\(2\)-rotund norms for generalized
Baernstein spaces and their duals


Abstract

We consider a generalized Baernstein space associated to a compact family of finite subsets of an uncountable set. We show that for certain transfinitely defined families such spaces admit an equivalent \(2\)-rotund norm. We also show that for an arbitrary family the dual space admits an equivalent \(2\)-rotund norm.

1

1 Introduction↩︎

The notions of \(2\)-rotund and weakly \(2\)-rotund norms were introduced by Milman [1] and are defined as follows.

Definition 1. Let \(X\) be a Banach space. We say that a norm \(\|\cdot\|\) on \(X\) is \(2\)-rotund (\(2R\)) (resp.  weakly \(2\)-rotund (\(W2R\))) if for every \((x_n) \subset X\) such that \(\|x_n\| \le 1\) (\(n \ge 1\)) and \[\lim_{m,n \rightarrow \infty} \|x_m + x_n\| = 2,\] there exists \(x \in X\) such that \(x = \lim_{n \rightarrow \infty} x_n\) strongly (resp.  weakly).

It follows from a characterization of reflexivity due to James [2] that if \(X\) admits an equivalent \(W2R\) norm then \(X\) is reflexive. Hájek and Johanis proved the converse: every reflexive Banach space admits an equivalent \(W2R\) norm [3]. Odell and Schlumprecht [4] proved that every separable reflexive Banach space \(X\) admits an equivalent \(2R\) norm (cf. [5]). However, it is an open question whether every reflexive Banach space admits an equivalent \(2R\) norm.

Let \(\Gamma\) be an infinite set. Throughout, \(\mathcal{F}\) denotes a collection of finite subsets of \(\Gamma\) satisfying the following:

  • \(\mathcal{F}\) contains all singletons;

  • \(\mathcal{F}\) is hereditary, i.e., if \(F \in \mathcal{F}\) and \(G \subseteq F\) then \(G \in \mathcal{F}\);

  • \(\mathcal{F}\) is compact, i.e., \(\{ 1_F \colon F \in \mathcal{F}\}\) is a compact subset of \(\{0,1\}^{\Gamma}\) in the topology of pointwise convergence.

Let \((e_\gamma)_{\gamma \in \Gamma}\) denote the unit vector basis of \(c_{00}(\Gamma)\) and let \((e_\gamma^*)\) denote the dual basis. We define a norm \(\|\cdot\|\) on \(c_{00}(\Gamma)\) as follows: \[\label{eq:32Baernsteinnorm} \|\sum a_\gamma e_\gamma \| = \sup (\sum_{i=1}^n (\sum_{\gamma \in F_i} |a_\gamma|)^2)^{1/2},\tag{1}\] where the supremum is taken over all \(n \ge 1\) and all disjoint \(F_i \in \mathcal{F}\) (\(1 \le i \le n\)).

The generalized Baernstein space \((B(\mathcal{F}),\|\cdot\|)\) is the completion of \(c_{00}(\Gamma)\) with respect to \(\|\cdot\|\). Note that \((e_\gamma)_{\gamma \in \Gamma}\) is a \(1\)-unconditional basis of \(B(\mathcal{F})\) and that \(\|\cdot\|\) satisfies a lower \(2\)-estimate for disjointly supported vectors \(x,y\): \[\label{eq:32lower2estimate} \|x + y\|^2 \ge \|x\|^2 + \|y\|^2.\tag{2}\]

The first space of this type was introduced by Baernstein [6] with \(\Gamma= \mathbb{N}\) and \(\mathcal{F}= \mathcal{S}_1= \{E \subset \mathbb{N}\colon |E| \le \min E\}\) (the Schreier family), with the extra assumption that \(\max F_i < \min F_{i+1}\) for \(1 \le i \le n-1\) in 1 . It was the first example of a reflexive Banach space with a normalized basis (weakly null by reflexivity) whose arithmetic means do not converge strongly to zero.

The space \(B(\mathcal{F})\) is reflexive (for arbitrary \(\Gamma\) and \(\mathcal{F}\)). For completeness we present a proof at the end of the paper.

The norm of \((B(\mathcal{F},\|\cdot\|)\) and its dual norm \(\|\cdot\|_*\) are not \(2R\) in general. For example, for the original Baernstein space, we have \[\|e_n + e_m\| =2,\quad \|e^*_3 + e_n^*\|_*=1, \|(e_3^* + e_n^* )+(e_3^* + e_m^*)\|_* = 2\quad (m,n > 3),\] and hence \(\|\cdot\|\) and \(\|\cdot\|_*\) are not \(2R\) norms.

The following question is open to the best of our knowledge.

Question 1. Suppose \(\Gamma\) is uncountable. Does \(B(\mathcal{F})\) have an equivalent \(2R\) norm?

In Section 2, motivated by the Schreier hierarchy introduced in [7], we present a general method for defining, for each countable ordinal \(\alpha\), a family \(\mathcal{F}_\alpha\) for certain uncountable \(\Gamma\). The construction is similar to that of the transfintely defined families introduced in [8]. In Section 3 we prove that, for each countable ordinal \(\alpha\), \(B(\mathcal{F}_\alpha)\) has an equivalent \(2R\) norm.

In Section 4 we prove, for arbitrary \(\Gamma\) and \(\mathcal{F}\), that \(B(\mathcal{F})^*\) admits an equivalent \(2R\) norm. The renorming is essentially the same as the \(W2R\) renorming given in [3].

As an application of these results we prove that the space constructed by Kutzarova and Troyanski [9] (based on a family of sets introduced in [10]) which does not admit an equivalent norm that is either uniformly rotund in every direction or uniformly differentiable in every direction does admit an equivalent \(2R\) norm.

In forthcoming articles we prove positive results for other classes of spaces. In particular, in [11] we consider the existence of equivalent symmetric \(2R\) norms for spaces with a symmetric basis.

2 Transfinitely defined families↩︎

  • Let \(S\) be any set of cardinality at least \(2\) and let \(\overline{S} := S^{\mathbb{N}}\).

  • For distinct \(p = (p(i))_{i=1}^\infty \in \overline{S}\) and \(q = (q(i))_{i=1}^\infty \in \overline{S}\), let \(d(p,q) = 1\) if \(p(1) \ne q(1)\) and, for \(k \ge 2\), let \(d(p,q) = k\) if \(p(k) \ne q(k)\) and \(p(j) = q(j)\) for \(1 \le j \le k-1.\)

  • For \(A \subset \overline{S}\), with \(|A| \ge 2\), let \[A^\sharp = \min \{d(p,q) \colon p,q \in A, p \ne q\}.\] We define, for each countable ordinal \(\alpha\), a hereditary family \(\mathcal{F}_\alpha\) of finite subsets of \(\overline{S}\).

