Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-16T13:54:44.986Z Has data issue: false hasContentIssue false

$\boldsymbol {C}^{*}$-ALGEBRAS FROM $\boldsymbol {K}$ GROUP REPRESENTATIONS

Published online by Cambridge University Press:  08 March 2022

VALENTIN DEACONU*
Affiliation:
Department of Mathematics and Statistics, University of Nevada, Reno, NV89557-0084, USA
*
Rights & Permissions [Opens in a new window]

Abstract

We introduce certain $C^*$ -algebras and k-graphs associated to k finite-dimensional unitary representations $\rho _1,\ldots ,\rho _k$ of a compact group G. We define a higher rank Doplicher-Roberts algebra $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ , constructed from intertwiners of tensor powers of these representations. Under certain conditions, we show that this $C^*$ -algebra is isomorphic to a corner in the $C^*$ -algebra of a row-finite rank k graph $\Lambda $ with no sources. For G finite and $\rho _i$ faithful of dimension at least two, this graph is irreducible, it has vertices $\hat {G}$ and the edges are determined by k commuting matrices obtained from the character table of the group. We illustrate this with some examples when $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ is simple and purely infinite, and with some K-theory computations.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

The study of graph $C^*$ -algebras was motivated, among other reasons, by the Doplicher–Roberts algebra $\mathcal {O}_\rho $ associated to a group representation $\rho $ (see [Reference Kajiwara, Pinzari and Watatani19, Reference Mann, Raeburn and Sutherland22]). It is natural to imagine that a rank k graph is related to a fixed set of k representations $\rho _1,\ldots ,\rho _k$ satisfying certain properties.

Given a compact group G and k finite-dimensional unitary representations $\rho _i$ on Hilbert spaces $\mathcal H_i$ of dimensions $d_i$ for $i=1,\ldots ,k$ , we first construct a product system $\mathcal E$ indexed by the semigroup $(\mathbb {N}^k,+)$ with fibers $\mathcal E_{n}=\mathcal H_1^{\otimes n_1}\otimes \cdots \otimes \mathcal H_k^{\otimes n_k}$ for $n=(n_1,\ldots ,n_k)\in \mathbb {N}^k$ . Using the representations $\rho _i$ , the group G acts on each fiber of $\mathcal {E}$ in a compatible way, so we obtain an action of G on the Cuntz–Pimsner algebra $\mathcal {O}(\mathcal {E})$ . This action determines the crossed product $\mathcal {O}(\mathcal {E})\rtimes G$ and the fixed point algebra $\mathcal {O}(\mathcal {E})^G$ .

Inspired by Section 7 of [Reference Kajiwara, Pinzari and Watatani19] and Section 3.3 of [Reference Albandik and Meyer1], we define a higher rank Doplicher–Roberts algebra $\mathcal O_{\rho _1,\ldots ,\rho _k}$ associated to the representations $\rho _1,\ldots ,\rho _k$ . This algebra is constructed from intertwiners $Hom (\rho ^n, \rho ^m)$ , where $\rho ^n=\rho _1^{\otimes n_1}\otimes \cdots \otimes \rho _k^{\otimes n_k}$ is acting on $\mathcal {H}^n=\mathcal H_1^{\otimes n_1}\otimes \cdots \otimes \mathcal H_k^{\otimes n_k}$ for $n=(n_1,\ldots ,n_k)\in \mathbb N^k$ . We show that $\mathcal O_{\rho _1,\ldots ,\rho _k}$ is isomorphic to $\mathcal {O}(\mathcal {E})^G$ .

If the representations $\rho _1,\ldots ,\rho _k$ satisfy some mild conditions, we construct a k-colored graph $\Lambda $ with vertex space $\Lambda ^0=\hat {G}$ , and with edges $\Lambda ^{\varepsilon _i}$ given by some matrices $M_i$ indexed by $\hat {G}$ . Here $\varepsilon _i=(0,\ldots ,1,\ldots ,0)\in \mathbb {N}^k$ with $1$ in position i are the canonical generators. For $v,w\in \hat {G}$ , the matrices $M_i$ have entries

$$ \begin{align*}M_i(w,v)=|\{e\in \Lambda^{\varepsilon_i}: s(e)=v, r(e)=w\}|=\dim Hom(v,w\otimes \rho_i),\end{align*} $$

which is the multiplicity of v in $w\otimes \rho _i$ for $i=1,\ldots ,k$ . Note that the matrices $M_i$ commute because $\rho _i\otimes \rho _j\cong \rho _j\otimes \rho _i$ for all $i,j=1,\ldots ,k$ and therefore

$$ \begin{align*}\dim Hom(v,w\otimes \rho_i\otimes\rho_j)=\dim Hom(v,w\otimes \rho_j\otimes\rho_i).\end{align*} $$

By a particular choice of isometric intertwiners in $Hom(v,w\otimes \rho _i)$ for each $v,w\in \hat {G}$ and for each i, we can choose bijections

$$ \begin{align*}\lambda_{ij}:\Lambda^{\varepsilon_i}\times_{\Lambda^0}\Lambda^{\varepsilon_j}\to \Lambda^{\varepsilon_j}\times_{\Lambda^0}\Lambda^{\varepsilon_i},\end{align*} $$

obtaining a set of commuting squares for $\Lambda $ . For $k\ge 3$ , we need to check the associativity of the commuting squares, that is,

$$ \begin{align*}(id_\ell\times \lambda_{ij})(\lambda_{i\ell}\times id_j)(id_i\times \lambda_{j\ell})=(\lambda_{j\ell}\times id_i)(id_j\times \lambda_{i\ell})(\lambda_{ij}\times id_\ell)\end{align*} $$

as bijections from $\Lambda ^{\varepsilon _i}\times _{\Lambda ^0}\Lambda ^{\varepsilon _j}\times _{\Lambda ^0}\Lambda ^{\varepsilon _\ell }$ to $\Lambda ^{\varepsilon _\ell }\times _{\Lambda ^0}\Lambda ^{\varepsilon _j}\times _{\Lambda ^0}\Lambda ^{\varepsilon _i}$ for all $i<j<\ell $ (see [Reference Fowler and Sims14]). If these conditions are satisfied, we obtain a rank k graph $\Lambda $ , which is row-finite with no sources but, in general, is not unique.

In many situations, $\Lambda $ is cofinal and it satisfies the aperiodicity condition, so $C^*(\Lambda )$ is simple. For $k=2$ , the $C^*$ -algebra $C^*(\Lambda )$ is unique when it is simple and purely infinite, because its K-theory depends only on the matrices $M_1, M_2$ . It is an open question what happens for $k\ge 3$ .

Assuming that the representations $\rho _1,\ldots ,\rho _k$ determine a rank k graph $\Lambda $ , we prove that the Doplicher–Roberts algebra $\mathcal O_{\rho _1,\ldots ,\rho _k}$ is isomorphic to a corner of $C^*(\Lambda )$ , so if $C^*(\Lambda )$ is simple, then $\mathcal O_{\rho _1,\ldots ,\rho _k}$ is Morita equivalent to $C^*(\Lambda )$ . In particular cases, we can compute its K-theory using results from [Reference Evans11].

2 The product system

Product systems over arbitrary semigroups were introduced by Fowler [Reference Fowler13], inspired by work of Arveson, and studied by several authors (see [Reference Albandik and Meyer1, Reference Carlsen, Larsen, Sims and Vittadello4, Reference Sims and Yeend26]). In this paper, we are mostly interested in product systems $\mathcal {E}$ indexed by $( \mathbb {N}^k , +)$ , associated to some representations $\rho _1,\ldots ,\rho _k$ of a compact group G. We remind the reader of some general definitions and constructions with product systems, but we restrict our attention to the Cuntz–Pimsner algebra $\mathcal {O}(\mathcal {E})$ and we mention some properties in particular cases only (see Example 2.3 for $P=\mathbb {N}^k$ ).

Definition 2.1. Let $(P, \cdot )$ be a discrete semigroup with identity e and let A be a $C^*$ -algebra. A product system of $C^*$ -correspondences over A indexed by P is a semigroup $\mathcal {E}=\bigsqcup _{p\in P}\mathcal {E}_p$ and a map $\mathcal {E}\to P$ such that:

  • for each $p\in P$ , the fiber $\mathcal {E}_p\subset \mathcal {E}$ is a $C^*$ -correspondence over A with inner product $\langle \cdot ,\cdot \rangle _p$ ;

  • the identity fiber $\mathcal {E}_e$ is A viewed as a $C^*$ -correspondence over itself;

  • for $p,q\in P\setminus \{e\}$ , the multiplication map

    $$ \begin{align*}\mathcal{M}_{p,q}:\mathcal{E}_p\times \mathcal{E}_q\to \mathcal{E}_{pq},\;\; \mathcal{M}_{p,q}(x,y)= xy\end{align*} $$
    induces an isomorphism $\mathcal {M}_{p,q}:\mathcal {E}_p\otimes _A \mathcal {E}_q\to \mathcal {E}_{pq}$ ; and
  • multiplication in $\mathcal {E}$ by elements of $\mathcal {E}_e=A$ implements the right and left actions of A on each $\mathcal {E}_p$ . In particular, $\mathcal {M}_{p,e}$ is an isomorphism.

Let $\phi _p:A\to \mathcal {L}(\mathcal {E}_p)$ be the homomorphism implementing the left action. The product system $\mathcal {E}$ is said to be essential if each $\mathcal {E}_p$ is an essential correspondence, that is, if the span of $\phi _p(A)\mathcal {E}_p$ is dense in $\mathcal {E}_p$ for all $p\in P$ . In this case, the map $\mathcal {M}_{e,p}$ is also an isomorphism.

If the maps $\phi _p$ take values in $\mathcal {K}(\mathcal {E}_p)$ , then the product system is called row-finite or proper. If all maps $\phi _p$ are injective, then $\mathcal {E}$ is called faithful.

Definition 2.2. Given a product system $\mathcal {E}\to P$ over A and a $C^*$ -algebra B, a map $\psi :\mathcal {E}\to B$ is called a Toeplitz representation of $\mathcal {E}$ if:

  • denoting $\psi _p:=\psi |_{\mathcal {E}_p}$ , each $\psi _p:\mathcal {E}_p\to B$ is linear, $\psi _e:A\to B$ is a $*$ -homomorphism, and

    $$ \begin{align*}\psi_e(\langle x,y\rangle_p)=\psi_p(x)^*\psi_p(y)\end{align*} $$
    for all $x,y\in \mathcal {E}_p$ ; and
  • $\psi _p(x)\psi _q(y)=\psi _{pq}(xy)$ for all $p,q\in P, x\in \mathcal {E}_p, y\in \mathcal {E}_q$ .

For each $p\in P$ , we write $\psi ^{(p)}$ for the homomorphism $\mathcal {K}(\mathcal {E}_p)\to B$ obtained by extending the map $\theta _{\xi , \eta }\mapsto \psi _p(\xi )\psi _p(\eta )^*$ , where

$$ \begin{align*}\theta_{\xi, \eta}(\zeta)=\xi\langle \eta, \zeta\rangle.\end{align*} $$

The Toeplitz representation $\psi :\mathcal {E}\to B$ is Cuntz–Pimsner covariant if $\psi ^{(p)}(\phi _p(a))=\psi _e(a)$ for all $p\in P$ and all $a\in A$ such that $\phi _p(a)\in \mathcal {K}(\mathcal {E}_p)$ .

