, then every infinite dimensional complemented subspace of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is decomposable.

Sketch of proof.

The first statement follows directly from Proposition 8.1 (i). For the second one, by Proposition 8.1 (ii) it suffices to check for a subspace of the form WFW_{F}. Write FF as the disjoint union of two non-empty clopen sets F1F_{1}, F2F_{2} and observe that WF=WF1WF2W_{F}=W_{F_{1}}\oplus W_{F_{2}}. It is perhaps of some interest that this process can be combined with Proposition 4.1 to construct an uncomplemented subspace of WFW_{F} with an infinite dimensional Schauder decomposition. Note that all that was required is that every non-empty clopen subset of KK contains a further proper non-empty clopen subset. This is strictly weaker than KK being totally disconnected. ∎

Remark 8.3.

In all cases, 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} has, up to isomorphism, countably many complemented subspaces. This is because Clopen(K)\mathrm{Clopen}(K) is countable and Proposition 8.1 (ii). An argument similar to that of Proposition 8.1 (iii) shows that for FGClopen(K)F\neq G\in\mathrm{Clopen}(K), WFW_{F} is not isomorphic to a finite codimensional subspace of WGW_{G}. This can be used to resolve the isomorphic containment relation on the complemented subspaces of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} based on the inclusion relation on Clopen(K)\mathrm{Clopen}(K).

8.2. The subspace structure of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}

This section discusses the lack of unconditional sequences in 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} and whether 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is HI.

Proposition 8.4.
  1. (i)

    The space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} does not contain unconditional sequences and therefore it is HI-saturated

  2. (ii)

    The space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is hereditarily indecomposable if and only if KK is a singleton.

  3. (iii)

    A complemented subspace of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is not isomorphic to any of its proper subspaces.

  4. (iv)

    The space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} does not admit an infinite dimensional Schauder decomposition.

Sketch of proof.

For the first two statements, it is necessary to show that every block subspace of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} contains a 2-RIS. The main difference to how this is proved in [4] is here it is necessary to use Proposition 7.1 and then perform an argument similar to that in [4, Lemma 8.2].

To show that 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} contains no unconditional sequence, on any block subspace one builds a RIS (xk)k(x_{k})_{k} that is normalized by a very fast growing sequence (eηkPIk)k(e_{\eta_{k}}^{*}\circ P_{I_{k}})_{k} with κ(ηk)κ0\kappa(\eta_{k})\to\kappa_{0} and to build a one-exact pair. The rest of the process is similar to [4, Section 8]. By Gowers’ famous dichotomy theorem from [16], 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is HI-saturated.

If KK is a singleton then the metric constraint trivializes; for every γ,γΓ\gamma,\gamma^{\prime}\in\Gamma, κ(γ)=κ(γ)\kappa(\gamma)=\kappa(\gamma^{\prime}). Therefore exact pairs coming from different block subspaces can be connected to show the HI property, as in [4, Section 8]. If KK is not a singleton, then take two non-empty open sets UU, VV with positive distance. Using the metric, define a Lipschitz function ϕ\phi with ϕ|U=1\phi|_{U}=1 and ϕV=0\phi_{V}=0. Then, ϕ^U\hat{\phi}_{U} restricted on {dγ:κ(γ)UV}¯\overline{\langle\{d_{\gamma}:\kappa(\gamma)\in U\cup V\}\rangle} is a projection onto {dγ:κ(γ)U}¯\overline{\langle\{d_{\gamma}:\kappa(\gamma)\in U\}\rangle} with kernel {dγ:κ(γ)V}¯\overline{\langle\{d_{\gamma}:\kappa(\gamma)\in V\}\rangle}.

Statement (iii) is an application of the Fredhold index on spaces of the form WFW_{F}. Note that every semi-Fredhold diagonal operator on WFW_{F} is Fredholm of index zero. By Proposition 8.1 (iii), every semi-Fredholm operator on WFW_{F} is Fredholm of index zero.