  • Let \[\mathcal{F}_0 = \{ \emptyset \} \cup \{\{p\} \colon p \in \overline{S}\}.\]

  • If \(k \ge 1\) and \(\mathcal{F}\) is any collection of finite subsets of \(\overline{S}\) satisfying the conditions set out in the Introduction, let \[\mathcal{F}^{(k)} = \mathcal{F}_0 \cup \{A \in \mathcal{F} \colon A^{\sharp} \ge k\}.\] Note that since \(\mathcal{F}\) is hereditary, \(\mathcal{F}^{(k)}\) is also hereditary.

  • If \(\alpha = \beta^+\) is a successor ordinal, let \(\mathcal{F}_\alpha\) be any hereditary family satisfying the following:

    • \(\mathcal{F}_\beta \subseteq \mathcal{F}_\alpha\).

    • If \(A \in \mathcal{F}_\alpha\) and \(|A| \ge 2\), then there exist \(A_i \in \mathcal{F}_\beta\) (\(1 \le i \le A^{\sharp}\)) such that \[A = \cup_{i=1}^{A^{\sharp}} A_i.\]

  • If \(\alpha\) is a limit ordinal, choose \(\alpha_r \uparrow \alpha\) (\(r \ge 1\)) and define \[\mathcal{F}_\alpha = \cup_{r=1}^\infty \mathcal{F}^{(r)}_{\alpha_r}.\] Note that, for each \(k \ge 1\), \[\mathcal{F}^{(k)}_\alpha = \cup_{r=1}^\infty \mathcal{F}^{(r \vee k)}_{\alpha_r},\] where \(r \vee k := \max(r,k)\).

3 \(B(\mathcal{F}_\alpha^{(k)})\) admits an equivalent \(2R\) norm↩︎

Theorem 2. For each countable ordinal \(\alpha\) and \(k \ge 1\), \(B(\mathcal{F}_\alpha^{(k)})\) admits a \(2R\) renorming.

We shall use the following characterization of \(2\)-rotundity (see e.g., [12] or [3]): \(\|\cdot\|\) is a \(2R\) norm on \(X\) if for all \((x_n) \subset X\) such that \[\label{eq:32alternativedef} \lim_{m,n \rightarrow \infty} [\| x_m + x_n\|^2 - 2(\|x_m\|^2 + \|x_n\|^2)] = 0,\tag{3}\] there exists \(x \in X\) such that \(x=\lim_{n \rightarrow \infty} x_n\) strongly.

For \(x \in B(\mathcal{F})\), the support of \(x\), denoted \(\operatorname{supp} x\), is defined by \[\operatorname{supp} x = \{\gamma \in \overline{S}: e_\gamma^*(e_\gamma) \ne 0\}.\]

Let \(\|\cdot\|_{\alpha,k}\) denote the norm in \(B(\mathcal{F}_\alpha^{(k)}).\)

Lemma 1. Let \(\alpha\) be a limit ordinal (with \(\alpha_r \uparrow \alpha\) as above) and let \(k \ge 1\). Suppose that \(\|x_n\|_{\alpha, k} \le 1\) (\(n\ge1\)) and that \[\label{eq:32lim95m95n} \lim_{m,n\rightarrow \infty} \|x_m + x_n\|_{\alpha,k} = 2.\tag{4}\] Then, for some \(r \ge 1\), \(\limsup_{n\rightarrow\infty}\|x_n\|_{\alpha_r, r \vee k} >0\)

Proof. Suppose, to derive a contradiction, that \(\lim_{n \rightarrow \infty}\|x_n\|_{\alpha_r, r \vee k} =0\) for all \(r \ge 1\); in particular, \(x_n \rightarrow 0\) in \(\ell_2(\overline{S})\). Hence, by a gliding hump argument, approximating by finitely disjointly supported vectors, and after passing to a subsequence and relabelling, we may assume that \(\operatorname{supp} x_n\) is finite and that \(\operatorname{supp} x_n \cap \operatorname{supp} x_m = \emptyset\) if \(m \ne n\).

Fix \(n \ge 1\) and \(F \in \mathcal{F}^{(k)}_\alpha\) satisfying \[|F \cap \operatorname{supp} x_n| \ge 2.\] Let \[N = \max \{ d(p,q) \colon p,q \in \operatorname{supp} x_n, p \ne q\}.\] It follows that \(F^\sharp \le N\), and hence \[F \in \cup_{r=1}^N \mathcal{F}_{\alpha_r}^{(r \vee k)}.\]

Let \[x_n = \sum a_\gamma e_\gamma\] and, for \(m > n\), \[x_m = \sum b^m_\gamma e_\gamma.\] Since \(x_n \rightarrow 0\) in \(\ell_2(\overline{S})\), \[\label{eq:32sumsofsquares} \lim_{n \rightarrow \infty} \sum a_\gamma^2 =0 .\tag{5}\] Since the supports of the \(x_m\)’s are disjoint, we may assume that \(a_\gamma\ge0\) and \(b^m_\gamma\ge0\).

By assumption, \(\|x_m\|_{\alpha_r, r \vee k} \rightarrow 0\) as \(m \rightarrow \infty\) for all \(r \ge 1.\) Hence \[\label{eq:32uniform1} \lim_{m \rightarrow \infty} \sum_{\gamma \in F} b^m_\gamma = 0.\tag{6}\] uniformly over all \(F \in \mathcal{F}^{(k)}_\alpha\) satisfying \(|F \cap \operatorname{supp} x_n| \ge 2\).

