Skip to main content
\(\DeclareMathOperator{\dist}{dist} \DeclareMathOperator{\Lim}{Lim^1} \DeclareMathOperator{\diag}{diag} \DeclareMathOperator{\trace}{Tr} \DeclareMathOperator{\spans}{span} \DeclareMathOperator{\spec}{\sigma} \DeclareMathOperator{\essspec}{\sigma_{\mathrm{ess}}} \DeclareMathOperator{\ind}{ind} \newcommand{\Hil}{\mathcal{H}} \newcommand{\vmid}{\,\middle\vert\,} \renewcommand{\epsilon}{\varepsilon} \renewcommand{\emptyset}{\varnothing} \newcommand{\norm}[1]{\left\lVert #1 \right\rVert } \newcommand{\abs}[1]{\left\lvert #1 \right\rvert} \newcommand{\angles}[1]{\left\langle #1 \right\rangle} \newcommand{\snorm}[1]{\lVert #1 \rVert } \newcommand{\sabs}[1]{\lvert #1 \rvert} \newcommand{\sangles}[1]{\langle #1 \rangle} \newcommand{\lt}{<} \newcommand{\gt}{>} \newcommand{\amp}{&} \)

Section4Arveson's Theorem Revisited

In this section we apply the results concerning restricted diagonalization to prove a few key facts which will yield a reformulation and extension of Arveson's theorem (Theorem 4.3). Our first result in this direction is Theorem 4.2 which characterizes the condition \((d_n) \in \Lim\big(\spec(N)\big)\) in terms of restricted diagonalization. In order to prove Theorem 4.2, we use a straightforward geometric lemma which serves a similar purpose as Lemma 1 [2].

Relabel the \(\lambda_j\) if necessary so that \(k=1\text{.}\) By applying a rotation, translation and scaling (which preserve proportional distances), we may suppose that \(\lambda_1 = 1\) and \(L = -a + i\mathbb{R}\) for some \(a \ge 0\) so that the real part \(\Re(x) = 0\text{.}\) Note that \(-a \ge \max_{j \ge 2} \{ \Re(\lambda_j)\}\text{.}\) Since \(0 \in [-a,1]\) we may write

\begin{equation*} t (-a) + (1-t) 1 = 0, \quad\text{for}\quad t = \frac{1}{1 + a} \end{equation*}

Now

\begin{equation*} 0 = \Re(x) = \sum_{j=1}^m c_j \Re(\lambda_j) \le \Bigg( \sum_{j=2}^m c_j \Bigg) \max_{j \ge 2} \{\Re(\lambda_j)\} + c_1 \lambda_1 \le \Bigg( \sum_{j=2}^m c_j \Bigg) (-a) + c_1 1. \end{equation*}

Since we have two convex combinations of \(-a, 1\) and the latter is closer to \(1\) than the former, the convexity coefficients satisfy

\begin{equation*} \sum_{j = 2}^m c_j \le t = \frac{1}{1+a} = \frac{\dist(\Re(x),\lambda_1)}{\dist(\lambda_1,L)} \le \frac{\abs{x-\lambda_1}}{\dist(\lambda_1,L)}. \qedhere \end{equation*}

We first reduce to the case when \(\spec(N) = \essspec(N)\text{.}\) Since \(N\) is a normal operator with finite spectrum, by the spectral theorem there is a finite rank perturbation \(N'\) of \(N\) for which \(N'\) is normal and \(\spec(N') = \essspec(N') = \essspec(N)\text{.}\) In particular, if \(P_{\lambda}\) are the spectral projections of \(N\) onto \(\{\lambda\}\text{,}\) and \(\lambda' \in \essspec(N)\) is a distinguished element, then we can choose

\begin{equation*} N' := \lambda' P + \sum_{\lambda \in \essspec(N)} \lambda P_{\lambda}, \quad\text{where}\quad P = \sum_{\lambda \notin \essspec(N)} P_{\lambda}. \end{equation*}

Since \(N'-N\) is finite rank, the diagonals of \(N'\) and \(N\) differ by an absolutely summable sequence, so \((d_n) \in \Lim(X)\) if and only if the diagonal of \(N'\) is in \(\Lim(X)\text{.}\) Moreover, the spectral projections of \(N\) and \(N'\) differ from one another by finite projections. Therefore, the spectral projections of \(N\) each differ from a diagonal projection by a Hilbert–Schmidt operator if and only if the same holds true for \(N'\text{.}\) By Theorem 3.4, \(N\) is diagonalizable by a unitary which is a Hilbert–Schmidt perturbation of the identity if and only if \(N'\) is as well. Therefore, by the above reduction, it suffices to prove the theorem with the added assumption that \(\spec(N) = \essspec(N)\text{.}\)