There is a $C^*$ -algebra $\mathcal {T}_A(\mathcal {E})$ called the Toeplitz algebra of $\mathcal {E}$ and a representation $i_{\mathcal {E}}:\mathcal {E}\to \mathcal {T}_A(\mathcal {E})$ which is universal in the following sense: $\mathcal {T}_A(\mathcal {E})$ is generated by $i_{\mathcal {E}}(\mathcal {E})$ and, for any representation $\psi :\mathcal {E}\to B$ , there is a homomorphism $\psi _*:\mathcal {T}_A(\mathcal {E})\to B$ such that $\psi _*\circ i_{\mathcal {E}}=\psi $ .

The Cuntz–Pimsner algebra $\mathcal {O}_A(\mathcal {E})$ of a product system $\mathcal {E}\to P$ is universal for Cuntz–Pimsner covariant representations.

There are various extra conditions on a product system $\mathcal {E}\to P$ and several other notions of covariance besides the Cuntz–Pimsner covariance from Definition 2.2, which allow one to define the Cuntz–Pimsner algebra $\mathcal {O}_A(\mathcal {E})$ or the Cuntz–Nica–Pimsner algebra $\mathcal {N}\mathcal {O}_A(\mathcal {E})$ satisfying certain properties (see [Reference Albandik and Meyer1, Reference Carlsen, Larsen, Sims and Vittadello4, Reference Dor-On and Kakariadis10, Reference Fowler13, Reference Sims and Yeend26], among others). We mention that $\mathcal {O}_A(\mathcal {E})$ (or $\mathcal {N}\mathcal {O}_A(\mathcal {E})$ ) comes with a covariant representation $j_{\mathcal {E}}:\mathcal {E}\to \mathcal {O}_A(\mathcal {E})$ and is universal in the following sense: $\mathcal {O}_A(\mathcal {E})$ is generated by $j_{\mathcal {E}}(\mathcal {E})$ and, for any covariant representation $\psi :\mathcal {E}\to B$ , there is a homomorphism $\psi _*:\mathcal {O}_A(\mathcal {E})\to B$ such that $\psi _*\circ j_{\mathcal {E}}=\psi $ . Under certain conditions, $\mathcal {O}_A(\mathcal {E})$ satisfies a gauge invariant uniqueness theorem.

Example 2.3. For a product system $\mathcal {E}\to P$ with fibers $\mathcal {E}_p$ that are nonzero finite-dimensional Hilbert spaces, and, in particular, $A=\mathcal {E}_e=\mathbb {C}$ , let us fix an orthonormal basis $\mathcal {B}_p$ in $\mathcal {E}_p$ . Then a Toeplitz representation $\psi :\mathcal {E}\to B$ gives rise to a family of isometries $\{\psi (\xi ): \xi \in \mathcal {B}_p\}_{p\in P}$ with mutually orthogonal range projections. In this case, $\mathcal {T}(\mathcal {E})=\mathcal {T}_{\mathbb {C}}(\mathcal {E})$ is generated by a collection of Cuntz–Toeplitz algebras which interact according to the multiplication maps $\mathcal {M}_{p,q}$ in $\mathcal {E}$ .

A representation $\psi :\mathcal {E}\to B$ is Cuntz–Pimsner covariant if

$$ \begin{align*} \sum_{\xi\in \mathcal{B}_p}\psi(\xi)\psi(\xi)^*=\psi(1)\end{align*} $$

for all $p\in P$ . The Cuntz–Pimsner algebra $\mathcal {O}(\mathcal {E})=\mathcal {O}_{\mathbb {C}}(\mathcal {E})$ is generated by a collection of Cuntz algebras, so it could be thought of as a multidimensional Cuntz algebra. Fowler proved in [Reference Fowler12] that if the function $p\mapsto \dim \mathcal {E}_p$ is injective, then the algebra ${\mathcal O}(\mathcal {E})$ is simple and purely infinite. For other examples of multidimensional Cuntz algebras, see [Reference Burgstaller3].

Example 2.4. A row-finite k-graph with no sources $\Lambda $ (see [Reference Kumjian, Pask, Raeburn and Renault18]) determines a product system $\mathcal {E}\to \mathbb {N}^k$ with $\mathcal {E}_0=A=C_0(\Lambda ^0)$ and $\mathcal {E}_n=\overline {C_c(\Lambda ^n)}$ for $n\neq 0$ such that we have a $\mathbb {T}^k$ -equivariant isomorphism $\mathcal {O}_A(\mathcal {E})\cong C^*(\Lambda )$ . Recall that, for product systems indexed by $\mathbb {N}^k$ , the universal property induces a gauge action on $\mathcal {O}_A(\mathcal {E})$ defined by $\gamma _z(j_{\mathcal {E}}(\xi ))=z^nj_{\mathcal {E}}(\xi )$ for $z\in \mathbb {T}^k$ and $\xi \in \mathcal {E}_n$ .

The following two definitions and two results are taken from [Reference Deaconu, Huang and Sims7]; see also [Reference Hao and Ng15, Reference Kumjian and Pask17].

Definition 2.5. An action $ \beta $ of a locally compact group $ G $ on a product system $ \mathcal {E} \to P $ over A is a family $ (\beta ^{p})_{p \in P} $ such that $ \beta ^{p} $ is an action of $ G $ on each fiber $\mathcal {E}_{p} $ compatible with the action $\alpha =\beta ^e$ on A, and, furthermore, the actions $(\beta ^p)_{p\in P}$ are compatible with the multiplication maps $\mathcal {M}_{p,q}$ in the sense that

$$ \begin{align*}\beta^{p q}_g(\mathcal{M}_{p,q}(x \otimes y)) = \mathcal{M}_{p,q}(\beta^{p}_g(x) \otimes \beta^{q}_g(y)) \end{align*} $$

for all $ g \in G $ , $ x \in \mathcal {E}_{p} $ and $ y \in \mathcal {E}_{q} $ .

Definition 2.6. If $ \beta $ is an action of $ G $ on the product system $\mathcal {E} \to P $ , we define the crossed product $\mathcal {E} \rtimes _{\beta } G $ as the product system indexed by $ P $ with fibers $ \mathcal {E}_{p} \rtimes _{\beta ^{p}} G $ , which are $ C^{\ast } $ -correspondences over $ A \rtimes _{\alpha } G $ . For $ \zeta \in C_c(G,\mathcal {E}_{p}) $ and $ \eta \in C_c(G,\mathcal {E}_{q}) $ , the product $ \zeta \eta \in C_c(G,\mathcal {E}_{p q}) $ is defined by

$$ \begin{align*}(\zeta \eta)(s) = \int_G\mathcal{M}_{p,q}(\zeta(t) \otimes \beta^{q}_t(\eta(t^{- 1} s)))\,dt. \end{align*} $$

Proposition 2.7. The set $ \mathcal {E} \rtimes _{\beta } G = \bigsqcup _{p \in P} \mathcal {E}_{p} \rtimes _{\beta ^{p}} G $ with the above multiplication satisfies all the properties of a product system of $ C^{\ast } $ -correspondences over $ A \rtimes _{\alpha } G $ .

Proposition 2.8. Suppose that a locally compact group $ G $ acts on a row-finite and faithful product system $ \mathcal {E} $ indexed by $ P = (\mathbb {N}^{k},+) $ via automorphisms $ \beta ^{p}_{g} $ . Then $ G $ acts on the Cuntz–Pimsner algebra $\mathcal {O}_{A}(\mathcal {E}) $ via automorphisms denoted by $ \gamma _{g} $ . Moreover, if $ G $ is amenable, then $ \mathcal {E} \rtimes _{\beta } G $ is row-finite and faithful, and

$$ \begin{align*}\mathcal{O}_{A}(\mathcal{E}) \rtimes_{\gamma} G \cong \mathcal{O}_{A \rtimes_{\alpha} G}(\mathcal{E} \rtimes_{\beta} G). \end{align*} $$

Now we define the product system associated to k representations of a compact group G. We limit ourselves to finite-dimensional unitary representations, even though the definition makes sense in greater generality.

Definition 2.9. Given a compact group G and k finite-dimensional unitary representations $\rho _i$ of G on Hilbert spaces $\mathcal H_i$ for $i=1,\ldots ,k$ , we construct the product system $\mathcal {E}=\mathcal {E}(\rho _1,\ldots ,\rho _k)$ indexed by the commutative monoid $(\mathbb N^k,+)$ , with fibers

$$ \begin{align*}\mathcal E_n=\mathcal{H}^n=\mathcal H_1^{\otimes n_1}\otimes\cdots\otimes \mathcal H_k^{\otimes n_k}\end{align*} $$

for $n=(n_1,\ldots ,n_k)\in \mathbb {N}^k$ ; in particular, $A=\mathcal E_0=\mathbb C$ . The multiplication maps

$$ \begin{align*}\mathcal {M}_{n,m}:\mathcal {E}_n\times \mathcal {E}_m\to \mathcal {E}_{n+m} \end{align*}$$

in $\mathcal {E}$ are defined by using the standard isomorphisms $\rho _i\otimes \rho _j\cong \rho _j\otimes \rho _i$ for all $i<j$ . The associativity in $\mathcal {E}$ follows from the fact that

$$ \begin{align*}\mathcal{M}_{n+m,p}\circ (\mathcal{M}_{n,m}\times id)=\mathcal{M}_{n,m+p}\circ ( id\times \mathcal{M}_{m,p})\end{align*} $$

as maps from $\mathcal {E}_n\times \mathcal {E}_m\times \mathcal {E}_p$ to $\mathcal {E}_{n+m+p}.$ Then $\mathcal {E}=\mathcal {E}(\rho _1,\ldots ,\rho _k)$ is called the product system of the representations $\rho _1,\ldots ,\rho _k$ .

Remark 2.10. Similarly, a semigroup P of unitary representations of a group G determines a product system $\mathcal {E}\to P$ .

Proposition 2.11. With notation as in Definition 2.9, assume that $d_i=\dim \mathcal {H}_i\ge 2$ . Then the Cuntz–Pimsner algebra $\mathcal {O}(\mathcal {E})$ associated to the product system $\mathcal {E}\to \mathbb {N}^k$ described above is isomorphic with the $C^*$ -algebra of a rank k graph $\Gamma $ with a single vertex and with $|\Gamma ^{\varepsilon _i}|=d_i$ . This isomorphism is equivariant for the gauge action. Moreover,

$$ \begin{align*}\mathcal{O}(\mathcal{E})\cong \mathcal O_{d_1}\otimes\cdots\otimes \mathcal O_{d_k},\end{align*} $$

where $\mathcal O_n$ is the Cuntz algebra.