The final statement follows from Proposition 8.1 (i) and compactness. Assume that (Pn)n(P_{n})_{n} is a sequence of infinite dimensional projections with increasing ranges (Wn)n(W_{n})_{n}, such that dim(Wn+1/Wn)=\dim(W_{n+1}/W_{n})=\infty for all nn\in\mathbb{N}, that converges in the strong operator topology to the identity. For each nn\in\mathbb{N} Write Pn=PFn+AnP_{n}=P_{F_{n}}+A_{n}. It follows that F1F2F_{1}\subsetneq F_{2}\subsetneq\cdots and nFn=K\cup_{n}F_{n}=K. Compactness prohibits this. ∎

Remark 8.5.

Although it is likely that no subspace of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} is isomorphic to its proper subspaces, the lack of convex combinations in the definition of Γ\Gamma does not allow the usual proofs to go through.

8.3. The space of bounded linear operators (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})

This brief section visits the space (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) with regards to its ideal and subspace structure.

Proposition 8.6.
  1. (i)

    The space (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) does not contain an isomorphic copy of c0c_{0}. In particular, unless KK is finite, (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) does not contain an isomorphic copy of C(K)C(K) and therefore 𝒦(𝔛C(K))\mathcal{K}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is not complemented in (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}).

  2. (ii)

    The quotient map []:(𝔛C(K))𝒞𝒶𝓁(𝔛𝒞(𝒦))[\cdot]:\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}})\to\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}) is strictly singular if an only if KK is countable.

  3. (iii)

    Denote by Open(K)\mathrm{Open}(K) the collection of open subsets of 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}. There exists an order preserving bijection between Open(K)\mathrm{Open}(K) and the non-zero closed two sided ideals of (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}).

Comment on Proof.

The proof of item (i) is outlined in [33, Remark 4.5, page 65]. The fact that for KK countable the quotient map is strictly singular is also explained in Remark 4.6 of that same paper. If KK is uncountable, then C(K)C(K) contains an isomorphic copy of 1\ell_{1} that can be lifted by [][\cdot] to (𝔛C(K))\mathcal{L}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}). Item (iii) was explained in [23, Remark 1.5 (vi), page 1022]. ∎

8.4. Very incomparable spaces with C(K)C(K) Calkin algebras

Similarly to [4, Section 10.2, page 46], by varying the the sequence (mj,nj)j(m_{j},n_{j})_{j} is is possible to create different versions of the space 𝔛C(K)\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}} that are very incomparable to one another. Recall that, up to homeomorhism, there are continuum many compact metrizable spaces, as they may be identified with the closed subsets of [0,1][0,1]^{\mathbb{N}}.

Proposition 8.7.

There exists a collection of Banach spaces

{𝔛C(K)α:α<𝔠,KClosed([0,1])}\big{\{}\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{\alpha}:\alpha<\mathfrak{c},\;K\in\mathrm{Closed}([0,1]^{\mathbb{N}})\big{\}}

with the following properties.

  1. (i)

    For each 𝔛C(K)α\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{\alpha}, 𝒞𝒶𝓁(𝔛𝒞(𝒦)α)\mathpzc{Cal}(\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{\alpha}) is homomorphically isometric to C(K)C(K).

  2. (ii)

    For (α,K)(β,L)(\alpha,K)\neq(\beta,L) every bounded linear operator T:𝔛C(K)α𝔛C(L)βT:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{\alpha}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!L\>\!\!)}}^{\beta} is compact.

Comment on Proof.