Note that if \(F_1,F_2,\dots,F_s\) are disjoint sets in \(\mathcal{F}^{(k)}_\alpha\) satisfying \(|F_i \cap \operatorname{supp} x_n| \ge 2\) (\(1 \le i \le s\)) then \(s \le |\operatorname{supp} x_n|\). Hence 6 implies that \[\label{eq:32uniform2} \sum_{\gamma \in \cup_{i=1}^s F_i} b^m_\gamma \rightarrow 0\tag{7}\] as \(m \rightarrow \infty\) uniformly over all such collections \((F_i)_{i=1}^s\). Let \(A_i = \sum_{\gamma \in F_i} a_\gamma\) and let \(B^m_i = \sum_{\gamma \in F_i} b^m_\gamma\). Then \[\begin{align} \sum_{i=1}^s (A_i + B^m_i)^2 &= \sum_{i=1}^s (A_i^2 +( B^m_i)^2 + 2A_iB^m_i) \\ &\le \sum_{i=1}^s A_i^2 + (\sum_{i=1}^s B^m_i)^2 + 2(\sum_{i=1}^s B^m_i) (\sum A_i^2)^{1/2}\\ &\le \sum_{i=1}^s A_i^2 + (\sum_{i=1}^s B^m_i)^2 + 2\|x_n\|_{\alpha,k} \sum_{i=1}^s B^m_i\\ & \le \sum_{i=1}^s A_i^2 + (\sum_{i=1}^s B^m_i)^2 + 2 \sum_{i=1}^s B^m_i. \end{align}\] Note that 7 implies that \(\sum_{i=1}^s B^m_i \rightarrow 0\) as \(m \rightarrow \infty\) uniformly over all such \((F_i)_{i=1}^s\). Let \(\varepsilon>0\). It follows that for all \(m \ge M(n, \varepsilon)\), \[\label{eq:32atleast2} \sum_{i=1}^s (A_i + B^m_i)^2 < \sum_{i=1}^s A_i^2 + \varepsilon \le \|x_n\|_{\alpha_k} + \varepsilon \le 1 + \varepsilon\tag{8}\] uniformly over all \((F_i)_{i=1}^s\). Moreover, it follows from 5 that for all \(n \ge N(\varepsilon)\) \[\sum a_\gamma^2 < \varepsilon^2.\] Let \(J\subset \operatorname{supp} x_n\). Consider disjoint sets \(G_\lambda \in \mathcal{F}^{(k)}_\alpha\) (\(\lambda \in J\)) satisfying \(G_\lambda \cap \operatorname{supp} x_n = \{\lambda\}\) (\(\lambda \in J\)). Let \(C^m_\lambda = \sum_{\gamma \in G_\lambda} b^m_\gamma\). Then for all \(m> n> N(\varepsilon)\), \[\label{eq:32exactlyone} \begin{align} \sum_{\lambda \in J} (a_\lambda + C^m_\lambda)^2 &\le \sum_{\lambda \in J} a_\gamma^2 + \sum_{\lambda \in J}(C^m_\lambda)^2 + 2( \sum_{\lambda \in J} a_\lambda^2)^{1/2} (\sum_{\lambda\in J}(C^m_\lambda)^2)^{1/2}\\ &\le \varepsilon + \sum_{\lambda \in J}(C^m_\lambda)^2 + 2\varepsilon \|x_m\|_{\alpha,k}\\ &\le \varepsilon + 2\varepsilon + \|x_m\|_{\alpha,k}^2 \\ &\le 1 + 3\varepsilon. \end{align}\tag{9}\] Hence, combining 8 and 9 , for all \(n \ge N(\varepsilon)\) and \(m> M(n,\varepsilon)\), \[\|x_n + x_m\|_{\alpha,k}^2 \le 2 + 4\varepsilon.\] Since \(\varepsilon>0\) is arbitrary, we have \[\label{eq:32sqrt2} \limsup_{n \rightarrow \infty} \limsup_{m \rightarrow \infty} \|x_n + x_m\|_{\alpha,k} \le \sqrt{2},\tag{10}\] which contradicts 4 . ◻

The following analogue for successor ordinals has a similar (but simpler) proof.

Lemma 2. Let \(\alpha = \beta^+\) be a successor ordinal. Suppose that \(\|x_n\|_{\alpha, k} \le 1\) (\(n\ge1\)) and that \[\label{eq:32lim95m95n95295successor} \lim_{m,n\rightarrow \infty} \|x_m + x_n\|_{\alpha,k} = 2.\tag{11}\] Then \[\limsup_{n \rightarrow \infty} \|x_n\|_{\beta,k} >0.\]

Remark 3. 10 shows that Lemma 1 and Lemma 2 can be strengthened by replacing 4 and 11 by \[\limsup_{n \rightarrow \infty} \limsup_{m \rightarrow \infty} \|x_n + x_m\|_{\alpha,k} > \sqrt{2}.\]

The proof of the following lemma uses the fact that Hilbert space \((\ell_2,|\cdot|)\) is uniformly convex; specifically, for \(0< \varepsilon < 2\), \[|x| \le1, |y| \le 1, |x-y| = \varepsilon \Rightarrow |\frac{x+y}{2}| \le 1 - \frac{\varepsilon^2}{8}.\] We will also use the following notation: for \(x =\sum_{\gamma \in \overline{S}} x_\gamma e_\gamma\) and disjoint sets \(F_i \subset \overline{S}\) (\(1 \le i \le n\)), \[|(x; F_1,\dots,F_n)|_2:=(\sum_{i=1}^n (\sum_{\gamma \in F_i} x_\gamma)^2)^{1/2}.\] Note that if \(x \ge 0\), then \[\|x\|_{\alpha,k} = \sup |(x; F_1,\dots,F_n)|_2,\] where the supremum is taken over all \(n \ge 1\) and disjoint \(F_i \in \mathcal{F}^{(k)}_\alpha\).

Lemma 3. Let \(\alpha\) be a limit ordinal (with \(\alpha_r \uparrow \alpha\) as above) and let \(k \ge 1\). Suppose that \(\|x_n\|_{\alpha, k} \le 1\) (\(n\ge1\)), that \[\label{eq:32lim95m95n952} \lim_{m,n\rightarrow \infty} \|x_m + x_n\|_{\alpha,k} = 2,\tag{12}\] and that there exists \(x \in \ell_2(\overline{S})\) such that, for each \(r \ge 1\), \[\lim_{n \rightarrow \infty} \|x_n - x\|_{\alpha_r,r \vee k} = 0.\] Then \(\lim_{n \rightarrow \infty} \|x_n - x\|_{\alpha,k}=0\).

Proof. Note that \[\|x\|_{\alpha,k} \le \limsup_{n \rightarrow \infty} \|x_n\|_{\alpha,k} \le 1,\] since \(x_n \rightarrow x\) pointwise. Suppose, to derive a contradiction, that the conclusion is false. Then, after passing to a subsequence and relabelling, we may assume that \[\lim_{n \rightarrow\infty}\|x_n - x\|_{\alpha,k} = \delta > 0.\] Let \(x_n^\prime= x_n - x\). By assumption, for all \(r \ge 1\), \[\lim_{n \rightarrow \infty}\|x_n^\prime\|_{\alpha_r, r\vee k} = 0.\] Let \(\varepsilon>0\). Choose a finitely supported vector \(y\) such that \[\|x - y\|_{\alpha,k} < \frac{\varepsilon^2}{10}.\] By a gliding hump argument, passing to a further subsequence and relabelling, we may choose disjointly supported vectors \(y_n\) (\(n \ge 1\)), each with finite support disjoint from the support of \(y\), such that \(\|x_n^\prime - y_n\|_{\alpha,k} \rightarrow 0\) as \(n \rightarrow \infty\) and, for all \(m,n \ge 1\), \[\|y + y_n||_{\alpha,k} \le 1,\] and also \[\|2y + y_n + y_m \|_{\alpha,k} > 2 -\frac{\varepsilon^2}{4}.\] Hence \[\lim_{n \rightarrow \infty}\| y_n\|_{\alpha,k} = \delta,\] and, for all \(r \ge 1\), \[\label{eq:3232nullsequence} \lim_{n \rightarrow \infty}\| y_n\|_{\alpha_r, r \vee k} =\lim_{n \rightarrow \infty}\|x_n^\prime\|_{\alpha_r, r\vee k} =0.\tag{13}\] Without loss of generality, we may assume that \(y\ge0\) and \(y_n \ge 0\) for all \(n \ge 1\). Fix \(n \ge 1\) and let \(m>n\). Suppose that \(2y + y_n + y_m\) is normed by disjoint sets \(F_1,\dots,F_u\) in \(\mathcal{F}_{\alpha,k}\) (we suppress the dependence of \(F_i\) on \(n\) and \(m\) to simplify notation), i.e., \[|(2y + y_n + y_m; F_1,\dots,F_u)|_2 = \|2y + y_n + y_m \|_{\alpha,k}> 2 -\frac{\varepsilon^2}{4}.\] Since \[|(y+y_n; F_1,\dots,F_u)|_2 \le \|y+ y_n\|_{\alpha,k}\le 1\] and \[|(y+y_m; F_1,\dots,F_u)|_2 \le \|y+ y_m\|_{\alpha,k}\le 1,\] the uniform convexity of \(\ell_2\) yields \[|(y_n-y_m; F_1,\dots,F_u)|_2 < \varepsilon.\]