Proof. Indeed, by choosing a basis in each $\mathcal {H}_i$ , we get the edges $\Gamma ^{\varepsilon _i}$ in a k-colored graph $\Gamma $ with a single vertex. The isomorphisms $\rho _i\otimes \rho _j\cong \rho _j\otimes \rho _i$ determine the factorization rules of the form $ef=fe$ for $e\in \Gamma ^{\varepsilon _i}$ and $f\in \Gamma ^{\varepsilon _j}$ , which obviously satisfy the associativity condition. In particular, the corresponding isometries in $C^*(\Gamma )$ commute and determine, by the universal property, a surjective homomorphism $\varphi $ onto $\mathcal {O}(\mathcal {E})$ , preserving the gauge action. Using the gauge invariant uniqueness theorem for k-graph algebras, the map $\varphi $ is an isomorphism. In particular, $\mathcal {O}(\mathcal {E})\cong \mathcal O_{d_1}\otimes \cdots \otimes \mathcal O_{d_k}$ .

Remark 2.12. For $d_i\ge 2$ , the $C^*$ -algebra $\mathcal {O}(\mathcal {E})\cong C^*(\Gamma )$ is always simple and purely infinite since it is a tensor product of simple and purely infinite $C^*$ -algebras. If $d_i=1$ for some i, then the isomorphism in Proposition 2.11 still holds, but $C^*(\Gamma )\cong \mathcal {O}(\mathcal {E})$ contains a copy of $C(\mathbb {T})$ , so it is not simple. Of course, if $d_i=1$ for all i, then $\mathcal {O}(\mathcal {E})\cong C(\mathbb {T}^k)$ . For more on single vertex rank k graphs, see [Reference Davidson and Yang5, Reference Davidson and Yang6].

Proposition 2.13. The compact group G acts on each fiber $\mathcal {E}_n$ of the product system $\mathcal E$ via the representation $\rho ^n=\rho _1^{\otimes n_1}\otimes \cdots \otimes \rho _k^{\otimes n_k}$ . This action is compatible with the multiplication maps and commutes with the gauge action of $\mathbb {T}^k$ . The crossed product $\mathcal E\rtimes G$ becomes a row-finite and faithful product system indexed by $\mathbb N^k$ over the group $C^*$ -algebra $C^*(G)$ . Moreover,

$$ \begin{align*}\mathcal{O}(\mathcal{E}) \rtimes G \cong \mathcal{O}_{C^*(G)}(\mathcal{E} \rtimes G).\end{align*} $$

Proof. Indeed, for $g\in G$ and $\xi \in \mathcal {E}_n=\mathcal {H}^n$ , we define $g\cdot \xi =\rho ^n(g)(\xi )$ , and since $\rho _i\otimes \rho _j\cong \rho _j\otimes \rho _i$ , we have $g\cdot (\xi \otimes \eta )=g\cdot \xi \otimes g\cdot \eta $ for $\xi \in \mathcal {E}_n, \eta \in \mathcal {E}_m$ . Clearly,

$$ \begin{align*}g\cdot\gamma_z(\xi)=g\cdot(z^n\xi)=z^n(g\cdot\xi)=\gamma_z(g\cdot\xi),\end{align*} $$

so the action of G commutes with the gauge action. Using Proposition 2.7, $\mathcal E\rtimes G$ becomes a product system indexed by $\mathbb N^k$ over $C^*(G)\cong \mathbb {C}\rtimes G$ with fibers $\mathcal {E}_n\rtimes G$ . The isomorphism $\mathcal {O}(\mathcal {E}) \rtimes G \cong \mathcal {O}_{C^*(G)}(\mathcal {E} \rtimes G)$ follows from Proposition 2.8.

Corollary 2.14. Since the action of G commutes with the gauge action, the group G acts on the core algebra $\mathcal {F}=\mathcal {O}(\mathcal {E})^{\mathbb {T}^k}$ .

Remark 2.15. In some cases, $\mathcal {O}(\mathcal {E})\rtimes G$ is isomorphic to the self-similar k-graph $C^*$ -algebras $\mathcal {O}_{G,\Lambda }$ introduced in [Reference Li and Yang21]. Moreover, for a self-similar k-graph $(G,\Lambda )$ with $|\Lambda ^0|=1$ , we have $\mathcal {O}_{G,\Lambda }\cong \mathcal {Q}(\Lambda \bowtie G)$ , where $\Lambda \bowtie G$ is a Zappa–Szép product and $\mathcal {Q}(\Lambda \bowtie G)$ is its boundary quotient $C^*$ -algebra (see Example 3.10(4) in [Reference Li and Yang21] and Theorem 3.3 in [Reference Li and Yang20]). I thank the referee for bringing this relationship to my attention.

3 The Doplicher–Roberts algebra

The Doplicher–Roberts algebras $\mathcal {O}_\rho $ , denoted by ${\mathcal O}_G$ in [Reference Doplicher and Roberts8], were introduced to construct a new duality theory for compact Lie groups G that strengthens the Tannaka–Krein duality. Here $\rho $ is the n-dimensional representation of G defined by the inclusion $G\subseteq U(n)$ in some unitary group $U(n)$ . Let ${\mathcal T}_G$ denote the representation category whose objects are tensor powers $\rho ^p=\rho ^{\otimes p}$ for $p\ge 0$ , and whose arrows are the intertwiners $Hom(\rho ^p, \rho ^q)$ . The group G acts via $\rho $ on the Cuntz algebra ${\mathcal O}_n$ and ${\mathcal O}_G={\mathcal O}_\rho $ is identified in [Reference Doplicher and Roberts8] with the fixed point algebra ${\mathcal O}_n^G$ . If $\sigma $ denotes the restriction to ${\mathcal O}_\rho $ of the canonical endomorphism of $\mathcal {O}_n$ , then ${\mathcal T}_G$ can be reconstructed from the pair $({\mathcal O}_\rho ,\sigma )$ . Subsequently, Doplicher–Roberts algebras were associated to any object $\rho $ in a strict tensor $C^*$ -category (see [Reference Doplicher and Roberts9]).

Given finite-dimensional unitary representations $\rho _1,\ldots ,\rho _k$ of a compact group G on Hilbert spaces $\mathcal H_1,\ldots , \mathcal H_k$ , we construct a Doplicher–Roberts algebra $\mathcal O_{\rho _1,\ldots ,\rho _k}$ from intertwiners

$$ \begin{align*}Hom (\rho^n, \rho^m)=\{T\in\mathcal{L}({\mathcal H}^n, {\mathcal H}^m)\;\mid \; T\rho^n(g)=\rho^m(g)T \text{ for all } g\in G\},\end{align*} $$

where, for $n=(n_1,\ldots ,n_k)\in \mathbb N^k$ , the representation $\rho ^n=\rho _1^{\otimes n_1}\otimes \cdots \otimes \rho _k^{\otimes n_k}$ acts on $\mathcal {H}^n=\mathcal H_1^{\otimes n_1}\otimes \cdots \otimes \mathcal H_k^{\otimes n_k}$ . Note that $\rho ^0=\iota $ is the trivial representation of G, acting on $\mathcal {H}^0=\mathbb {C}$ . This Doplicher–Roberts algebra is a subalgebra of $\mathcal {O}(\mathcal {E})$ for the product system $\mathcal {E}$ , as in Definition 2.9.

Lemma 3.1. Consider

$$ \begin{align*} \mathcal{A}_0=\bigcup_{m,n\in \mathbb{N}^k}\mathcal{L}(\mathcal H^n,\mathcal H^m).\end{align*} $$

Then the linear span of $\mathcal {A}_0$ becomes a $*$ -algebra $\mathcal {A}$ with appropriate multiplication and involution. This algebra has a natural $\mathbb {Z}^k$ -grading coming from a gauge action of $\mathbb {T}^k$ . Moreover, the Cuntz–Pimsner algebra $\mathcal {O}(\mathcal {E})$ of the product system $\mathcal {E}=\mathcal {E}(\rho _1,\ldots ,\rho _k)$ is equivariantly isomorphic to the $C^*$ -closure of $\mathcal {A}$ in the unique $C^*$ -norm for which the gauge action is isometric.

Proof. Recall that the Cuntz algebra $\mathcal {O}_n$ contains a canonical Hilbert space $\mathcal {H}$ of dimension n and it can be constructed as the closure of the linear span of $ \bigcup _{p,q\in \mathbb {N}}\mathcal {L}(\mathcal H^p,\mathcal H^q)$ using embeddings

$$ \begin{align*}\mathcal{L}(\mathcal{H}^p,\mathcal{H}^q)\subseteq \mathcal{L}(\mathcal{H}^{ p+1},\mathcal{H}^{ q+1}),\quad T\mapsto T\otimes I,\end{align*} $$

where $\mathcal {H}^p=\mathcal H^{\otimes p}$ and $I:\mathcal {H}\to \mathcal {H}$ is the identity map. This linear span becomes a $*$ -algebra with a multiplication given by composition and an involution (see [Reference Doplicher and Roberts8] and Proposition 2.5 in [Reference Katsoulis16]).

Similarly, for all $r\in \mathbb {N}^k$ , we consider embeddings $\mathcal {L}(\mathcal {H}^n,\mathcal {H}^m)\subseteq \mathcal {L}(\mathcal {H}^{n+r},\mathcal {H}^{m+r})$ given by $T\mapsto T\otimes I_r$ , where $I_r:{\mathcal H}^r\to {\mathcal H}^r$ is the identity map, and we endow $\mathcal {A}$ with a multiplication given by composition and an involution. More precisely, if $S\in \mathcal {L}(\mathcal {H}^n,\mathcal {H}^m)$ and $T\in \mathcal {L}(\mathcal {H}^q,\mathcal {H}^p)$ , then the product $ST$ is

$$ \begin{align*}(S\otimes I_{p\vee n-n})\circ (T\otimes I_{p\vee n-p})\in \mathcal{L}(\mathcal{H}^{q+p\vee n-p},\mathcal{H}^{m+p\vee n-n}),\end{align*} $$

where we write $p\vee n$ for the coordinatewise maximum. This multiplication is well defined in $\mathcal {A}$ and is associative. The adjoint of $T\in \mathcal {L}(\mathcal {H}^n,\mathcal {H}^m)$ is $T^*\in \mathcal {L}(\mathcal {H}^m,\mathcal {H}^n)$ .

There is a natural $\mathbb Z^k$ -grading on $\mathcal {A}$ given by the gauge action $\gamma $ of $\mathbb {T}^k$ , where, for $z=(z_1,\ldots ,z_k)\in \mathbb {T}^k$ and $T\in \mathcal {L}(\mathcal {H}^n,\mathcal {H}^m)$ , we define

$$ \begin{align*}\gamma_z(T)(\xi)=z_1^{m_1-n_1}\cdots z_k^{m_k-n_k}T(\xi).\end{align*} $$

Adapting the argument in Theorem 4.2 in [Reference Doplicher and Roberts9] for $\mathbb {Z}^k$ -graded $C^*$ -algebras, the $C^*$ -closure of $\mathcal {A}$ in the unique $C^*$ -norm for which $\gamma _z$ is isometric is well defined. The map

$$ \begin{align*}(T_1,\ldots,T_k)\mapsto T_1\otimes\cdots\otimes T_k, \end{align*} $$

where

$$ \begin{align*}T_1\otimes\cdots\otimes T_k: \mathcal H^n\to \mathcal H^m,\; (T_1\otimes\cdots\otimes T_k)(\xi_1\otimes\cdots\otimes \xi_k)=T_1(\xi_1)\otimes\cdots\otimes T_k(\xi_k)\end{align*} $$

for $T_i\in \mathcal {L}(\mathcal H_i^{n_i},\mathcal H_i^{m_i})$ for $i=1,\ldots ,k$ preserves the gauge action and it can be extended to an equivariant isomorphism from $\mathcal {O}(\mathcal {E})\cong \mathcal {O}_{d_1}\otimes \cdots \otimes \mathcal {O}_{d_k}$ to the $C^*$ -closure of $\mathcal {A}$ . Note that the closure of $ \bigcup _{n\in \mathbb {N}^k}\mathcal {L}(\mathcal H^n,\mathcal H^n)$ is isomorphic to the core $\mathcal {F}=\mathcal {O}(\mathcal {E})^{\mathbb {T}^k}$ , that is the fixed point algebra under the gauge action, which is a UHF-algebra.