This is achieved by choosing an almost disjoint family of infinite subsets of the natural numbers {LKα:α<𝔠,KClosed([0,1])}\{L^{\alpha}_{K}:\alpha<\mathfrak{c},\;K\in\mathrm{Closed}([0,1]^{\mathbb{N}})\}. Then for each such LKαL^{\alpha}_{K} define a space 𝔛C(K)α\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{\alpha} with Calkin algebra C(K)C(K) using (mj,nj)jLKα(m_{j},n_{j})_{j\in L^{\alpha}_{K}}. The proof that for each (α,K)(β,L)(\alpha,K)\neq(\beta,L) every bounded linear operator T:𝔛C(K)α𝔛C(L)βT:\mathfrak{X}_{{}^{C\>\!\!(\>\!\!K\>\!\!)}}^{\alpha}\to\mathfrak{X}_{{}^{C\>\!\!(\>\!\!L\>\!\!)}}^{\beta} is compact is very similar to [4, Theorem 10.4, page 47], with the assistance of Proposition 7.1. ∎

Corollary 8.8.

Let k1,,knk_{1},\ldots,k_{n}\in\mathbb{N} and K1,,KnK_{1},\ldots,K_{n} be compact metric spaces. There exists a Banach space XX with 𝒞𝒶𝓁(𝒳)\mathpzc{Cal}(X) isomorphic as a Banach algebra to Mk1(C(K1))Mkn(C(Kn))M_{k_{1}}(C(K_{1}))\oplus\cdots\oplus M_{k_{n}}(C(K_{n})).

Comment on Proof.

It was observed by Kania and Laustsen in [23, Note added in proof, page 1022] that every finite dimensional semi-simple complex algebra is a Calkin algebra. Using Proposition 8.7, their argument works perfectly well to find Calkin algebras of the above form. ∎

9. Open problems

With regards to the question of what unital algebras can be realized as Calkin algebras of a Banach space (see, e.g., Tarbard’s PhD thesis [44, page 134] in 2012), there is a lot of progress to be made. For example, there does not exist a known property of unital Banach algebras that precludes them from being Calkin algebras (some progress was made in [20] which is discussed further below). The appearance of the Argyros-Haydon construction led to the description of a plethora of explicit Calkin algebras, both finite and infinite dimensional. Unlike the Calkin algebras of classical spaces, these examples are always separable and (in essence) commutative. Additionally, all infinite dimensional ones are non-reflexive.

From a Banach-space-theoretic perspective, the construction of a reflexive Calkin algebra seems particularly intriguing and challenging. In [32] a quasireflexive 𝒞𝒶𝓁(𝒳)\mathpzc{Cal}(X) was found. Although this may seem very close, the underlying space XX has an infinite dimensional Schauder decomposition. This must be avoided in order to achieve reflexivity of the Calkin algebra (compare this to Proposition 8.4 (iv)).

Problem 1.

Does there exist a reflexive and infinite dimensional Calkin algebra?

It is worth mentioning that, if viewed simply as Banach spaces, there exists a large variety of Calkin algebras. In [32] the author, Puglisi, and Tolias proved that there exist HI Calkin algebras and that every non-reflexive space with an unconditional basis is a Calkin algebra, albeit with a non-standard multiplication.

Constructing non-separable explicit Calkin algebras would require the development of additional tools. Of particular relevance is the question of the existence of a space with an unconditional basis that has the diagonal-plus-compact property. Such a space’s Calkin algebra would automatically be isomorphic to C(β)C(\beta\mathbb{N}\!\setminus\!\mathbb{N}). Another path worth exploring is the existence of some type of non-separable Argyros-Haydon space (e.g., based on the Bourgain-Pisier construction from [9] and its non-separable version by Lopez-Abad from [28]). It is worth pointing out that there already exist known examples of non-separable C(K)C(K) algebras with representations of the form (X)/𝒮𝒮(X)\mathcal{L}(X)/\mathcal{SS}(X) (Koszmider, [24] and Plebanek, [40]). It unclear however how the methods from these two papers could be used to study the following.

Problem 2.

Does there exist a non-separable C(K)C(K) space that is a Calkin algebra?

Horváth and Kania proved in [20] that for any cardinal λ\lambda there exists a C(K)C(K) space of density 2λ2^{\lambda} that is not the Calkin algebra of any space of density λ\lambda. Of course, this does not mean that C(K)C(K) is not the Calkin algebra of a space with larger density.