We may assume that \(F_1,\dots,F_s\) have nonempty intersection with both \(\operatorname{supp} y\) and \(\operatorname{supp} y_n\), that \(F_{s+1},\dots,F_t\) intersect \(\operatorname{supp} y\) but not \(\operatorname{supp} y_n\), and that \(F_{t+1},\dots,F_u\) do not intersect \(\operatorname{supp} y\). Note that \(s \le |\operatorname{supp} y|\) and \(|F_i \cap \operatorname{supp} (y + y_n)| \ge 2\) for \(1 \le i \le s\). Hence, repeating the argument used to prove 7 , we deduce that \[\label{eq:32uniform3} \lim_{m\rightarrow\infty} \sum_{\gamma \in \cup_{i=1}^s F_i} b^m_\gamma = 0\tag{14}\] for \(y_m = \sum b^m_\gamma e_\gamma\). Hence \[|(y_m; F_1,\dots,F_s)|_2 < \frac{\varepsilon}{2}\] for all \(m > M_1(n,\varepsilon)\).

Note that \(y_n\) vanishes on \(F_i\) for \(s+1 \le i \le t\). Hence, for all \(m > M_1(n,\varepsilon)\), \[\begin{align} |(y_n+y_m; F_1,\dots,F_t)|_2 &=(\sum_{i=1}^s (\sum_{\gamma \in F_i} (b^n_\gamma + b^m_\gamma))^2 + \sum_{i = s+1}^t (\sum_{\gamma \in F_i} b^m_\gamma)^2)^{1/2} \\ &\le (\sum_{i=1}^s (\sum_{\gamma \in F_i} (b^n_\gamma -b^m_\gamma))^2 + \sum_{i = s+1}^t (\sum_{\gamma \in F_i} b^m_\gamma)^2)^{1/2} \\ & + 2 (\sum_{i=1}^s (\sum_{\gamma \in F_i} b^m_\gamma)^2)^{1/2}\\ \intertext{(by the triangle inequality in \ell_2)} &= |(y_n-y_m; F_1,\dots,F_t)|_2 + 2|(y_m; F_1,\dots F_s)|_2\\ &\le \varepsilon + \varepsilon = 2\varepsilon. \end{align}\] So \[\begin{align} |(2y + y_n + y_m; F_1,\dots,F_t)|_2 &\le 2|(y;F_1,\dots, F_t)|_2 + |(y_n+y_m;F_1,\dots,F_t)|_2\\ &\le 2\|y\|_{\alpha,k} + 2\varepsilon. \end{align}\] Thus, \[\label{eq:32firstest}\begin{align} (2 - \frac{\varepsilon^2}{4})^2 &< |(2y+y_n+y_m; F_1,\dots,F_u)|_2^2\\ &= |(2y+y_n+y_m; F_1,\dots,F_t)|_2^2 + |(y_n+y_m; F_{t+1},\dots,F_u)|_2^2\\ \intertext{(since y vanishes on F_i for t+1 \le i \le u)} &\le ( 2\|y\|_{\alpha,k} + 2\varepsilon)^2 + \|y_n + y_m\|_{\alpha,k}^2. \end{align}\tag{15}\] Since \(y\), \(y_n\), and \(y_m\) are disjointly supported, we have \[\label{eq:32secondest} \|y\|_{\alpha,k}^2 + \|y_n\|_{\alpha,k}^2 \le \|y+y_n\|_{\alpha,k}^2 \le 1\tag{16}\] and \[\label{eq:32thirdest} \|y\|_{\alpha,k}^2 + \|y_m\|_{\alpha,k}^2 \le \|y+y_m\|_{\alpha,k}^2 \le 1.\tag{17}\] Combining 15 , 16 , and 17 , \[\begin{align} 4\|y\|_{\alpha,k}^2 +2(\|y_n\|_{\alpha,k}^2 + \|y_m\|_{\alpha,k}^2) &\le 4\\ & = (2 - \frac{\varepsilon^2}{4})^2 + \varepsilon^2 -\frac{\varepsilon^4}{16}\\ &\le ( 2\|y\|_{\alpha,k} + 2\varepsilon)^2 + \|y_n + y_m\|_{\alpha,k}^2 + \varepsilon^2 \\ &\le 4\|y\|_{\alpha,k}^2 + \|y_n + y_m\|_{\alpha,k}^2 +(8\varepsilon + 5\varepsilon^2). \end{align}\] Hence for all \(m>M_1(n,\varepsilon)\), \[\label{eq:328epsilon} \|y_n + y_m\|_{\alpha,k}^2 +(8\varepsilon + 5\varepsilon^2) \ge 2(\|y_n\|_{\alpha,k}^2 + \|y_m\|_{\alpha,k}^2).\tag{18}\] Now suppose \(\varepsilon\) is chosen so that \(8\varepsilon + 5\varepsilon^2 < 2\delta^2\). Since \(\lim_{n \rightarrow \infty} \|y_n\|_{\alpha,k} = \delta\), it follows from 18 that \[\liminf_{n \rightarrow \infty} \liminf_{m \rightarrow\infty} \|y_n + y_m\|_{\alpha,k} > \sqrt2 \delta.\] which contradicts Remark 3 since, for all \(r \ge 1\), \[\lim_{n \rightarrow \infty} \|y_n\|_{\alpha_r, r \vee k} = 0.\] ◻

The following analogue for successor ordinals has a similar (but simpler) proof.

Lemma 4. Let \(\alpha = \beta^+\) be a successor ordinal. Suppose that \(\|x_n\|_{\alpha, k} \le 1\) (\(n\ge1\)), that \[\label{ymxldhrn} \lim_{m,n\rightarrow \infty} \|x_m + x_n\|_{\alpha,k} = 2,\tag{19}\] and that there exists \(x \in \ell_2(\overline{S})\) such that \[\lim_{n \rightarrow \infty} \|x_n - x\|_{\beta, k} = 0.\] Then \(\lim_{n \rightarrow \infty} \|x_n - x\|_{\alpha,k}=0\).