(Proof of \(\Rightarrow\)) Enumerate the elements of \(\essspec(N) = \spec(N)\) as \(\lambda_1,\ldots,\lambda_m\text{.}\) Let \(P_j\) denote the spectral projection corresponding to the eigenvalue \(\lambda_j\text{,}\) so that \(N = \sum_{j=1}^m \lambda_j P_j\text{.}\) Let \(\{e_n\}_{n=1}^{\infty}\) denote the orthonormal basis corresponding to the diagonal \((d_n)\text{.}\) Suppose \((d_n) \in \Lim(X)\text{,}\) and so there exist \(x_n \in X\) for which \((d_n - x_n) \in \ell^1\text{.}\) Let \(\Lambda_k := \{ n \in \mathbb{N} \mid x_n = \lambda_k\}\) be the index set where the sequence \((x_n)\) takes the value \(\lambda_k \in X\text{.}\)

The projections \(P_j\) sum to the identity, so for each \(n \in \mathbb{N}\text{,}\) \(\sum_{j=1}^m \angles{P_j e_n,e_n} = 1\) and therefore

\begin{equation*} d_n = \sangles{Ne_n, e_n} = \sum_{j=1}^m \sangles{P_j e_n, e_n} \lambda_j \end{equation*}

is a convex combination of the spectrum.

For \(\lambda_k \in X\text{,}\) let \(L_k\) be a line separating \(\lambda_k\) from the remaining elements of \(\essspec(N)\text{.}\) Such a line \(L_k\) exists because \(\lambda_k\) is an extreme point of the convex hull of \(\essspec(N)\text{,}\) and this is a finite set. Since \((d_n) \in \Lim (X)\) we know that \((d_n - \lambda_k)_{n \in \Lambda_k}\) is absolutely summable for every \(k\text{.}\) Therefore, for all but finitely many indices \(n \in \Lambda_k\text{,}\) the diagonal entry \(d_n\) lies on a line parallel to \(L_k\) separating \(\lambda_k\) from \(L_k\) and hence also \(\essspec(N) \setminus \{ \lambda_k \}\text{.}\)

By Lemma 4.1, for these indices \(n \in \Lambda_k\text{,}\)

\begin{equation} \sum_{\substack{j=1 \\ j\not=k}}^m \sangles{P_j e_n, e_n} \le \frac{\abs{d_n - \lambda_k}}{\dist(\lambda_k,L_k)}.\label{eq-P_j-in-ell-1}\tag{4.1} \end{equation}

Since this inequality holds for all but finitely many \(n \in \Lambda_k\text{,}\) and \(\dist(\lambda_k,L_k)\) is independent of \(n \in \Lambda_k\text{,}\) and \((d_n - \lambda_k)_{n \in \Lambda_k}\) is absolutely summable, (4.1) proves \(\big(\langle P_j e_n, e_n \rangle\big)_{n \in \Lambda_k}\) lies in \(\Lim (\{0\}) = \ell^1\) when \(j \not= k\text{.}\) If \(\lambda_j \in \essspec(N) \setminus X\text{,}\) by letting \(\lambda_k\) run through \(X\text{,}\) we find \(\big(\langle P_j e_n, e_n \rangle\big)_{n \in \mathbb{N}}\) is absolutely summable since \(\bigcup_{\lambda_k \in X} \Lambda_k = \mathbb{N}\text{.}\) This implies \(P_j\) is trace-class and hence a finite projection, contradicting the fact that \(\lambda_j \in \essspec(N)\text{.}\) Therefore \(X = \essspec(N)\text{.}\)