The current paper yields that every separable and commutative CC^{*}-algebra can be represented as a Calkin algebra. Outside this class, there still remain classical commutative Banach algebras for which the existence of such a representation is unknown, e.g., the convolution algebra L1(G)L_{1}(G) for an abelian locally compact polish group GG. Note that Tarbard’s Calkin algebra from [44] is not of this type, because it is 1(0)\ell_{1}(\mathbb{N}_{0}), a semigroup algebra. This space is closely related to the disk algebra, another example of interest. In a personal communication with the author, J. Pachl asked the following.

Problem 3.

What semigroup algebras admit representations as Calkin algebras?

There exist non-commutative explicit Calkin algebras, such as all finite dimensional semi-simple complex algebras (as observed by Laustsen and Kania in [23, Note added in proof, page 1022]) and algebras of the form Mk1(C(K1))Mkn(C(Kn))M_{k_{1}}(C(K_{1}))\oplus\cdots\oplus M_{k_{n}}(C(K_{n})) (see Corollary 8.8). However, these examples are built by applying elementary processes to commutative ones.

There are additional challenges associated to describing explicit (and “genuinely”) non-commutative Calkin algebras. In [18] Gower and Maurey gave an example of a quotient algebra of some (X)\mathcal{L}(X) that resembles the Cuntz algebra 𝒪n\mathcal{O}_{n}. However, in that construction it is unclear what the kernel of corresponding homomorphism is.

The direction of focusing on the description of explicit non-commutative CC^{*}-algebras as Calkin algebras was proposed to the author in a personal communication by N. C. Phillips, who specifically asked the following.

Problem 4.

Do the following non-commutative CC^{*}-algebras admit representations as Calkin algebras?

  1. (a)

    The UHF algebra of type 22^{\infty}.

  2. (b)

    The Cuntz algebra 𝒪n\mathcal{O}_{n}.

  3. (c)

    The reduced CC^{*}-algebra of the free group on two generators, Cr(𝔽2)C_{r}^{*}(\mathbb{F}_{2}).

  4. (d)

    The full CC^{*}-algebra of the free group on two generators, C(𝔽2)C^{*}(\mathbb{F}_{2}).

Let 𝒜\mathcal{A} denote a specific unital CC^{*}-algebra, e.g., one of the above. The first step towards representing it as a Calkin algebra is to identify the right class of operators 𝒞\mathcal{C} acting on a separable Hilbert space that generates 𝒜\mathcal{A}. The next logical step it to represent 𝒜\mathcal{A} as a quotient algebra of an (X)\mathcal{L}(X) space à-la Gowers-Maurey. This is achieved by creating a space XX where a class modeled on 𝒞\mathcal{C} reigns supreme in (X)\mathcal{L}(X), in the sense that it can be used to approximate all operators, modulo perhaps some small ideal (e.g., the strictly singular or compact operator ideal). The first hurdle in achieving this task is that the classical Gowers-Maurey HI space, being based on Schlumprecht space, resembles 1\ell_{1} and not a Hilbert space. This is the precise reason why in the Gowers-Maurey shift space XX from [18], (X)\mathcal{L}(X) has 1()\ell_{1}(\mathbb{Z}) as a quotient algebra instead of C(𝕋)C(\mathbb{T}). The explanation is the following. Considers the class 𝒞\mathcal{C} of all integer powers of the right shift operator on \mathbb{Z}. Acting on 1()\ell_{1}(\mathbb{Z}) this class generates the convolution algebra 1()\ell_{1}(\mathbb{Z}) whereas acting on 2()\ell_{2}(\mathbb{Z}) it generates C(𝕋)C(\mathbb{T}).

A reasonable approach would be to first focus on Banach algebras of operators on 1\ell_{1} that are similar to (a), (b), (c), (d) e.g.,

  1. (1)

    spatial LpL_{p} UHF algebras (Phillips, [38]),

  2. (2)

    LpL_{p}-Cuntz algebras 𝒪np\mathcal{O}_{n}^{p} (Phillips, [36]),

  3. (3)

    reduced group LpL_{p}-operator algebras Fλp(G)F_{\lambda}^{p}(G) (Herz, [19]), and

  4. (4)

    full group LpL_{p}-operator algebras Fp(G)F^{p}(G) (Phillips, [37]).