Proof of Theorem 2. We will prove the result for a fixed \(\alpha\) and for all \(k \ge 1\) by transfinite induction on \(\alpha\). The result clearly holds for \(\alpha = 0\) since \(B(\mathcal{F}_0^{(k)}) = B(\mathcal{F}_0)= \ell_2(\overline{S})\) for all \(k \ge 1\). So suppose the result holds for all \(\beta<\alpha\) and for all \(k \ge 1\).

Case I: \(\alpha\) is a limit ordinal. So \(\mathcal{F}^{(k)}_\alpha = \cup_{r=1}^\infty \mathcal{F}^{(r \vee k)}_{\alpha_r},\) where \(\alpha_r \uparrow \alpha\). By inductive hypothesis, each \(B(\mathcal{F}^{(r \vee k)}_{\alpha_r})\) admits an equivalent \(2R\) norm \(|\!|\!|\cdot |\!|\!|_{\alpha_r, r \vee k}\). Note that \[|\!|\!|\cdot |\!|\!|_{\alpha_r, r \vee k} \le C_r \|\cdot\|_{\alpha, k}\] for some \(C_r < \infty\). Thus, \[|\!|\!|\cdot |\!|\!|_{\alpha,k}^2 := \|\cdot\|_{\alpha,k}^2 + \sum_{r=1}^\infty \frac{1}{2^rC_r^2} |\!|\!|\cdot |\!|\!|_{\alpha_r, r \vee k}^2\] defines an equivalent norm \(|\!|\!|\cdot |\!|\!|_{\alpha,k}\) on \(B(\mathcal{F}_\alpha^{(k)})\). Let us show that \(|\!|\!|\cdot |\!|\!|_{\alpha,k}\) is a \(2R\) norm. Suppose that \((x_n) \subset B(\mathcal{F}_\alpha^{(k)})\) satisfies \[\lim_{m,n \rightarrow \infty} |\!|\!|x_n + x_m |\!|\!|_{\alpha,k}^2 - 2( |\!|\!|x_n |\!|\!|_{\alpha,k}^2 + |\!|\!|x_m| |\!|\!|_{\alpha,k}^2) = 0.\] Note that \[\begin{align} & |\!|\!|x_n + x_m |\!|\!|_{\alpha,k}^2 - 2( |\!|\!|x_n |\!|\!|_{\alpha,k}^2 + |\!|\!|x_m |\!|\!|_{\alpha,k}^2)\\ &\le -(\|x_n\|_{\alpha,k} -\| x_m\|_{\alpha,k})^2 - \sum_{r=1}^\infty \frac{1}{2^rC_r^2} ( |\!|\!|x_n |\!|\!|_{\alpha_r,k} - |\!|\!| x_m |\!|\!|_{\alpha+r,k})^2. \end{align}\] It follows that \(\lim_{n \rightarrow \infty} \|x_n\|_{\alpha,k} = L\) for some \(L \ge 0\), that \[\label{eq:32limitalphak} \lim_{m,n \rightarrow \infty} \|x_n + x_m\|_{\alpha,k}^2 - 2(\|x_n\|_{\alpha,k}^2 + \|x_m\|_{\alpha,k}^2) = 0,\tag{20}\] and that, for all \(r\ge1\), \[\lim_{m,n \rightarrow \infty} |\!|\!|x_n + x_m |\!|\!|_{\alpha_r,k \vee r}^2 - 2( |\!|\!|x_n |\!|\!|_{\alpha_r,k \vee r}^2 + |\!|\!|x_m |\!|\!|_{\alpha_r,k \vee r}^2) = 0.\] Since each \(|\!|\!|\cdot |\!|\!|_{\alpha_r, r \vee k}\) is a \(2R\) norm, it follows from 3 that there exists \(x \in \ell_2(\overline{S})\) such that, for all \(r\ge 1\), \[\lim_{n \rightarrow \infty} |\!|\!|x_n - x |\!|\!|_{\alpha_r, r \vee k} = 0.\] Moreover, 20 implies that \[\lim_{m,n \rightarrow \infty}\| x_n + x_m \|_{\alpha,k} = 2L.\] So, by Lemma 3, \[\lim_{n \rightarrow \infty} \| x_n - x \|_{\alpha,k} = 0,\] and hence \[\lim_{n \rightarrow \infty} |\!|\!| x_n - x |\!|\!|_{\alpha,k} = 0,\] as desired.

Case II: \(\alpha= \beta^+\) is a successor ordinal. The proof is very similar to the limit ordinal case. By the inductive hypothesis, \(B(\mathcal{F}_\beta^{(k)})\) admits an equivalent \(2R\) norm \(|\!|\!| \cdot |\!|\!|_{\beta,k}\). Let \[|\!|\!| \cdot |\!|\!|^2_{\alpha, k}= \|\cdot\|_{\alpha,k}^2 + |\!|\!| \cdot |\!|\!|^2_{\beta,k}.\] Using Lemma 4 instead of Lemma 3 and repeating the argument of Case I shows that \(|\!|\!| \cdot |\!|\!|_{\alpha, k}\) is a \(2R\) norm. ◻

4 \(B(\mathcal{F})^*\) admits an equivalent \(2R\) norm↩︎

Let \(\mathcal{F}\) be a compact, hereditary family of finite subsets of an infinite set \(\Gamma\) containing all singleton sets. We prove in Section 5 that \((B(\mathcal{F}),\|\cdot\|)\) is reflexive. Day [13] introduced the norm \(\|\cdot \|_{\textrm{Day}}\) on \(c_0(\Gamma)\) defined by \[\|\sum a_\gamma e_\gamma\|_{\textrm{Day}} = \sup (\sum_{i=1}^n 4^{-i} |a_{\gamma_i}|^2)^{1/2},\] where the supremum is taken over all \(n \ge 1\) and all choices of distinct \(\gamma_i\in \Gamma\) (\(1 \le i \le n\)). We define an equivalent norm on \(B(\mathcal{F})^*\) thus: \[|\!|\!|x \; |\!|\!|^2 = \|x\|_*^2 + \|x \|_{\textrm{Day}}^2 \qquad (x \in B(\mathcal{F})^*).\]

The following result is essentially due to Hájek and Johannis. It is a consequence of Theorem 3 and Corollary 4 of [3] and the reflexivity of \(B(\mathcal{F})^*\).