To define the Doplicher–Roberts algebra $\mathcal O_{\rho _1,\ldots ,\rho _k}$ , we again identify $Hom(\rho ^n,\rho ^m)$ with a subset of $Hom(\rho ^{n+r},\rho ^{m+r})$ for each $r\in \mathbb N^k$ , via $T\mapsto T\otimes I_r$ . After this identification, it follows that the linear span ${}^0{\mathcal O}_{\rho _1,\ldots , \rho _k}$ of $ \bigcup _{m,n\in \mathbb {N}^k}Hom(\rho ^n, \rho ^m)\subseteq \mathcal {A}_0$ has a natural multiplication and involution inherited from $\mathcal {A}$ . Indeed, a computation shows that if $S\in Hom(\rho ^n, \rho ^m)$ and $T\in Hom(\rho ^q,\rho ^p)$ , then $S^*\in Hom(\rho ^m, \rho ^n)$ and

$$ \begin{align*} &((S\otimes I_{p\vee n-n})\circ (T\otimes I_{p\vee n-p}))\rho^{q+p\vee n-p}(g)\\ &\quad=\rho^{m+p\vee n-n}(g)((S\otimes I_{p\vee n-n})\circ (T\otimes I_{p\vee n-p})),\end{align*} $$

so $(S\otimes I_{p\vee n-n})\circ (T\otimes I_{p\vee n-p})\in Hom(\rho ^{q+p\vee n-p}, \rho ^{m+p\vee n-n})$ and ${}^0{\mathcal O}_{\rho _1,\ldots , \rho _k}$ is closed under these operations. Since the action of G commutes with the gauge action, there is a natural $\mathbb Z^k$ -grading of ${}^0{\mathcal O}_{\rho _1,\ldots ,\rho _k}$ given by the gauge action $\gamma $ of $\mathbb {T}^k$ on $\mathcal {A}$ .

It follows that the closure ${\mathcal O}_{\rho _1,\ldots , \rho _k}$ of ${}^0{\mathcal O}_{\rho _1,\ldots ,\rho _k}$ in $\mathcal {O}(\mathcal {E})$ is well defined, obtaining the Doplicher–Roberts algebra associated to the representations $\rho _1,\ldots ,\rho _k$ . This $C^*$ -algebra also has a $\mathbb Z^k$ -grading and a gauge action of $\mathbb {T}^k$ . By construction, ${\mathcal O}_{\rho _1,\ldots , \rho _k}\subseteq \mathcal {O}(\mathcal {E})$ .

Remark 3.2. For a compact Lie group G, our Doplicher–Roberts algebra ${\mathcal O}_{\rho _1,\ldots , \rho _k}$ is Morita equivalent with the higher rank Doplicher–Roberts algebra $\mathcal {D}$ defined in [Reference Albandik and Meyer1]. It is also the section $C^*$ -algebra of a Fell bundle over $\mathbb {Z}^k$ .

Theorem 3.3. Let $\rho _i$ be finite-dimensional unitary representations of a compact group G on Hilbert spaces $\mathcal H_i$ of dimensions $d_i\ge 2$ for $i=1,\ldots ,k$ . Then the Doplicher–Roberts algebra ${\mathcal O}_{\rho _1,\ldots ,\rho _k}$ is isomorphic to the fixed point algebra ${\mathcal O}(\mathcal {E})^G\cong (\mathcal O_{d_1}\otimes \cdots \otimes \mathcal O_{d_k})^G$ , where $\mathcal {E}=\mathcal {E}(\rho _1,\ldots ,\rho _k)$ is the product system described in Definition 2.9.

Proof. We know from Lemma 3.1 that ${\mathcal O}(\mathcal {E})$ is isomorphic to the $C^*$ -algebra generated by the linear span of $ \mathcal {A}_0= \bigcup _{m,n\in \mathbb {N}^k}\mathcal {L}({\mathcal H}^n, {\mathcal H}^m)$ . The group G acts on $\mathcal {L}({\mathcal H}^n, {\mathcal H}^m)$ by

$$ \begin{align*}(g\cdot T)(\xi)=\rho^m(g)T(\rho^n(g^{-1})\xi)\end{align*} $$

and the fixed point set is $Hom(\rho ^n, \rho ^m)$ . Indeed, we have $g\cdot T=T$ if and only if $T\rho ^n(g)=\rho ^m(g)T$ . This action is compatible with the embeddings and the operations, so it extends to the $*$ -algebra $\mathcal {A}$ and the fixed point algebra is the linear span of $ \bigcup _{m,n\in \mathbb {N}^k}Hom(\rho ^n, \rho ^m)$ .

It follows that ${}^0{\mathcal O}_{\rho _1,\ldots ,\rho _k}\subseteq {\mathcal O}(\mathcal {E})^G$ and therefore its closure ${\mathcal O}_{\rho _1,\ldots ,\rho _k}$ is isomorphic to a subalgebra of ${\mathcal O}(\mathcal {E})^G$ . For the other inclusion, any element in ${\mathcal O}(\mathcal {E})^G$ can be approximated with an element from ${}^0{\mathcal O}_{\rho _1,\ldots ,\rho _k}$ , and hence ${\mathcal O}_{\rho _1,\ldots ,\rho _k}=\mathcal {O}(\mathcal {E})^G$ .

Remark 3.4. By left tensoring with $I_r$ for $r\in \mathbb {N}^k$ , we obtain some canonical unital endomorphisms $\sigma _r$ of ${\mathcal O}_{\rho _1,\ldots ,\rho _k}$ .

In the next section, we show that, in many cases, $\mathcal O_{\rho _1,\ldots ,\rho _k}$ is isomorphic to a corner of $C^*(\Lambda )$ for a rank k graph $\Lambda $ , so, in some cases, we can compute its K-theory. It would be nice to express the K-theory of $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ in terms of the maps $\pi \mapsto \pi \otimes \rho _i$ defined on the representation ring $\mathcal {R}(G)$ .

4 The rank k graphs

For convenience, we first collect some facts about higher rank graphs, introduced in [Reference Kumjian, Pask, Raeburn and Renault18]. A rank k graph or k-graph $(\Lambda , d)$ consists of a countable small category $\Lambda $ with range and source maps r and s together with a functor $d : \Lambda \rightarrow \mathbb {N}^k$ called the degree map, satisfying the factorization property: for every $\lambda \in \Lambda $ and all $m, n \in \mathbb {N}^k$ with $d( \lambda ) = m + n$ , there are unique elements $\mu , \nu \in \Lambda $ such that $\lambda = \mu \nu $ and $d( \mu ) = m$ , $d( \nu ) = n$ . For $n \in \mathbb {N}^k$ , we write $\Lambda ^n := d^{-1} (n)$ and call it the set of paths of degree n. For $\varepsilon _i=(0,\ldots ,1,\ldots ,0)$ with $1$ in position i, the elements in $\Lambda ^{\varepsilon _i}$ are called edges and the elements in $\Lambda ^0$ are called vertices.

A k-graph $\Lambda $ can be constructed from $\Lambda ^0$ and from its k-colored skeleton $\Lambda ^{\varepsilon _1}\cup \cdots \cup \Lambda ^{\varepsilon _k}$ using a complete and associative collection of commuting squares or factorization rules (see [Reference Sims25]).

The k-graph $\Lambda $ is row-finite if, for all $n\in \mathbb {N}^k$ and all $v\in \Lambda ^0$ , the set $v\Lambda ^n := \{\lambda \in \Lambda ^n : r(\lambda ) = v\}$ is finite. It has no sources if $v\Lambda ^n\neq \emptyset $ for all $v\in \Lambda ^0$ and $n\in \mathbb {N}^k$ . A k-graph $\Lambda $ is said to be irreducible (or strongly connected) if, for every $u,v\in \Lambda ^0$ , there is $\lambda \in \Lambda $ such that $u = r(\lambda )$ and $v = s(\lambda )$ .

Recall that $C^*(\Lambda )$ is the universal $C^*$ -algebra generated by a family $\{S_\lambda : \lambda \in \Lambda \}$ of partial isometries satisfying:

  • $\{S_v:v\in \Lambda ^0\}$ is a family of mutually orthogonal projections;

  • $S_{\lambda \mu }=S_\lambda S_\mu $ for all $\lambda , \mu \in \Lambda $ such that $s(\lambda ) = r(\mu )$ ;

  • $S_\lambda ^*S_\lambda = S_{s(\lambda )}$ for all $\lambda \in \Lambda $ ; and

  • for all $v\in \Lambda ^0$ and $n\in \mathbb {N}^k$ ,

    $$ \begin{align*} S_v=\sum_{\lambda\in v\Lambda^n}S_\lambda S_\lambda^*.\end{align*} $$

A k-graph $\Lambda $ is said to satisfy the aperiodicity condition if, for every vertex $v\in \Lambda ^0$ , there is an infinite path $x\in v\Lambda ^\infty $ such that $\sigma ^mx\neq \sigma ^nx$ for all $m\neq n$ in $\mathbb {N}^k$ , where $\sigma ^m:\Lambda ^\infty \to \Lambda ^\infty $ are the shift maps. We say that $\Lambda $ is cofinal if, for every $x\in \Lambda ^\infty $ and $v\in \Lambda ^0$ , there is $\lambda \in \Lambda $ and $n\in \mathbb {N}^k$ such that $s(\lambda )=x(n)$ and $r(\lambda )=v$ .

Assume that $\Lambda $ is row-finite with no sources and that it satisfies the aperiodicity condition. Then $C^*(\Lambda )$ is simple if and only if $\Lambda $ is cofinal (see Proposition 4.8 in [Reference Kumjian, Pask, Raeburn and Renault18] and Theorem 3.4 in [Reference Robertson and Sims23]).

We say that a path $\mu \in \Lambda $ is a loop with an entrance if $s(\mu )=r(\mu )$ , and there exists $\alpha \in s(\mu )\Lambda $ such that $d(\mu )\ge d(\alpha )$ and there is no $\beta \in \Lambda $ with $\mu = \alpha \beta $ . We say that every vertex connects to a loop with an entrance if, for every $v\in \Lambda ^0$ , there is a loop with an entrance $\mu \in \Lambda $ , and a path $\lambda \in \Lambda $ with $r(\lambda )=v$ and $s(\lambda )=r(\mu )=s(\mu )$ . If $\Lambda $ satisfies the aperiodicity condition and every vertex connects to a loop with an entrance, then $C^*(\Lambda )$ is purely infinite (see Proposition 4.9 in [Reference Kumjian, Pask, Raeburn and Renault18] and Proposition 8.8 in [Reference Sims24]).