After some progress has been made in the case p=1p=1, there is some available technology that may be used to generalize, namely the asymptotic-p\ell_{p} HI spaces of Deliyanni and Manoussakis from [13]. It is not entirely clear how one would then proceed to the next step, i.e., representing 𝒜\mathcal{A} as a Calkin algebra, but making it this far would most certainly provide a lot of insight. Calkin algebras not based on 1\ell_{1} were achieved by the author, Puglisi, and Tolias in [32] by combining techniques of Argyros, Deliyanni, and Tolias from [2] and Zisimopoulou from [46].

KK-theory of operator spaces on Banach spaces has been studied since the 1990s when Gowers and Maurey used it in [18] to prove the existence of a Banach space isomorphic to its cube, but not its square. This highlighted the connections between KK-theory, Fredholm theory, and quotients of operator algebras in general Banach spaces. Since then, the KK-theory of (X)\mathcal{L}(X) has been computed for various XX and examples of spaces with interesting KK-theories have been constructed (see, e.g., [25], [26], [47], and [22]). The Gowers-Maurey and Argyros-Haydon spaces have been an important component of this endeavour. Phillips asked the author the following question, which has also been attributed to Laustsen (see, e.g., [47, page 748]).

Problem 5.

Which pairs of abelian groups (G0,G1)(G_{0},G_{1}) can arise as KK-groups (K0((X)),K1((X)))(K_{0}(\mathcal{L}(X)),K_{1}(\mathcal{L}(X))) for some Banach space XX?

It is known among K-theory experts that for every pair of countable abelian groups (G0,G1)(G_{0},G_{1}), there exists a compact metric space XX such that (K0(X),K1(X))=(G0,G1)(K_{0}(X),K_{1}(X))=(\mathbb{Z}\oplus G_{0},G_{1}). Despite the author’s best effort, a reference for this general statement could not be found. An outline of a construction of such XX for finitely generated abelian groups can be found in, e.g., [41, Exercise 13.2, page 228]. The author and Phillips showed in [31] that for every compact metric space XX, if EE is the Banach space given by Theorem A with Calkin algebra C(X)C(X), then (K0((E)),K1((E)))=(K0(X),K1(X))(K_{0}(\mathcal{L}(E)),K_{1}(\mathcal{L}(E)))=(\mathbb{Z}\oplus K_{0}(X),K_{1}(X)). Therefore in Problem 5, for any pair of countable abelian groups (G0,G1)(G_{0},G_{1}), the pair (G0,G1)(\mathbb{Z}\oplus\mathbb{Z}\oplus G_{0},G_{1}) is realizable. A related question is the following.

Problem 6.

Which pairs of abelian groups (G0,G1)(G_{0},G_{1}) can arise as KK-groups (K0(𝒞𝒶𝓁(𝒳)),𝒦1(𝒞𝒶𝓁(𝒳)))(K_{0}(\mathpzc{Cal}(X)),K_{1}(\mathpzc{Cal}(X))) for some Banach space XX?

Acknowledgements

The author would like to thank the anonymous referee for helpful suggestions and, in particular, for recommending the inclusion of Problem 6 and a discussion on KK-theory in Section 9.