Lemma 5. Suppose \((y_n) \subset B(\mathcal{F})^*\) satisfies \[\label{eq:32triplenormcondition} \lim_{m,n \rightarrow \infty} |\!|\!|y_n + y_m |\!|\!|^2- 2(|\!|\!|y_n|\!|\!|^2 + |\!|\!|y_m|\!|\!|^2) = 0.\tag{21}\] Then there exists \(y \in B(\mathcal{F})^*\) such that \[y_n\rightarrow y\quad \text{weakly as n \rightarrow \infty}\] and \[\lim_{n \rightarrow \infty} \|y_n - y\|_\infty = 0.\]

Dualizing 2 , the dual space \((B(\mathcal{F})^*,\|\cdot\|_*)\) satisfies an upper \(2\)-estimate for disjointly supported vectorrs \(x,y \in B(\mathcal{F})^*\): \[\|x + y\|_*^2 \le \|x\|_*^2 + \|y\|_*^2.\] Moreover, for all \(x \in B(\mathcal{F})^*\) and \(F \in \mathcal{F}\), \[\|x\cdot 1_F\|_* = \|x\cdot 1_F\|_\infty \le \|x\|_\infty.\]

Lemma 6. Suppse that \(y\) and \(y_n\) have disjoint finite supports (\(n \ge 1\)), that \[\|y\|_* = \|y_n\|_* =1 \qquad(n \ge 1),\] and that \[\lim_{n \rightarrow \infty} \|y_n\|_\infty = 0.\] Then, for all \(\delta > 0\), \[\lim_{n \rightarrow \infty} \|y+\delta y_n\|_* = (1 + \delta^2)^{1/2}.\]

Proof. We may assume that \(y \ge0\) and \(y_n \ge.\) Choose positive norming vectors \(x, x_n \in B(\mathcal{F})\) with \[(x,y) = \|x\| = \|y\|_* =1, \quad (x_n,y_n)=\|x_n\|= \|y_n\|_*=1,\] where \((\cdot,\cdot)\) denotes the duality pairing for \(B(\mathcal{F})\times B(\mathcal{F})^*\). Note that \(x\) and \(x_n\) have disjoint finite supports (\(n \ge 1\)). Fixing \(n \ge 1\), choose disjoint \(F_i \in \mathcal{F}\) (\(1 \le i \le N\)) such that \[\|x + \delta x_n\| = |(x+ \delta x_n; F_1,\dots,F_N)|_2.\] We may assume that only \(F_1,\dots,F_k\) have non-empty intersection with both \(\operatorname{supp} x\) and \(\operatorname{supp} x_n\). Note that \[k \le M:= |\operatorname{supp} x|.\] For each \(1 \le i \le k\), \[\|y_n \cdot 1_{F_i}\|_* \le \|y_n\|_\infty.\] Hence \[\sum_{i=1}^k \|y_n \cdot 1_{F_i}\|_* \le M \|y_n\|_\infty \rightarrow 0 \quad\text{as n \rightarrow \infty}.\] Let \(F = \cup_{i=1}^k F_i\). (To simplify notation we suppress the dependence of \(F\) on \(n\).) Then \[\begin{align} \|x_n - x_n \cdot 1_F\| &\ge (x_n -x_n\cdot 1_F, y_n)\\ &= (x_n,y_n -y_n\cdot 1_F)\\ &= 1 - (x_n, y_n\cdot 1_F)\\ &\ge 1 - \sum_{i=1}^k \|y_n \cdot 1_{F_i}\|_*\\ &\rightarrow 1 \quad \text{as n \rightarrow \infty.} \end{align}\]Hence \[\lim_{n \rightarrow \infty} (\|x_n\|^2 - \|x_n - x_n\cdot 1_F\|^2)= 1 - \lim_{n \rightarrow \infty} \|x_n - x_n\cdot 1_F\|^2=0.\] Since \((B(\mathcal{F},\|\cdot\|)\) satisfies a lower \(2\)-estimate, it follows that \(\|x_n \cdot 1_F\| \rightarrow 0\) as \(n \rightarrow \infty\). Hence \[\begin{align} 1 + \delta^2 &\le \liminf_{n \rightarrow \infty}\|x + \delta x_n\|^2\\ & \le \limsup_{n \rightarrow \infty}\|x + \delta x_n\|^2\\ & = \limsup_{n \rightarrow \infty} |(x+\delta x_n; F_1,\dots F_N)|_2^2\\ &= \limsup_{n \rightarrow \infty} |(x+ \delta x_n - \delta x_n \cdot 1_F; F_1,\dots,F_N)|_2^2\\ & \le \limsup_{n \rightarrow \infty}[ \|x\|^2+ \delta^2 \| x_n - x_n \cdot 1_F\|^2]\\ \intertext{(since no F_i intersects both \operatorname{supp} x and \operatorname{supp} (x_n - x_n \cdot 1_F))} &= 1 + \delta^2. \end{align}\] Thus, \[\lim_{n \rightarrow \infty}\|x+ \delta x_n\| ^2= 1 + \delta^2,\] and hence \[\begin{align} (1 + \delta^2)^{1/2} &\ge \limsup_{n \rightarrow \infty} \|y + \delta y_n\|_*\\ \intertext{(since \|\cdot\|_* satisfies an upper 2-estimate)} &\ge \liminf_{n \rightarrow \infty} \|y + \delta y_n\|_*\\ & \ge \liminf_{n \rightarrow \infty} \frac{(x + \delta x_n, y + \delta y_n)}{\|x + \delta x_n\|}\\ &= \frac{1 + \delta^2}{(1 + \delta^2)^{1/2}}\\ &=(1 + \delta^2)^{1/2}. \end{align}\] ◻

Theorem 4. \(|\!|\!|\cdot |\!|\!|\) is an equivalent \(2R\) norm for \(B(\mathcal{F})^*\).