Now consider \(\lambda_j \in X = \essspec(N)\text{.}\) In analogy with the previous paragraph, using the fact that \(\big(\langle P_j e_n, e_n \rangle\big)_{n \in \Lambda_k} \in \ell^1\) when \(j \not= k\) and letting \(\lambda_k\) run through \(X \setminus \lambda_j\text{,}\) we find \(\big(\langle P_j e_n, e_n \rangle\big)_{n \notin \Lambda_j} \in \ell^1\text{.}\) Finally, for \(n \in \Lambda_j\text{,}\)

\begin{equation*} 1 - \sangles{P_j e_n, e_n} = \sum_{\substack{k=1 \\ k\not=j}}^m \sangles{P_k e_n, e_n}, \end{equation*}

and hence \(\big(1- \sangles{P_j e_n, e_n}\big)_{n \in \Lambda_k}\) is a finite sum of absolutely summable sequences, and is therefore absolutely summable. Thus \(\big(\sangles{P_j e_n, e_n}\big)_{n \in \Lambda_k} \in \Lim(\{1\})\text{,}\) so \(\big(\sangles{P_j e_n,e_n}\big) \in \Lim(\{0,1\})\text{.}\) Therefore, by Corollary 3.1, \(P_j\) differs from a diagonal projection by a Hilbert–Schmidt operator. Since this is true of all the spectral projections of \(N\text{,}\) we may apply Theorem 3.4 to conclude that \(N\) is a diagonalizable by a Hilbert–Schmidt perturbation of the identity.

(Proof of \(\Leftarrow\)) This implication is a direct corollary of Theorem 3.8. To see this, suppose \(\essspec(N) = X\) and \(N\) is diagonalizable by a unitary \(U\) which is a Hilbert–Schmidt perturbation of the identity. Thus \(UNU^{*} = \diag(x_n)\) for some sequence \(x_n \in \essspec(N) = X\text{.}\) Then by Theorem 3.8, \(E(N-UNU^{*})\) is trace-class. That is, \(\trace \big(E(N-UNU^{*})\big) = \sum_{n=1}^{\infty} (d_n-x_n)\) is an absolutely summable series, so \((d_n) \in \Lim(X)\text{.}\)

We now establish our generalized operator-theoretic reformulation of Arveson's Theorem 1.5 by means of Theorem 3.8. After the proof we will explain how to derive Theorem 1.5 from Theorem 4.3.

Suppose \(N\) is normal operator with finite spectrum which is diagonalizable by a unitary \(U\) that is a Hilbert–Schmidt perturbation of the identity. Then by Theorem 3.8, \(E(UNU^{*}-N)\) is trace-class with trace zero. Moreover, \(\spec(UNU^{*}) = \spec(N)\text{,}\) thereby proving that an \(N'\) as in the statement exists.

Now, let \(N'\) be any diagonal operator with \(\spec(N') \subseteq \spec(N)\) for which \(E(N-N')\) is trace-class. Since \(N'\) and \(UNU^{*}\) are diagonal, we find

\begin{equation} UNU^{*}-N' = E(UNU^{*}-N') = E(UNU^{*}-N)+E(N-N')\label{eq-diagonal-split}\tag{4.3} \end{equation}

is trace-class, diagonal, and has finite spectrum contained in the set of differences \(\spec(N)-\spec(N)\text{.}\) Together, these conditions imply this operator is finite rank. Moreover, the (diagonal) spectral projections of \(UNU^{*}, N'\text{,}\) which we denote \(R_{\lambda},Q_{\lambda}\text{,}\) respectively for \(\lambda \in \spec(N)\text{,}\) each differ by a finite rank operator. Here we allow for the case \(Q_{\lambda} = 0\) when \(\lambda \in \spec(N) \setminus \spec(N')\text{.}\) This guarantees

\begin{equation*} [R_{\lambda}:Q_{\lambda}] = \trace(R_{\lambda}-Q_{\lambda}), \end{equation*}

using, for example, Proposition 2.4; however, this formula for essential codimension holds whenever the difference of the projections is trace-class and is widely known (see for instance [4] Theorem 4.1, [3] Theorem 3, or [9] Corollary 3.3).

Therefore,

\begin{equation} \trace(UNU^{*}-N') = \trace \left( \sum_{\lambda \in \spec(N)} (\lambda R_{\lambda} - \lambda Q_{\lambda}) \right) = \sum_{\lambda \in \spec(N)} [R_{\lambda} : Q_{\lambda}] \lambda.\label{eq-trace-ess-codim-formula}\tag{4.4} \end{equation}

Moreover, we can replace \(R_{\lambda}\) with \(P_{\lambda}\) in the right-most side of the above display. Indeed, since \(U\) conjugates \(P_{\lambda}, R_{\lambda}\text{,}\) \([P_{\lambda} : R_{\lambda}] = 0\) by Proposition 2.3, and furthermore \([P_{\lambda} : Q_{\lambda}] = [P_{\lambda} : R_{\lambda}] + [R_{\lambda} : Q_{\lambda}]\) by Proposition 2.2(iii).