Given finite-dimensional unitary representations $\rho _i$ of a compact group G on Hilbert spaces $\mathcal H_i$ for $i=1,\ldots ,k$ , we want to construct a rank k graph $\Lambda =\Lambda (\rho _1,\ldots ,\rho _k)$ . Let R be the set of equivalence classes of irreducible summands $\pi :G\to U(\mathcal {H}_\pi )$ which appear in the tensor powers $\rho ^n=\rho _1^{\otimes n_1}\otimes \cdots \otimes \rho _k^{\otimes n_k}$ for $n\in \mathbb {N}^k$ , as in [Reference Mann, Raeburn and Sutherland22]. Take $\Lambda ^0=R$ and, for each $i=1,\ldots ,k$ , consider the set of edges $\Lambda ^{\varepsilon _i}$ which are uniquely determined by the matrices $M_i$ with entries

$$ \begin{align*}M_i(w,v)=|\{e\in \Lambda^{\varepsilon_i}: s(e)=v, r(e)=w\}|=\dim Hom(v,w\otimes \rho_i),\end{align*} $$

where $v,w\in R$ . The matrices $M_i$ commute since $\rho _i\otimes \rho _j\cong \rho _j\otimes \rho _i$ and therefore

$$ \begin{align*}\dim Hom(v,w\otimes \rho_i\otimes\rho_j)=\dim Hom(v,w\otimes \rho_j\otimes\rho_i)\end{align*} $$

for all $i<j$ . This allows us to fix some bijections

$$ \begin{align*}\lambda_{ij}:\Lambda^{\varepsilon_i}\times_{\Lambda^0}\Lambda^{\varepsilon_j}\to \Lambda^{\varepsilon_j}\times_{\Lambda^0}\Lambda^{\varepsilon_i}\end{align*} $$

for all $1\le i<j\le k$ , which determine the commuting squares of $\Lambda $ . As usual,

$$ \begin{align*}\Lambda^{\varepsilon_i}\times_{\Lambda^0}\Lambda^{\varepsilon_j}=\{(e,f)\in \Lambda^{\varepsilon_i}\times \Lambda^{\varepsilon_j}: s(e)=r(f)\}.\end{align*} $$

For $k\ge 3$ , we also need to verify that $\lambda _{ij}$ can be chosen to satisfy the associativity condition, that is,

$$ \begin{align*}(id_\ell\times \lambda_{ij})(\lambda_{i\ell}\times id_j)(id_i\times \lambda_{j\ell})=(\lambda_{j\ell}\times id_i)(id_j\times \lambda_{i\ell})(\lambda_{ij}\times id_\ell)\end{align*} $$

as bijections from $\Lambda ^{\varepsilon _i}\times _{\Lambda ^0}\Lambda ^{\varepsilon _j}\times _{\Lambda ^0}\Lambda ^{\varepsilon _\ell }$ to $\Lambda ^{\varepsilon _\ell }\times _{\Lambda ^0}\Lambda ^{\varepsilon _j}\times _{\Lambda ^0}\Lambda ^{\varepsilon _i}$ for all $i<j<\ell $ .

Remark 4.1. Many times $R=\hat {G}$ , so $\Lambda ^0=\hat {G}$ , for example, if $\rho _i$ are faithful and $\rho _i(G)\subseteq SU(\mathcal {H}_i)$ or if G is finite, $\rho _i$ are faithful and $\dim \rho _i\ge 2$ for all $i=1,\ldots ,k$ (see Lemma 7.2 and Remark 7.4 in [Reference Kajiwara, Pinzari and Watatani19]).

Proposition 4.2. Given representations $\rho _1,\ldots ,\rho _k$ as above, assume that $\rho _i$ are faithful and that $R=\hat {G}$ . Then each choice of bijections $\lambda _{ij}$ satisfying the associativity condition determines a rank k graph $\Lambda $ which is cofinal and locally finite with no sources.

Proof. Indeed, the sets $\Lambda ^{\varepsilon _i}$ are uniquely determined and the choice of bijections $\lambda _{ij}$ satisfying the associativity condition is enough to determine $\Lambda $ . Since the entries of the matrices $M_i$ are finite and there are no zero rows, the graph is locally finite with no sources. To prove that $\Lambda $ is cofinal, fix a vertex $v\in \Lambda ^0$ and an infinite path $x\in \Lambda ^\infty $ . Arguing as in Lemma 7.2 in [Reference Kajiwara, Pinzari and Watatani19], any $w\in \Lambda ^0$ , in particular, $w=x(n)$ for a fixed n, can be joined by a path to v, so there is $\lambda \in \Lambda $ with $s(\lambda )=x(n)$ and $r(\lambda )=v$ . See also Lemma 3.1 in [Reference Mann, Raeburn and Sutherland22].

Remark 4.3. Note that the entry $M_i(w,v)$ is just the multiplicity of the irreducible representation v in $w\otimes \rho _i$ for $i=1,\ldots ,k$ . If $\rho _i^*=\rho _i$ , then the matrices $M_i$ are symmetric since

$$ \begin{align*}\dim Hom(v, w\otimes \rho_i)=\dim Hom(\rho^*_i\otimes v,w)\end{align*} $$

which implies $M_i(w; v) = M_i(v;w)$ . Here $\rho ^*_i$ denotes the dual representation defined by $\rho _i^*(g)=\rho _i(g^{-1})^t$ and equal, in our case, to the conjugate representation $\bar {\rho _i}$ .

For G finite, these matrices are finite, and the entries $M_i(w,v)$ can be computed using the character table of G. For G infinite, the Clebsch–Gordan relations can be used to determine the numbers $M_i(w,v)$ . Since the bijections $\lambda _{ij}$ are, in general, not unique, the rank k graph $\Lambda $ is not unique, as illustrated in some examples. It is an open question how the $C^*$ -algebra $C^*(\Lambda )$ depends, in general, on the factorization rules.

To relate the Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ to a rank k graph $\Lambda $ , we mimic the construction in [Reference Mann, Raeburn and Sutherland22]. For each edge $e\in \Lambda ^{\varepsilon _i}$ , choose an isometric intertwiner

$$ \begin{align*}T_e: \mathcal{H}_{s(e)}\to \mathcal{H}_{r(e)}\otimes \mathcal{H}_i\end{align*} $$

in such a way that

$$ \begin{align*}\mathcal{H}_\pi\otimes \mathcal{H}_i=\bigoplus_{e\in \pi\Lambda^{\varepsilon_i}}T_eT_e^*(\mathcal{H}_\pi\otimes \mathcal{H}_i)\end{align*} $$

for all $\pi \in \Lambda ^0$ , that is, the edges in $\Lambda ^{\varepsilon _i}$ ending at $\pi $ give a specific decomposition of $\mathcal {H}_\pi \otimes \mathcal {H}_i$ into irreducibles. When $\dim Hom(s(e), r(e)\otimes \rho _i)\ge 2$ , we must choose a basis of isometric intertwiners with orthogonal ranges, so, in general, $T_e$ is not unique. In fact, specific choices for the isometric intertwiners $T_e$ determine the factorization rules in $\Lambda $ and whether or not they satisfy the associativity condition.

Given $e\in \Lambda ^{\varepsilon _i}$ and $f\in \Lambda ^{\varepsilon _j}$ with $r(f)=s(e)$ , we know how to multiply $T_e\in Hom(s(e),r(e)\otimes \rho _i)$ with $T_f\in Hom(s(f),r(f)\otimes \rho _j)$ in the algebra $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ , by viewing $Hom(s(e),r(e)\otimes \rho _i)$ as a subspace of $Hom(\rho ^n,\rho ^m)$ for some $m,n$ , and similarly for $Hom(s(f),r(f)\otimes \rho _j)$ . We choose edges $e'\in \Lambda ^{\varepsilon _i}, f'\in \Lambda ^{\varepsilon _j}$ with $s(f)=s(e'), r(e)=r(f'), r(e')=s(f')$ such that $T_eT_f=T_{f'}T_{e'}$ , where $T_{f'}\in Hom(s(f'),r(f')\otimes \rho _j)$ and $T_{e'}\in Hom(s(e'),r(e')\otimes \rho _i)$ . This is possible since

$$ \begin{align*}T_eT_f=(T_e\otimes I_j)\circ T_f\in Hom(s(f),r(e)\otimes \rho_i\otimes \rho_j),\end{align*} $$
$$ \begin{align*}T_{f'}T_{e'}=(T_{f'}\otimes I_i)\circ T_{e'}\in Hom(s(e'),r(f')\otimes \rho_j\otimes \rho_i),\end{align*} $$

and $\rho _i\otimes \rho _j\cong \rho _j\otimes \rho _i$ . In this case, we declare that $ef=f'e'$ . Repeating this process, we obtain bijections $\lambda _{ij}:\Lambda ^{\varepsilon _i}\times _{\Lambda ^0}\Lambda ^{\varepsilon _j}\to \Lambda ^{\varepsilon _j}\times _{\Lambda ^0}\Lambda ^{\varepsilon _i}$ . Assuming that the associativity conditions are satisfied, we obtain a k-graph $\Lambda $ .

We write $T_{ef}=T_eT_f=T_{f'}T_{e'}=T_{f'e'}$ . A finite path $\lambda \in \Lambda ^n$ is a concatenation of edges and determines by composition a unique intertwiner

$$ \begin{align*}T_\lambda:\mathcal{H}_{s(\lambda)}\to \mathcal{H}_{r(\lambda)}\otimes\mathcal{H}^n.\end{align*} $$

Moreover, the paths $\lambda \in \Lambda ^n$ with $r(\lambda )=\iota $ , the trivial representation, provide an explicit decomposition of $\mathcal {H}^n=\mathcal {H}_1^{\otimes n_1}\otimes \cdots \otimes \mathcal {H}_k^{\otimes n_k}$ into irreducibles, and hence

$$ \begin{align*}\mathcal{H}^n=\bigoplus_{\lambda\in\iota \Lambda^n}T_\lambda T_\lambda^*(\mathcal{H}^n).\end{align*} $$

Proposition 4.4. Assuming that the choices of isometric intertwiners $T_e$ , as above, determine a k-graph $\Lambda $ , the family

$$ \begin{align*}\{T_\lambda T^*_\mu: \lambda\in\Lambda^m, \mu\in\Lambda^n, r(\lambda)=r(\mu)=\iota, s(\lambda)=s(\mu)\}\end{align*} $$

is a basis for $Hom(\rho ^n, \rho ^m)$ and each $T_\lambda T^*_\mu $ is a partial isometry.