Proof. Suppose \((y_n) \subset B(\mathcal{F})^*\) satisfies 21 . By Lemma 5 there exists \(y \in B(\mathcal{F})^*\) such that \(y = w-\lim_{n \rightarrow \infty} y_n\) and \(\lim_{n \rightarrow \infty}\|y_n - y\|_\infty =0\). Suppose, to derive a contradiction, that \((y_n)\) does not converge strongly to \(y\). Passing to a subsequence and relabelling, we may assume that \(y_n = y+z_n\) where \[\lim_{n \rightarrow\infty} \|z_n\|_* = \delta>0, \lim_{n \rightarrow\infty} \|z_n\|_\infty = 0,\] and that the following limits exist: \[\lim_{n \rightarrow\infty} \|y+z_n\|_*, \lim_{n,m \rightarrow\infty} \|2y+z_n+z_m\|_* .\] Let \(\varepsilon>0\). By passing to a further subsequence, a gliding hump argument and the fact that \(\lim_{n \rightarrow\infty} \|z_n\|_\infty = 0\) show that there exist vectors \(y^\prime\) and \(z_n^\prime\) (\(n \ge 1\)) with disjoint finite supports such that \[\label{eq:32approxi} \|y - y^\prime\|_* < \varepsilon, \lim_{n \rightarrow \infty}\| z_n - z_n^\prime\|_* =0.\tag{22}\] Note that 21 implies that \[\label{eq:32limitnorm42} \lim_{m,n \rightarrow \infty} \|2y+ z_n + z_m \|_*^2- 2(\|y+z_n|\|_*^2 + \|y+z_m\|_*^2) = 0.\tag{23}\] Since \(\lim_{n \rightarrow \infty} \|z^\prime_n\|_\infty = 0\), Lemma 6 yields \[\lim_{n \rightarrow \infty} \|y^\prime + z_n^\prime\|_*^2 = \|y^\prime\|_*^2 + \lim_{n \rightarrow \infty}\|z_n^\prime\|_*^2 = \|y^\prime\|_*^2 + \delta^2.\] Since \((B(\mathcal{F},\|\cdot\|_*)\) satisfies an upper \(2\)-estimate, \[\lim_{n,m \rightarrow \infty} \|2y^\prime + z_n^\prime + z_m^\prime \|_*^2 \le 4\|y^\prime\|_*^2 + 2\delta^2.\] Hence \[\begin{align} &\limsup_{n,m \rightarrow \infty}[\|2y^\prime + z_n^\prime + z_m^\prime \|_*^2 - 2(\|y^\prime + z_n^\prime\|_*^2 + \|y^\prime + z_m^\prime\|_*^2)]\\ &\le 4\|y^\prime\|_*^2 + 2\delta^2 - 2(2\|y^\prime\|_*^2 + 2\delta^2)\\ &= -2\delta^2, \end{align}\] which contradicts 22 and 23 provided \(\varepsilon\) is sufficiently small. ◻

Based on a family of sets introduced in [10], Kutzarova and Troyanski [9] constructed a Banach space \(Y\) which does not admit an equivalent norm that is uniformly rotund or uniformly differentiable in every direction. As an application of our results, we show that \(Y\) does admit an equivalent \(2R\) norm.

Corollary 1. The Banach space \(Y\) defined in [9] admits an equivalent \(2R\) norm.

Proof. The space \(Y\) is defined as \(X \oplus X^*\), where \(X\) is defined below.

Let \(S = \mathbb{N}\). Let \(\mathcal{F}_1\) be the collection of all finite subsets \(F\) of \(\overline{S}\) such that, if \(|F| \ge 2\), then for all \(p \in F\), \(p(1)=1\) and \(p(i) \in \{1,2,\dots,i-1\}\) for all \(i \ge 2\) and such that for all distinct \(p,q \in F\), there exists \(m \ge 3\) such that \(p(i) = q(i)\) for all \(1 \le i \le m-1\) and \(p(m) \ne q(m)\), which implies that \(F^\sharp = m-1\) and \(|F| \le m-1\) as required. The space \(X\) is defined to be the closed linear span of \[\{e_p \colon p(1)=1, p(i) \in \{1,2,\dots,i-1\}, i \ge 2\}.\] in \(B(\mathcal{F}_1)\). The successor case of the proof of Theorem 2 shows that \[|\!|\!| \cdot |\!|\!|^2 = \|\cdot\|_1^2 + \|\cdot\|^2_{\ell_2(\overline{S})}\] is an equivalent \(2R\) norm on \(B(\mathcal{F}_1)\). Hence \(|\!|\!| \cdot |\!|\!|\) restricts to an equivalent \(2R\) norm on \(X\). By Theorem 4, \(B(\mathcal{F}_1)^*\) admits an equivalent \(2R\) norm. Note that \(X^*\) is isomorphic to a quotient space of \(B(\mathcal{F}_1)^*\). It is easily seen that a quotient norm of a \(2R\) norm is \(2R\). Hence \(X^*\) admits an equivalent \(2R\) norm, \(|\!|\!| \cdot |\!|\!|^\prime\) say. Finally, \[\|(x,x^*)\| =\sqrt{|\!|\!|x|\!|\!|^2 + |\!|\!| x^* |\!|\!|^{\prime2}}\qquad ( (x,x^*) \in X \oplus X^*)\] is an equivalent \(2R\) norm on \(X \oplus X^* =Y\). ◻

5 \(B(\mathcal{F})\) is reflexive↩︎

Theorem 5. For arbitrary \(\Gamma\) and \(\mathcal{F}\), \(B(\mathcal{F})\) is reflexive.

Proof. First, we consider the case \(\Gamma = \mathbb{N}\). Let \((e_n)\) denote the unit vector basis of \(B(\mathcal{F})\).

Let \((F_i)_{i=1}^\infty \subset \mathcal{F}\) be a collection of disjoint elements of \(\mathcal{F}\) and suppose that \(\sum_{i=1}^\infty |a_i|^2 \le 1\). For \(x = \sum_{i=1}^\infty x_i e_i \in B(\mathcal{F})\), \[\label{eq:32functionalrep} | \sum_{i=1}^\infty a_i(\sum_{j \in F_i} x_j)| \le (\sum_{i=1}^\infty |a_i|^2)^{1/2} \|x\| \le \|x\|.\tag{24}\] Hence we may identify \(\sum_{i=1}^\infty a_i 1_{F_i} \in \ell_\infty\) with the element \(k\) in the unit ball of \(B(\mathcal{F})^*\) defined by 24 .

Suppose \(x \in B(\mathcal{F})\) has finite support and that \[\|x\| = (\sum_{i=1}^n (\sum_{j \in G_i} |x_j|) ^2)^{1/2}\] for disjoint \(G_i \in \mathcal{F}\). There exist nonnegative \(a_1,\dots,a_n\) with \(\sum_{i=1}^n a_i^2 = 1\) such that \[\sum_{i=1}^n a_i (\sum_{j \in G_i} |x_j|) = \|x\|,\] and there exist \(H_i\subseteq G_i\) (\(1 \le i \le n\)) such that \[\sum_{i=1}^n a_i|\sum_{j \in H_i} x_j| \ge \frac{1}{2} \sum_{i=1}^n a_i (\sum_{j \in G_i} |x_j|) = \frac{1}{2} \|x\|.\] Note that \(H_i \in\mathcal{F}\) since \(\mathcal{F}\) is hereditary. Hence the collection of linear functionals with a representation of the form \[k=\sum_{i=1}^\infty a_i 1_{F_i} \in \ell_\infty\qquad (\sum_{i=1}^\infty | a_i|^2 =1, (F_i)_{i=1}^\infty\, \text{disjoint sets in\,} \mathcal{F})\] is a \(2\)-norming set for \(B(\mathcal{F})\). It follows that the discretized collection \[K=\{\sum_{r=1}^\infty \pm 2^{-s(r)} 1_{F_r} \colon \sum_{r=1}^\infty 2^{-2s(r)} \le 1 \}.\] is a \(4\)-norming set.