References

  • [1] S. A. Argyros and I. Deliyanni. Examples of asymptotic l1l_{1} Banach spaces. Trans. Amer. Math. Soc., 349(3):973–995, 1997.
  • [2] S. A. Argyros, I. Deliyanni, and A. G. Tolias. Hereditarily indecomposable Banach algebras of diagonal operators. Israel J. Math., 181:65–110, 2011.
  • [3] S. A. Argyros, I. Gasparis, and P. Motakis. On the structure of separable \mathscr{L}_{\infty}-spaces. Mathematika, 62(3):685–700, 2016.
  • [4] S. A. Argyros and R. G. Haydon. A hereditarily indecomposable \mathscr{L}_{\infty}-space that solves the scalar-plus-compact problem. Acta Math., 206(1):1–54, 2011.
  • [5] S. A. Argyros and P. Motakis. A reflexive hereditarily indecomposable space with the hereditary invariant subspace property. Proc. Lond. Math. Soc. (3), 108(6):1381–1416, 2014.
  • [6] S. A. Argyros and P. Motakis. The scalar-plus-compact property in spaces without reflexive subspaces. Trans. Amer. Math. Soc., 371(3):1887–1924, 2019.
  • [7] S. A. Argyros and P. Motakis. On the complete separation of asymptotic structures in Banach spaces. Adv. Math., 362:106962, 51, 2020.
  • [8] J. Bourgain and F. Delbaen. A class of special {\mathscr{L}}_{\infty} spaces. Acta Math., 145(3-4):155–176, 1980.
  • [9] J. Bourgain and G. Pisier. A construction of {\mathscr{L}}_{\infty}-spaces and related Banach spaces. Bol. Soc. Brasil. Mat., 14(2):109–123, 1983.
  • [10] L. G. Brown, R. G. Douglas, and P. A. Fillmore. Extensions of CC^{*}-algebras and KK-homology. Ann. of Math. (2), 105(2):265–324, 1977.
  • [11] J. W. Calkin. Two-sided ideals and congruences in the ring of bounded operators in Hilbert space. Ann. of Math. (2), 42:839–873, 1941.
  • [12] S. R. Caradus, W. E. Pfaffenberger, and B. Yood. Calkin algebras and algebras of operators on Banach spaces. Marcel Dekker, Inc., New York, 1974. Lecture Notes in Pure and Applied Mathematics, Vol. 9.
  • [13] I. Deliyanni and A. Manoussakis. Asymptotic lpl_{p} hereditarily indecomposable Banach spaces. Illinois J. Math., 51(3):767–803, 2007.
  • [14] I. Farah. All automorphisms of the Calkin algebra are inner. Ann. of Math. (2), 173(2):619–661, 2011.
  • [15] W. T. Gowers. A solution to Banach’s hyperplane problem. Bull. London Math. Soc., 26(6):523–530, 1994.
  • [16] W. T. Gowers. A new dichotomy for Banach spaces. Geom. Funct. Anal., 6(6):1083–1093, 1996.
  • [17] W. T. Gowers and B. Maurey. The unconditional basic sequence problem. J. Amer. Math. Soc., 6(4):851–874, 1993.
  • [18] W. T. Gowers and B. Maurey. Banach spaces with small spaces of operators. Math. Ann., 307(4):543–568, 1997.
  • [19] C. Herz. Harmonic synthesis for subgroups. Ann. Inst. Fourier (Grenoble), 23(3):91–123, 1973.
  • [20] B. Horváth and T. Kania. Unital Banach algebras not isomorphic to Calkin algebras of separable Banach spaces. Proc. Amer. Math. Soc., 149(11):4781–4787, 2021.
  • [21] S. Kakutani. A generalization of Brouwer’s fixed point theorem. Duke Math. J., 8:457–459, 1941.
  • [22] T. Kania, P. Koszmider, and N. J. Laustsen. Banach spaces whose algebra of bounded operators has the integers as their K0K_{0}-group. J. Math. Anal. Appl., 428(1):282–294, 2015.
  • [23] T. Kania and N. J. Laustsen. Ideal structure of the algebra of bounded operators acting on a Banach space. Indiana Univ. Math. J., 66(3):1019–1043, 2017.