Proof. Each pair of paths $\lambda , \mu $ with $d(\lambda )=m, d(\mu )=n$ and $r(\lambda )=r(\mu )=\iota $ determines a pair of irreducible summands $T_\lambda (\mathcal {H}_{s(\lambda )}), T_\mu (\mathcal {H}_{s(\mu )})$ of $\mathcal {H}^m$ and $ \mathcal {H}^n$ , respectively. By Schur’s lemma, the space of intertwiners of these representations is trivial unless $s(\lambda )=s(\mu )$ , in which case it is the one-dimensional space spanned by $T_\lambda T_\mu ^*$ . It follows that any element of $Hom(\rho ^n, \rho ^m)$ can be uniquely represented as a linear combination of elements $T_\lambda T_\mu ^*$ , where $s(\lambda )=s(\mu )$ . Since $T_\mu $ is isometric, $T_\mu ^*$ is a partial isometry with range $\mathcal {H}_{s(\mu )}$ and hence $T_\lambda T_\mu ^*$ is also a partial isometry whenever $s(\lambda )=s(\mu )$ .

Theorem 4.5. Consider $\rho _1,\ldots , \rho _k$ finite-dimensional unitary representations of a compact group G and let $\Lambda $ be the k-colored graph with $\Lambda ^0=R\subseteq \hat {G}$ and edges $\Lambda ^{\varepsilon _i}$ determined by the incidence matrices $M_i$ defined above. Assume that the factorization rules determined by the choices of $T_e\in Hom(s(e),r(e)\otimes \rho _i)$ for all edges $e\in \Lambda ^{\varepsilon _i}$ satisfy the associativity condition, so $\Lambda $ becomes a rank k graph. If we consider $P\in C^*(\Lambda )$ ,

$$ \begin{align*}P=\sum_{\lambda\in\iota \Lambda^{(1,\ldots,1)}}S_\lambda S_\lambda^*,\end{align*} $$

where $\iota $ is the trivial representation, then there is a $*$ -isomorphism of the Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ onto the corner $PC^*(\Lambda )P$ .

Proof. Since $C^*(\Lambda )$ is generated by linear combinations of $S_\lambda S_\mu ^*$ with $s(\lambda )=s(\mu )$ (see Lemma 3.1 in [Reference Kumjian, Pask, Raeburn and Renault18]), we first define the maps

$$ \begin{align*}\phi_{n,m}:Hom(\rho^n, \rho^m)\to C^*(\Lambda),\quad \phi_{n,m}(T_\lambda T_\mu^*)=S_\lambda S_\mu^*,\end{align*} $$

where $s(\lambda )=s(\mu )$ and $r(\lambda )=r(\mu )=\iota $ . Since $S_\lambda S_\mu ^*=PS_\lambda S_\mu ^*P$ , the maps $\phi _{n,m}$ take values in $PC^*(\Lambda )P$ . We claim that, for any $r\in \mathbb {N}^k$ ,

$$ \begin{align*}\phi_{n+r,m+r}(T_\lambda T_\mu^*\otimes I_r)=\phi_{n,m}(T_\lambda T_\mu^*).\end{align*} $$

This is because

$$ \begin{align*}\mathcal{H}_{s(\lambda)}\otimes \mathcal{H}^r=\bigoplus_{\nu\in s(\lambda)\Lambda^r}T_\nu T_\nu^*(\mathcal{H}_{s(\lambda)}\otimes \mathcal{H}^r),\end{align*} $$

so that

$$ \begin{align*}T_\lambda T_\mu^*\otimes I_r=\sum_{\nu\in s(\lambda)\Lambda^r}(T_\lambda\otimes I_r)(T_\nu T_\nu^*)(T_\mu^*\otimes I_r)=\sum_ {\nu\in s(\lambda)\Lambda^r} T_{\lambda\nu}T_{\mu\nu}^*\end{align*} $$

and

$$ \begin{align*}S_\lambda S_\mu^*=\sum_{\nu\in s(\lambda)\Lambda^r}S_\lambda(S_\nu S_\nu^*)S_\mu^*=\sum_{\nu\in s(\lambda)\Lambda^r}S_{\lambda\nu}S_{\mu\nu}^*.\end{align*} $$

The maps $\phi _{n,m}$ determine a map $\phi : {}^0\mathcal {O}_{\rho _1,\ldots ,\rho _k}\to PC^*(\Lambda )P$ which is linear, $*$ -preserving and multiplicative. Indeed,

$$ \begin{align*}\phi_{n,m}(T_\lambda T_\mu^*)^*=(S_\lambda S_\mu^*)^*=S_\mu S_\lambda^*=\phi_{m,n}(T_\mu T_\lambda^*).\end{align*} $$

Consider now $T_\lambda T_\mu ^*\in Hom(\rho ^n, \rho ^m),\;\; T_\nu T_\omega ^*\in Hom (\rho ^q,\rho ^p)$ with $s(\lambda )=s(\mu ), s(\nu )=s(\omega ), r(\lambda )=r(\mu )=r(\nu )=r(\omega )=\iota $ . Since, for all $n\in \mathbb {N}^k$ ,

$$ \begin{align*}\sum_{\lambda\in\iota\Lambda^n}T_\lambda T_\lambda^*=I_n,\end{align*} $$

we get

$$ \begin{align*}T_\mu^*T_\nu= \begin{cases} T_\beta^*\quad& \text{if } \mu=\nu\beta,\\ T_\alpha \quad& \text{if } \nu=\mu\alpha,\\0 \quad&\text{otherwise,}\end{cases}\end{align*} $$

and hence

$$ \begin{align*}\phi((T_\lambda T_\mu^*)(T_\nu T_\omega^*))=\begin{cases} \phi(T_\lambda T_{\omega\beta}^*)=S_\lambda S_{\omega\beta}^*\quad&\text{if } \mu=\nu\beta,\\ \phi(T_{\lambda\alpha}T_\omega^*)=S_{\lambda\alpha}S_\omega^*\quad&\text{if } \nu=\mu\alpha,\\0 \quad&\text{otherwise.}\end{cases}\end{align*} $$

On the other hand, from Lemma 3.1 in [Reference Kumjian, Pask, Raeburn and Renault18],

$$ \begin{align*}S_\lambda S_\mu^* S_\nu S_{\omega}^*=\begin{cases} S_\lambda S_{\omega\beta}^*\quad&\text{if } \mu=\nu\beta,\\ S_{\lambda\alpha}S_\omega^*\quad&\text{if } \nu=\mu\alpha,\\0 \quad&\text{otherwise,}\end{cases}\end{align*} $$

and hence

$$ \begin{align*}\phi((T_\lambda T_\mu^*)(T_\nu T_\omega^*))=\phi(T_\lambda T_\mu^*)\phi(T_\nu T_\omega^*).\end{align*} $$

Since $PS_\lambda S_\mu ^*P=\phi _{n,m}(T_\lambda T_\mu ^*)$ if $r(\lambda )=r(\mu )=\iota $ and $s(\lambda )=s(\mu )$ , it follows that $\phi $ is surjective. Injectivity follows from the fact that $\phi $ is equivariant for the gauge action.

Corollary 4.6. If the k-graph $\Lambda $ associated to $\rho _1,\ldots ,\rho _k$ is cofinal, satisfies the aperiodicity condition and every vertex connects to a loop with an entrance, then the Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ is simple and purely infinite, and is Morita equivalent with $C^*(\Lambda )$ .

Proof. This follows from the fact that $C^*(\Lambda )$ is simple and purely infinite and because $PC^*(\Lambda )P$ is a full corner.

Remark 4.7. There is a groupoid $\mathcal {G}_\Lambda $ associated to a row-finite rank k graph $\Lambda $ with no sources (see [Reference Kumjian, Pask, Raeburn and Renault18]). By taking the pointed groupoid $\mathcal {G}_\Lambda (\iota )$ , the reduction to the set of infinite paths with range $\iota $ , under the same conditions as in Theorem 4.5, we get an isomorphism of the Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\ldots ,\rho _k}$ onto $C^*(\mathcal {G}_\Lambda (\iota ))$ .

5 Examples

Example 5.1. Let $G=S_3$ be the symmetric group with $\hat {G}=\{\iota , \epsilon , \sigma \}$ and character table

Here $\iota $ denotes the trivial representation, $\epsilon $ is the sign representation and $\sigma $ is an irreducible $2$ -dimensional representation, for example,

$$ \begin{align*}\sigma((12))=\left[\begin{array}{rr}-1&-1\\0&1\end{array}\right],\quad\sigma((123))=\left[\begin{array}{rr}-1&-1\\1&0\end{array}\right].\end{align*} $$

By choosing $\rho _1=\sigma $ on $\mathcal {H}_1=\mathbb {C}^2$ and $\rho _2=\iota +\sigma $ on $\mathcal {H}_2=\mathbb {C}^3$ , we get a product system $\mathcal {E}\to \mathbb {N}^2$ and an action of $S_3$ on $\mathcal {O}(\mathcal {E})\cong \mathcal O_2\otimes \mathcal O_3$ with fixed point algebra $\mathcal {O}(\mathcal {E})^{S_3}\cong \mathcal {O}_{\rho _1,\rho _2}$ isomorphic to a corner of the $C^*$ -algebra of a rank two graph $\Lambda $ . The set of vertices is $\Lambda ^0=\{\iota ,\epsilon , \sigma \}$ and the edges are given by the incidence matrices

$$ \begin{align*} M_1=\left[\begin{array}{ccc}0&0&1\\0&0&1\\1&1&1\end{array}\right], \quad M_2=\left[\begin{array}{ccc}1&0&1\\0&1&1\\1&1&2\end{array}\right].\end{align*} $$

This is because

$$ \begin{align*}\iota\otimes\rho_1=\sigma,\; \epsilon\otimes \rho_1=\sigma,\; \sigma\otimes \rho_1=\iota+\epsilon+\sigma,\end{align*} $$
$$ \begin{align*}\iota\otimes\rho_2=\iota+\sigma,\; \epsilon\otimes\rho_2=\epsilon+\sigma,\; \sigma\otimes\rho_2=\iota+\epsilon+2\sigma.\end{align*} $$

We label the blue (solid) edges by $e_1,\ldots , e_5$ and the red (dashed) edges by $f_1,\ldots ,f_8$ as in the figure below.