Let us show that \(K \subset \ell_\infty\) is compact in the topology of pointwise convergence on \(\ell_\infty\). For \(n\ge1\), let \[k_n = \sum_{r=1}^\infty 2^{-r}( 1_{U^n_r} - 1_{V^n_r}),\] where \(U^n_r= \cup_{i=1}^{p(n,r)} F^n_i\) and \(V^n_r=\cup_{i=1}^{q(n,r)} G^n_i\), and for each \(n \ge 1\), \[\{F^n_{r,i}, G^{n}_{r,j} \colon r\ge1, 1 \le i \le p(n,r), 1 \le j \le q(n,r)\}\] is a collection of nonempty disjoint elements of \(\mathcal{F}\), and \[\sum_{r=1}^\infty 2^{-2r}( p(n,r) + q(n,r)) \le 1.\] In particular, \(p(n,r) + q(n,r) \le 2^{2r}\) for all \(n,r\ge 1\). By a diagonal argument, passing to a subsequence and relabelling, we may assume that \[p(n,r) = p_r, q(n,r) = q_r \quad\text{for all n \ge r}.\] By compactness of \(\mathcal{F}\), we may also assume that \[\lim_{n \rightarrow \infty} F^n_{r,i} = F_{r,i}\quad (1 \le i \le p_r), \lim_{n \rightarrow \infty} G^n_{r,j} = G_{r,j} \quad (1 \le j \le q_r),\] where \[\{F_{r,i}, G_{r,j} \colon r\ge1, 1 \le i \le p_r, 1 \le j \le q_r\}\] is a collection of disjoint (possibly empty) elements of \(\mathcal{F}\) and \[\sum_{r=1}^\infty 2^{-2r}( p_r + q_r) \le 1.\] Set \(F_r = \cup_{i=1}^{p_r} F_{r,i}\) and \(G_r = \cup_{i=1}^{q_r} G_{r,i}\). It follows that \[k = \sum_{r=1}^\infty 2^{-r}(1_{F_r} - 1_{G_r}) \in K\] and \(k = \lim_{n \rightarrow \infty} k_n\) pointwise in \(\ell_\infty\). So \(K\) is compact (and metrizable) in the topology of pointwise convergence.

For \(x \in B(\mathcal{F})\), define \(\hat{x}\colon K \rightarrow \mathbb{R}\) by \(\hat{x}(k) = k(x).\) Suppose that \((k_n) \subset K\) and \(k_n \rightarrow k\) pointwise in \(\ell_\infty\). Clearly, \(\hat{x}(k_n) \rightarrow \hat{x}(k)\) when \(x\) has finite support. Since the finitely supported vectors are norm-dense in \(B(\mathcal{F})\), it follows that \(\hat{x}(k_n) \rightarrow \hat{x}(k)\) for all \(x \in B(\mathcal{F})\), i.e., that \(\hat{x}\) is continuous on \(K\). Since \(K\) is \(4\)-norming for \(B(\mathcal{F})\), the mapping \(x \mapsto \hat{x}\) defines a linear isomorphism from \(B(\mathcal{F})\) onto a closed subspace of \(C(K)\).

Suppose that \((x_n) \subset B(\mathcal{F})\) is bounded and coordinatewise null with respect to \((e_n)\). It follows from 24 that \[\lim_{n\rightarrow \infty} \hat{x}_n(k) = 0 \qquad (k \in K).\] Hence, by the Riesz representation and bounded convergence theorems, \(\hat{x}_n \rightarrow 0\) weakly in \(C(K)\). In particular, if \(x_n =\sum_{i = p_{n-1}+1}^{p_n}a_i e_i\), where \(p_{n-1} < p_n\), is a bounded block basis of \((e_n)\), then \((\hat{x}_n)\) is weakly null in \(C(K)\). Hence \((x_n)\) is weakly null in \(B(\mathcal{F})\), which implies that \((e_n)\) is a shrinking basis. On the other hand, since \((e_n)\) satisfies a lower \(2\)-estimate, it is boundedly complete. It follows from a theorem of James [14] that \(B(\mathcal{F})\) is reflexive.

Next suppose that \(\Gamma\) is uncountable. Let \(\Gamma_0\) be a countably infinite subset of \(\Gamma\). Then \[X_0 = \{ x \in B(\mathcal{F}) \colon \operatorname{supp} x \subseteq \Gamma_0\}\] is the Baernstein space on \(\Gamma_0\) corresponding to the family \(\mathcal{F}_0 = \{F \cap \Gamma_0 \colon F \in \mathcal{F} \}\). By the first part of the proof, \(X_0\) is reflexive. But every separable subspace of \(B(\mathcal{F})\) is contained in \(X_0\) for some \(\Gamma_0\). Hence every separable subspace of \(B(\mathcal{F})\) is reflexive, which implies that \(B(\mathcal{F})\) is also reflexive since reflexivity is separably determined. ◻

References↩︎

[1]
V.D. Milman, Geometric theory of Banach spaces. II: geometry of the unit sphere, Russian Math. Survey 26(1972), 79–163, (transl. from Russian).
[2]
R.C. James, Reflexivity and the sup of linear functionals, Israel J. Math. 13(1972), 289–300.
[3]
Petr Hájek and Michal Johanis, Characterization of reflexivity by equivalent renorming, J. Funct. Anal. 211(2004), 163–172.
[4]
E. Odell and Th. Schlumprecht, Asymptotic properties of Banach spaces under renormings, J. Amer. Math. Soc. 11(1998), 175–188.
[5]
Gilles Godefroy, Renormings of Banach spaces, Handbook of the geometry of Banach spaces, Vol. I, 781–835, NorthHolland Publishing Co., Amsterdam, 2001.
[6]
A. Baernstein, On reflexivity and summability, Studia Math. 42(1972), 91–94.
[7]
D.E. Alspach and S.A. Argyros, Complexity of weakly null sequences, Diss. Math. 321(1992), 1–44.
[8]
Spiros A. Argyros and Pavlos Motakis, \(\alpha\)-Large families and applications to Banach space theory, Topology and its Applications 172(2014),47–67.
[9]
D.N. Kutzarova and S.L.Troyanski, Reflexive Banach spaces without equivalent norms that are uniformly convex or uniformly differentiable in every direction, Studia Math. 72(1982), 91–95.
[10]
Y. Benyamini and T. Starbird, Embedding weakly compact sets into Hilbert space, Israel J. Math. 23(1970), 137–141.
[11]
S.J. Dilworth, Denka Kutzarova and Pavlos Motakis, Symmetric \(2R\) renormings(in preparation).
[12]
R. Deville, G. Godefroy, V. Zizler, Smoothness and renormings in Banach spaces, Monographs and Surveys in Pure and Applied Mathematics, Vol. 64, Pitman, London, 1993.
[13]
M. Day, Strict convexity and smoothness, Trans. Amer. Math. Soc. 78(1955), 516–528.
[14]
R.C. James, Bases and reflexivity of Banach spaces, Ann. Math. 52(1950), 518–527.

  1. The first author was supported by Simons Foundation Collaboration Grant No. 849142. The second author was supported by Simons Foundation Collaboration Grant No. 636954.↩︎