Finally, since \(\trace\big(E(UNU^{*}-N)\big) = 0\text{,}\) using (4.3) and (4.4) we find that

\begin{equation*} \trace\big(E(N-N')\big) = \trace(UNU^{*}-N') = \sum_{\lambda \in \spec(N)} [P_{\lambda}:Q_{\lambda}] \lambda. \qedhere \end{equation*}

We now illustrate how our results may be used to provide a new proof of Arveson's theorem.

Let \(X = \{\lambda_1,\ldots,\lambda_m\}\) and \(d = (d_1,d_2,\ldots)\) be as in Theorem 1.5. That is, \(X\) is the set of vertices of a convex polygon in \(\mathbb{C}\text{,}\) and \(d\) satisfies

\begin{equation*} \sum_{n=1}^{\infty} \abs{f(d_n)} \lt \infty, \end{equation*}

where \(f(z) = (z-\lambda_1)(z-\lambda_2)\cdots(z-\lambda_m)\text{.}\) As we remarked after Theorem 1.5, this summability condition is equivalent to \(d \in \Lim (X)\) by Proposition 2 of [2]. Now suppose \(d\) is the diagonal of an operator \(N \in \mathcal{N}(X)\) (i.e., \(N\) is normal with \(\spec(N) = \essspec(N) = X\)). Then by Theorem 4.2, \(N\) is diagonalizable by a unitary \(U = I+K\) with \(K\) Hilbert–Schmidt. Therefore, we may apply Theorem 4.3 to conclude that \(\trace\big(E(N-N')\big) \in K_{\spec(N)} = K_X\) for some diagonal operator \(N'\) with \(\spec(N') \subseteq \spec(N)\) and \(E(N-N')\) is trace-class. Finally, equation (3.2) of Proposition 3.5 establishes

\begin{equation*} \sum_{n=1}^{\infty} (d_n - x_n) = \trace\big(E(N-N')\big) \in K_X \end{equation*}

where \((x_n)\) is the diagonal of \(N'\text{,}\) so \(x_n \in \spec(N') = \spec(N) = X\text{.}\) Hence \(s(d) = 0\text{.}\)

Remark4.4

In [6], Bownik and Jasper completely characterized the diagonals of selfadjoint operators with finite spectrum. A few of the results we have presented herein are generalizations of Theorem 4.1 of [6], which consists of some necessary conditions for a sequence to be the diagonal of a finite spectrum selfadjoint operator. In particular, the statement \((d_n) \in \Lim(X)\) implies \(X = \essspec(N)\) of our Theorem 4.2 is an extension to finite spectrum normal operators of their corresponding result [6] Theorem 4.1(ii) for selfadjoint operators. Similarly, our formula (4.2) of Theorem 4.3 generalizes [6] Theorem 4.1(iii).

We conclude with another perspective on the trace \(\trace\big(E(N-N')\big)\text{.}\) Our next corollary shows that when the \(\mathbb{Z}\)-module \(K_{\spec(N)}\) has full rank (i.e., \(\rank K_{\spec(N)}\) is one less than the number of elements in the spectrum), this trace is zero if and only if \(N'\) is a diagonalization of \(N\) by a unitary \(U = I + K\) with \(K\) Hilbert–Schmidt.

By Theorem 4.3, the differences \(P_k - Q_k\) are Hilbert–Schmidt and

\begin{equation*} 0 = \trace\big(E(N-N')\big) = \sum_{k=1}^m [P_k:Q_k]\lambda_k \end{equation*}

Since \(\sum_{k=1}^m [P_k:Q_k] = 0\text{,}\) we have \([P_1:Q_1] = - \sum_{k=2}^m [P_k:Q_k]\) and so we may rearrange the equality above to

\begin{equation*} 0 = \sum_{k=2}^m [P_k:Q_k](\lambda_1 - \lambda_k). \end{equation*}

Since \(\lambda_1-\lambda_2,\ldots,\lambda_1-\lambda_m\) are linearly independent in \(K_{\spec(N)}\text{,}\) we conclude that the coefficients \([P_k:Q_k] = 0\) for \(2 \le k \le m\text{.}\) In turn, this implies \([P_1:Q_1] = 0\text{.}\) Therefore, by Lemma 3.2, there is a unitary \(U = I+K\) with \(K\) Hilbert–Schmidt conjugating each \(P_k\) to \(Q_k\text{.}\) Thus \(UNU^{*} = N'\text{.}\)