The isometric intertwiners are

$$ \begin{align*}T_{e_1}:\mathcal{H}_\iota\to \mathcal{H}_\sigma\otimes \mathcal{H}_1, \; T_{e_2}:\mathcal{H}_\sigma\to \mathcal{H}_\iota\otimes \mathcal{H}_1, \;T_{e_3}:\mathcal{H}_\epsilon\to \mathcal{H}_\sigma\otimes \mathcal{H}_1,\end{align*} $$
$$ \begin{align*}T_{e_4}:\mathcal{H}_\sigma\to \mathcal{H}_\epsilon\otimes \mathcal{H}_1,\; T_{e_5}:\mathcal{H}_\sigma\to\mathcal{H}_\sigma\otimes\mathcal{H}_1,\end{align*} $$
$$ \begin{align*}T_{f_1}:\mathcal{H}_\iota\to \mathcal{H}_\iota\otimes\mathcal{H}_2,\; T_{f_2}:\mathcal{H}_\epsilon\to \mathcal{H}_\epsilon\otimes\mathcal{H}_2,\; T_{f_3}:\mathcal{H}_\sigma\to \mathcal{H}_\iota\otimes\mathcal{H}_2,\end{align*} $$
$$ \begin{align*}T_{f_4}:\mathcal{H}_\iota\to \mathcal{H}_\sigma\otimes\mathcal{H}_2,\; T_{f_5}:\mathcal{H}_\sigma\to\mathcal{H}_\epsilon\otimes \mathcal{H}_2,\; T_{f_6}:\mathcal{H}_\epsilon\to\mathcal{H}_\sigma\otimes \mathcal{H}_2,\end{align*} $$
$$ \begin{align*} T_{f_7}, T_{f_8}:\mathcal{H}_\sigma\to \mathcal{H}_\sigma\otimes\mathcal{H}_2\end{align*} $$

such that

$$ \begin{align*}T_{e_1}T_{e_1}^*+T_{e_3}T_{e_3}^*+T_{e_5}T_{e_5}^*=I_\sigma\otimes I_1, \; T_{e_2}T_{e_2}^*=I_\iota\otimes I_1,\; T_{e_4}T_{e_4}^*=I_\epsilon\otimes I_1,\end{align*} $$
$$ \begin{align*}T_{f_1}T_{f_1}^*+T_{f_3}T_{f_3}^*=I_\iota\otimes I_2,\; T_{f_2}T_{f_2}^*+T_{f_5}T_{f_5}^*=I_\epsilon\otimes I_2,\end{align*} $$
$$ \begin{align*} T_{f_4}T_{f_4}^*+T_{f_6}T_{f_6}^*+T_{f_7}T_{f_7}^*+T_{f_8}T_{f_8}^*=I_\sigma\otimes I_2.\end{align*} $$

Here $I_\pi $ is the identity of $\mathcal {H}_\pi $ for $\pi \in \hat {G}$ and $I_i$ is the identity of $\mathcal {H}_i$ for $ i=1,2$ . Since

$$ \begin{align*}M_1M_2=\left[\begin{array}{ccc}1&1&2\\1&1&2\\2&2&4\end{array}\right]\end{align*} $$

and

$$ \begin{align*}T_{e_2}T_{f_4}, T_{f_3}T_{e_1}\in Hom(\iota, \iota\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_2}T_{f_6}, T_{f_3}T_{e_3}\in Hom(\epsilon, \iota\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_2}T_{f_7}, T_{e_2}T_{f_8}, T_{f_1}T_{e_2}, T_{f_3}T_{e_5}\in Hom(\sigma, \iota\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_4}T_{f_4}, T_{f_5}T_{e_1}\in Hom(\iota, \epsilon\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_4}T_{f_6}, T_{f_5}T_{e_3}\in Hom(\epsilon, \epsilon\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_4}T_{f_7}, T_{e_4}T_{f_8}, T_{f_2}T_{e_4}, T_{f_5}T_{e_5}\in Hom(\sigma, \epsilon\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_1}T_{f_1}, T_{e_5}T_{f_4}, T_{f_7}T_{e_1}, T_{f_8}T_{e_1}\in Hom(\iota, \sigma\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_3}T_{f_2}, T_{e_5}T_{f_6}, T_{f_7}T_{e_3}, T_{f_8}T_{e_3}\in Hom(\epsilon,\sigma\otimes\rho_1\otimes\rho_2),\end{align*} $$
$$ \begin{align*}T_{e_5}T_{f_7}, T_{e_5}T_{f_8}, T_{e_3}T_{f_5}, T_{e_1}T_{f_3}, T_{f_6}T_{e_4}, T_{f_4}T_{e_2}, T_{f_7}T_{e_5}, T_{f_8}T_{e_5}\in Hom(\sigma, \sigma\otimes\rho_1\otimes\rho_2),\end{align*} $$

a possible choice of commuting squares is

$$ \begin{align*}e_2f_4=f_3e_1,\; e_2f_6=f_3e_3,\; e_2f_7=f_1e_2,\; e_2f_8=f_3e_5,\; e_4f_4=f_5e_1,\; e_4f_6=f_5e_3,\end{align*} $$
$$ \begin{align*}e_4f_7= f_2e_4,\; e_4f_8=f_5e_5,\; e_1f_1=f_7e_1,\; e_5f_4=f_8e_1,\; e_3f_2=f_7e_3,\; e_5f_6=f_8e_3,\end{align*} $$
$$ \begin{align*}e_5f_7=f_6e_4,\; e_5f_8=f_4e_2,\; e_3f_5=f_7e_5,\; e_1f_3=f_8e_5.\end{align*} $$

This data is enough to determine a rank two graph $\Lambda $ associated to $\rho _1, \rho _2$ . But this is not the only choice, since, for example, we could have taken

$$ \begin{align*}e_2f_4=f_3e_1,\; e_2f_6=f_3e_3,\; e_2f_8=f_1e_2,\; e_2f_7=f_3e_5,\; e_4f_4=f_5e_1,\; e_4f_6=f_5e_3,\end{align*} $$
$$ \begin{align*}e_4f_8= f_2e_4,\; e_4f_7=f_5e_5,\; e_1f_1=f_7e_1,\; e_5f_4=f_8e_1,\; e_3f_2=f_8e_3,\; e_5f_6=f_7e_3,\end{align*} $$
$$ \begin{align*}e_5f_7=f_6e_4,\; e_5f_8=f_4e_2,\; e_3f_5=f_7e_5,\; e_1f_3=f_8e_5,\end{align*} $$

which determines a different $2$ -graph.

A direct analysis using the definitions shows that, in each case, the $2$ -graph $\Lambda $ is cofinal, satisfies the aperiodicity condition and every vertex connects to a loop with an entrance. It follows that $C^*(\Lambda )$ is simple and purely infinite and the Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\rho _2}$ is Morita equivalent with $C^*(\Lambda )$ .

The K-theory of $C^*(\Lambda )$ can be computed using Proposition 3.16 in [Reference Evans11] and it does not depend on the choice of factorization rules. We have

$$ \begin{align*}K_0(C^*(\Lambda))\cong\text{coker}[I-M_1^t\;\; I-M_2^t]\oplus\ker\left[\begin{array}{c}M_2^t-I\\I-M_1^t\end{array}\right]\cong \mathbb Z/2\mathbb Z,\end{align*} $$
$$ \begin{align*}K_1(C^*(\Lambda))\cong\ker[I-M_1^t\;\;I-M_2^t]/\text{im}\left[\begin{array}{c}M_2^t-I\\I-M_1^t\end{array}\right]\cong 0.\end{align*} $$

In particular, $\mathcal {O}_{\rho _1,\rho _2}\cong \mathcal {O}_3$ .

On the other hand, since $\rho _1, \rho _2$ are faithful, both Doplicher–Roberts algebras $\mathcal {O}_{\rho _1}, \mathcal {O}_{\rho _2}$ are simple and purely infinite with

$$ \begin{align*}K_0(\mathcal{O}_{\rho_1})\cong \mathbb{Z}/2\mathbb{Z},\; K_1(\mathcal{O}_{\rho_1})\cong 0,\; K_0(\mathcal{O}_{\rho_2})\cong \mathbb{Z},\; K_1(\mathcal{O}_{\rho_2})\cong \mathbb{Z},\end{align*} $$

so $\mathcal {O}_{\rho _1,\rho _2}\ncong \mathcal {O}_{\rho _1}\otimes \mathcal {O}_{\rho _2}$ .

Example 5.2. With $G=S_3$ and $\rho _1=2\iota , \rho _2=\iota +\epsilon $ , then $R=\{\iota , \epsilon \}$ , so $\Lambda $ has two vertices and incidence matrices

$$ \begin{align*}M_1=\left[\begin{array}{cc}2&0\\0&2\end{array}\right],\quad M_2=\left[\begin{array}{cc}1&1\\1&1\end{array}\right],\end{align*} $$

which give

Again, a corresponding choice of isometric intertwiners determines some factorization rules, for example,

$$ \begin{align*}e_1f_1=f_1e_2,\; e_2f_1=f_1e_1,\; e_1f_3=f_3e_3,\; e_2f_3=f_3e_4,\end{align*} $$
$$ \begin{align*}e_3f_2=f_2e_1,\; e_4f_2=f_2e_2,\; e_3f_4=f_4e_4,\; e_4f_4=f_4e_3.\end{align*} $$

Even though $\rho _1, \rho _2$ are not faithful, the obtained $2$ -graph is cofinal, satisfies the aperiodicity condition and every vertex connects to a loop with an entrance, so $\mathcal {O}_{\rho _1,\rho _2}$ is simple and purely infinite with trivial K-theory. In particular, $\mathcal {O}_{\rho _1,\rho _2}\cong \mathcal {O}_2$ .

Note that, since $\rho _1, \rho _2$ have kernel $N=\langle (123)\rangle \cong \mathbb {Z}/3\mathbb {Z}$ , we could replace G by $G/N\cong \mathbb {Z}/2\mathbb {Z}$ and consider $\rho _1,\rho _2$ as representations of $\mathbb {Z}/2\mathbb {Z}$ .

Example 5.3. Consider $G=\mathbb {Z}/2\mathbb {Z}=\{0,1\}$ with $\hat {G}=\{\iota ,\chi \}$ and character table

Choose the $2$ -dimensional representations

$$ \begin{align*}\rho_1=\iota+\chi,\; \rho_2=2\iota,\; \rho_3=2\chi,\end{align*} $$

which determine a product system $\mathcal {E}$ such that $\mathcal {O}(\mathcal {E})\cong \mathcal {O}_2\otimes \mathcal {O}_2\otimes \mathcal {O}_2$ and a Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\rho _2,\rho _3}\cong \mathcal {O}(\mathcal {E})^{\mathbb {Z}/2\mathbb {Z}}$ .

An easy computation shows that the incidence matrices of the blue (solid), red (dashed) and green (dotted) graphs are

$$ \begin{align*}M_1=\left[\begin{array}{cc}1&1\\1&1\end{array}\right],\quad M_2=\left[\begin{array}{cc}2&0\\0&2\end{array}\right],\quad M_3=\left[\begin{array}{cc}0&2\\2&0\end{array}\right].\end{align*} $$

With labels as in the figure, we choose the following factorization rules.

$$ \begin{align*}e_1f_1=f_2e_1,\; e_1f_2=f_1e_1,\; e_2f_1=f_4e_2,\; e_2f_2=f_3e_2,\end{align*} $$
$$ \begin{align*}e_3f_3=f_2e_3,\; e_3f_4=f_1e_3,\; e_4f_4=f_3e_4,\; e_4f_3=f_4e_4,\end{align*} $$
$$ \begin{align*}f_1g_1=g_2f_3,\; f_1g_2=g_1f_3,\; f_2g_1=g_2f_4,\; f_2g_2=g_1f_4,\end{align*} $$
$$ \begin{align*}f_3g_3=g_4f_1,\; f_3g_4=g_3f_1,\; f_4g_3=g_4f_2,\; f_4g_4=g_3f_2,\end{align*} $$
$$ \begin{align*}e_1g_1=g_2e_4,\; e_1g_2=g_1e_4,\; e_2g_1=g_3e_3,\; e_2g_2=g_4e_3,\end{align*} $$
$$ \begin{align*}e_3g_3=g_1e_2,\; e_3g_4=g_2e_2,\; e_4g_3=g_4e_1,\; e_4g_4=g_3e_1.\end{align*} $$

A tedious verification shows that all the following paths are well defined.

$$ \begin{align*}e_1f_1g_1,\; e_1f_1g_2,\; e_1f_2g_1, \; e_1f_2g_2,\; e_2f_1g_1,\; e_2f_1g_2,\; e_2f_2g_1,\; e_2f_2g_2,\end{align*} $$
$$ \begin{align*}e_3f_3g_3,\; e_3f_3g_4,\; e_3f_4g_3,\; e_3f_4g_4,\; e_4f_3g_3,\; e_4f_3g_4,\; e_4f_4g_3,\; e_4f_4g_4,\end{align*} $$

so the associativity property is satisfied and we get a rank three graph $\Lambda $ with two vertices. It is not difficult to check that $\Lambda $ is cofinal, satisfies the aperiodicity condition and every vertex connects to a loop with an entrance, so $C^*(\Lambda )$ is simple and purely infinite.