  • [24] P. Koszmider. Banach spaces of continuous functions with few operators. Math. Ann., 330(1):151–183, 2004.
  • [25] N. J. Laustsen. KK-theory for algebras of operators on Banach spaces. J. London Math. Soc. (2), 59(2):715–728, 1999.
  • [26] N. J. Laustsen. KK-theory for the Banach algebra of operators on James’s quasi-reflexive Banach spaces. KK-Theory, 23(2):115–127, 2001.
  • [27] J. Lindenstrauss and A. Pełczyński. Absolutely summing operators in LpL_{p}-spaces and their applications. Studia Math., 29:275–326, 1968.
  • [28] J. Lopez-Abad. A Bourgain-Pisier construction for general Banach spaces. J. Funct. Anal., 265(7):1423–1441, 2013.
  • [29] A. Manoussakis, A. Pelczar-Barwacz, and M. Świȩtek. An unconditionally saturated Banach space with the scalar-plus-compact property. J. Funct. Anal., 272(12):4944–4983, 2017.
  • [30] B. Maurey and H. P. Rosenthal. Normalized weakly null sequence with no unconditional subsequence. Studia Math., 61(1):77–98, 1977.
  • [31] P. Motakis and N. C. Phillips. New examples of KK-groups of algebras of operators on Banach spaces. (In preparation).
  • [32] P. Motakis, D. Puglisi, and A. Tolias. Algebras of diagonal operators of the form scalar-plus-compact are Calkin algebras. Michigan Math. J., 69(1):97–152, 2020.
  • [33] P. Motakis, D. Puglisi, and D. Zisimopoulou. A hierarchy of Banach spaces with C(K)C(K) Calkin algebras. Indiana Univ. Math. J., 65(1):39–67, 2016.
  • [34] E. Odell and Th. Schlumprecht. On the richness of the set of pp’s in Krivine’s theorem. In Geometric aspects of functional analysis (Israel, 1992–1994), volume 77 of Oper. Theory Adv. Appl., pages 177–198. Birkhäuser, Basel, 1995.
  • [35] E. Odell and Th. Schlumprecht. A Banach space block finitely universal for monotone bases. Trans. Amer. Math. Soc., 352(4):1859–1888, 2000.
  • [36] N. C. Phillips. Analogs of Cuntz algebras on LpL^{p} spaces. arxiv:1201.4196.
  • [37] N. C. Phillips. Crossed products of lpl^{p} operator algebras and the kk-theory of Cuntz algebras on lpl^{p} spaces. arxiv:1309.6406.
  • [38] N. C. Phillips. Simplicity of UHF and Cuntz algebras on LpL^{p} spaces. arxiv:1309.0115.
  • [39] N. C. Phillips and N. Weaver. The Calkin algebra has outer automorphisms. Duke Math. J., 139(1):185–202, 2007.
  • [40] G. Plebanek. A construction of a Banach space C(K)C(K) with few operators. Topology Appl., 143(1-3):217–239, 2004.
  • [41] M. Rørdam, F. Larsen, and N. Laustsen. An introduction to KK-theory for CC^{*}-algebras, volume 49 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 2000.
  • [42] Th. Schlumprecht. An arbitrarily distortable Banach space. Israel J. Math., 76(1-2):81–95, 1991.
  • [43] M. Tarbard. Hereditarily indecomposable, separable \mathscr{L}_{\infty} Banach spaces with 1\ell_{1} dual having few but not very few operators. J. Lond. Math. Soc. (2), 85(3):737–764, 2012.
  • [44] M. Tarbard. Operators on banach spaces of Bourgain-Delbaen type. ProQuest LLC, Ann Arbor, MI, 2013. Thesis (D.Phil.)–University of Oxford (United Kingdom).
  • [45] B. Yood. Difference algebras of linear transformations on a Banach space. Pacific J. Math., 4:615–636, 1954.
  • [46] D. Zisimopoulou. Bourgain-Delbaen \mathcal{L}^{\infty}-sums of banach spaces. arxiv:1402.6564.
  • [47] A. Zsák. A Banach space whose operator algebra has K0K_{0}-group \mathbb{Q}. Proc. London Math. Soc. (3), 84(3):747–768, 2002.