Since $\partial _1=[I-M_1^t\; I-M_2^t\; I-M_3^t]:\mathbb {Z}^6\to \mathbb {Z}^2$ is surjective, using Corollary 3.18 in [Reference Evans11], we obtain

$$ \begin{align*}K_0(C^*(\Lambda))\cong \ker\partial_2/\text{im}\; \partial_3\cong 0,\;K_1(C^*(\Lambda))\cong \ker\partial_1/\text{im}\; \partial_2\oplus \ker\partial_3\cong 0,\end{align*} $$

where

$$ \begin{align*}\partial_2=\left[\begin{array}{ccc} M_2^t-I&M_3^t-I&0\\I-M_1^t&0&M_3^t-I\\0&I-M_1^t&I-M_2^t\end{array}\right],\quad\partial_3=\left[\begin{array}{c}I-M_3^t\\M_2^t-I\\I-M_1^t\end{array}\right],\end{align*} $$

and, in particular, $\mathcal {O}_{\rho _1,\rho _2,\rho _3}\cong \mathcal {O}_2$ .

Example 5.4. Let $G=\mathbb {T}$ . We have $\hat {G}=\{\chi _k:k\in \mathbb {Z}\}$ , where $\chi _k(z)=z^k$ and $\chi _k\otimes \chi _\ell =\chi _{k+\ell }$ . The faithful representations

$$ \begin{align*}\rho_1=\chi_{-1}+\chi_0,\; \rho_2=\chi_0+\chi_1\end{align*} $$

of $\mathbb {T}$ determine a product system $\mathcal {E}$ with $\mathcal {O}(\mathcal {E})\cong \mathcal {O}_2\otimes \mathcal {O}_2$ and a Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\rho _2}\cong \mathcal {O}(\mathcal {E})^{\mathbb {T}}$ isomorphic to a corner in the $C^*$ -algebra of a rank $2$ graph $\Lambda $ with $\Lambda ^0=\hat {G}$ and infinite incidence matrices, where

$$ \begin{align*}M_1(\chi_k,\chi_\ell)=\begin{cases}1 \quad&\text{if } \ell=k \;\text{or}\; \ell=k-1,\\0\quad& \text{otherwise,}\end{cases}\end{align*} $$
$$ \begin{align*}M_2(\chi_k,\chi_\ell)=\begin{cases} 1 \quad&\text{if } \ell=k \;\text{or}\; \ell=k+1,\\0\quad& \text{otherwise.}\end{cases}\end{align*} $$

The skeleton of $\Lambda $ looks like

and this $2$ -graph is cofinal, satisfies the aperiodicity condition and every vertex connects to a loop with an entrance, so $C^*(\Lambda )$ is simple and purely infinite.

Example 5.5. Let $G=SU(2)$ . It is known (see page 84 in [Reference Bröcker and tom Dieck2]) that the elements in $\hat {G}$ are labeled by $V_n$ for $n\ge 0$ , where $V_0=\iota $ is the trivial representation on $\mathbb {C}$ , $V_1$ is the standard representation of $SU(2)$ on $\mathbb {C}^2$ , and, for $n\ge 2$ , $V_n=S^nV_1$ , the $n\, $ th symmetric power. In fact, $\dim V_n=n+1$ and $V_n$ can be taken as the representation of $SU(2)$ on the space of homogeneous polynomials p of degree n in variables $z_1,z_2$ , where, for $ g=\big [\begin {smallmatrix}a&b\\c&d\end {smallmatrix}\big ]\in SU(2)$ ,

$$ \begin{align*}(g\cdot p)(z)=p(az_1+cz_2, bz_1+dz_2).\end{align*} $$

The irreducible representations $V_n$ satisfy the Clebsch–Gordan formula

$$ \begin{align*}V_k\otimes V_\ell=\bigoplus_{j=0}^qV_{k+\ell-2j},\; q=\min\{k,\ell\}.\end{align*} $$

If we choose $\rho _1=V_1, \rho _2=V_2$ , then we get a product system $\mathcal {E}$ with $\mathcal {O}(\mathcal {E})\cong \mathcal {O}_2\otimes \mathcal {O}_3$ and a Doplicher–Roberts algebra $\mathcal {O}_{\rho _1,\rho _2}\cong \mathcal {O}(\mathcal {E})^{SU(2)}$ isomorphic to a corner in the $C^*$ -algebra of a rank two graph with $\Lambda ^0=\hat {G}$ and edges given by the matrices

$$ \begin{align*}M_1(V_k,V_\ell)=\begin{cases}1\quad&\text{if } k=0\;\text{and}\; \ell=1,\\ 1\quad& \text{if } k\ge 1\;\text{and}\; \ell \in\{k-1,k+1\},\\0\quad&\text{otherwise,}\end{cases}\end{align*} $$
$$ \begin{align*}M_2(V_k, V_\ell)=\begin{cases} 1 \quad&\text{if } k=0\;\text{and}\; \ell=2,\\1\quad&\text{if } k=1\;\text{and}\; \ell\in \{1,3\},\\ 1\quad&\text{if } k\ge 2\;\text{and}\; \ell\in\{k-2,k,k+2\},\\0\quad&\text{otherwise.}\end{cases}\end{align*} $$

The skeleton looks like

and this $2$ -graph is cofinal, satisfies the aperiodicity condition and every vertex connects to a loop with an entrance; in particular, $\mathcal {O}_{\rho _1,\rho _2}$ is simple and purely infinite.

Footnotes

Communicated by Lisa Orloff Clark

The author would like to thank the referee for useful suggestions.

References

Albandik, S. and Meyer, R., ‘Product systems over Ore monoids’, Doc. Math. 20 (2015), 13311402.Google Scholar
Bröcker, T. and tom Dieck, T., Representations of Compact Lie Groups , Graduate Texts in Mathematics, 98 (Springer, Berlin–Heidelberg, 1985).Google Scholar
Burgstaller, B., ‘Some multidimensional Cuntz algebras’, Aequationes Math. 76(1–2) (2008), 1932.CrossRefGoogle Scholar
Carlsen, T. M., Larsen, N., Sims, A. and Vittadello, S. T., ‘Co-universal algebras associated to product systems and gauge-invariant uniqueness theorems’, Proc. Lond. Math. Soc. (3) 103(4) (2011), 563600.CrossRefGoogle Scholar
Davidson, K. R. and Yang, D., ‘Periodicity in rank $2$ graph algebras’, Canad. J. Math. 61(6) 2009, 12391261.Google Scholar
Davidson, K. R. and Yang, D., ‘Representations of higher rank graph algebras’, New York J. Math. 15 (2009), 169198.Google Scholar
Deaconu, V., Huang, L. and Sims, A., ‘Group actions on product systems and K-theory’, to appear.Google Scholar
Doplicher, S. and Roberts, J. E., ‘Duals of compact Lie groups realized in the Cuntz algebras and their actions on C*-algebras’, J. Funct. Anal. 74 (1987), 96120.CrossRefGoogle Scholar
Doplicher, S. and Roberts, J. E., ‘A new duality theory for compact groups’, Invent. Math. 98 (1989), 157218.CrossRefGoogle Scholar
Dor-On, A. and Kakariadis, E. T. A., ‘Operator algebras for higher rank analysis and their application to factorial languages’, J. Anal. Math. 143(2) (2021), 555613.Google Scholar
Evans, D. G., ‘On the $K$ -theory of higher rank graph ${C}^{\ast }$ -algebras’, New York J. Math. 14 (2008), 131.Google Scholar
Fowler, N. J., ‘Discrete product systems of finite dimensional Hilbert spaces and generalized Cuntz algebras’, Preprint, 1999, https://arxiv.org/abs/math/9904116.Google Scholar
Fowler, N. J., ‘Discrete product systems of Hilbert bimodules’, Pacific J. Math. 204 (2002), 335375.CrossRefGoogle Scholar
Fowler, N. J. and Sims, A., ‘Product systems over right-angled Artin semigroups’, Trans. Amer. Math. Soc. 354 (2002), 14871509.CrossRefGoogle Scholar
Hao, G. and Ng, C.-K., ‘Crossed products of C*-correspondences by amenable group actions’, J. Math. Anal. Appl. 345(2) (2008), 702707.CrossRefGoogle Scholar
Kajiwara, T., Pinzari, C. and Watatani, Y., ‘Ideal structure and simplicity of the ${C}^{\ast }$ -algebras generated by Hilbert bimodules’, J. Funct. Anal. 159(2) (1998), 295322.CrossRefGoogle Scholar
Katsoulis, E., ‘Product systems of ${C}^{\ast }$ -correspondences and Takai duality’, Israel J. Math. 240(1) (2020), 223251.CrossRefGoogle Scholar
Kumjian, A. and Pask, D., ‘Higher rank graph ${C}^{\ast }$ -algebras’, New York J. Math. 6 (2000), 120.Google Scholar
Kumjian, A., Pask, D., Raeburn, I. and Renault, J., ‘Graphs, groupoids and Cuntz–Krieger algebras , J. Funct. Anal. 144(2) (1997), 505541.Google Scholar
Li, H. and Yang, D., ‘Boundary quotient ${C}^{\ast }$ -algebras of products of odometers’, Canad. J. Math. 71(1) (2019), 183212.CrossRefGoogle Scholar
Li, H. and Yang, D., ‘Self-similar $k$ -graph ${C}^{\ast }$ -algebras’, Int. Math. Res. Not. IMRN 15 (2021), 1127011305.CrossRefGoogle Scholar
Mann, M. H., Raeburn, I. and Sutherland, C. E., ‘Representations of finite groups and Cuntz–Krieger algebras’, Bull. Aust. Math. Soc. 46 (1992), 225243.CrossRefGoogle Scholar
Robertson, D. and Sims, A., ‘Simplicity of ${C}^{\ast }$ -algebras associated to row-finite locally convex higher-rank graphs’, Israel J. Math. 172 (2009), 171192.CrossRefGoogle Scholar
Sims, A., ‘Gauge-invariant ideals in ${C}^{\ast }$ -algebras of finitely aligned higher-rank graphs’, Canad. J. Math. 58(6) (2006), 12681290.Google Scholar
Sims, A., ‘Graphs and ${C}^{\ast }$ -algebras’, Austral. Math. Soc. Gaz. 39(1) (2012).Google Scholar
Sims, A. and Yeend, T., ‘ ${\mathrm{C}}^{\ast }$ -algebras associated to product systems of Hilbert bimodules’, J. Operator Theory 64(2) (2010), 349376.Google Scholar