1 Introduction

To give a unified framework like graph \(C^*\)-algebras for the Katsura’s [11] and Nekrashevyche’s algebras [17, 18], Exel and Pardo introduced self-similar graphs and their \(C^*\)-algebras in [7]. They then associated an inverse semigroup and groupoid model to this class of \(C^*\)-algebras and studied structural features by underlying self-similar graphs. Note that although only finite graphs are considered in [7], many of arguments and results may be easily generalized for countable row-finite graphs with no sources (see [8, 10] for example). Inspired from [7], Li and Yang in [15, 16] introduced self-similar action of a discrete countable group G on a row-finite k-graph \(\Lambda \). They then associated a universal \(C^*\)-algebra \({\mathcal {O}}_{G,\Lambda }\) to \((G,\Lambda )\) satisfying specific relations.

The algebraic analogues of Exel-Pardo \(C^*\)-algebras, denoted by \({\mathcal {O}}_{(G,E)}^{\textrm{alg}}\) in [6] and by \(L_R(G,E)\) in [9], were introduced and studied in [6, 9]. In particular, Hazrat et al. proved a \({\mathbb {Z}}\)-graded uniqueness theorem and gave a model of Steinberg algebras for \(L_R(G,E)\) [9]. The initial aims to write the present paper are to give a much easier proof for [9, Theorem B] (a groupoid model for \(L_R(G,E)\)) using the \({\mathbb {Z}}\)-graded uniqueness theorem and then to study the ideal structure. However, we do these here, among others, for a more general class of algebras associated to self-similar higher rank graphs \((G,\Lambda )\), which is introduced in Sect. 2.

This article is organized as follows. Let R be a unital commutative \(*\)-ring. In Sect. 2, we introduce a universal \(*\)-algebra \({\textrm{EP}}_R(G,\Lambda )\) of a self-similar k-graph \((G,\Lambda )\) satisfying specific properties, which is called the Exel-Pardo algebra of \((G,\Lambda )\). Our algebras are the higher rank generalization of those in [9, Theorem 1.6] and the algebraic analogue of \({\mathcal {O}}_{G,\Lambda }\) [15, 16]. Moreover, this class of algebras includes many important known algebras such as the algebraic Katsura algebras [9], Kumjian-Pask algebras [2], and the quotient boundary algebras \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\) of a Zappa-Sz\(\acute{\textrm{e}}\)p product \(\Lambda \bowtie G\) introduced in Sect. 3. In Sect. 3, we give a specific example of Exel-Pardo algebras using boundary quotient algebras of semigroups. Indeed, for a single-vertex self-similar k-graph \((G,\Lambda )\), the Zappa-Szép product \(\Lambda \bowtie G\) is a cancellative semigroup. We prove that the quotient boundary algebra \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\) (defined in Definition 3.1) is isomorphic to \({\textrm{EP}}_R(G,\Lambda )\). Section 4 is devoted to proving a graded uniqueness theorem for the Exel-Pardo algebras. Note that using the description in Proposition 2.7, there is a natural \({\mathbb {Z}}^k\)-grading on \({\textrm{EP}}_R(G,\Lambda )\). Then, in Theorem 4.2, a \({\mathbb {Z}}^k\)-graded uniqueness theorem is proved for \({\textrm{EP}}_R(G,\Lambda )\) which generalizes and modifies [9, Theorem A]. In particular, we will see in Sects. 5 and 6 that this modification makes it more applicable.

In Sects. 5 and 6, we assume that our self-similar k-graphs are pseudo-free (Definition 5.1). In Sect. 5, we prove that every Exel-Pardo algebra \({\textrm{EP}}_R(G,\Lambda )\) is isomorphic to the Steinberg algebra \(A_R({\mathcal {G}}_{G,\Lambda })\), where \({\mathcal {G}}_{G,\Lambda }\) is the groupoid introduced in [15]. We should note that the proof of this result is completely different from that of [9, Theorem B]. Indeed, the main difference between the proof of Theorem 5.5 and that of [9, Theorem B] is due to showing the injectivity of defined correspondence. In fact, in [9, Theorem B], the authors try to define a representation for \({\mathcal {S}}_{(G,E)}\) in \(\textrm{EP}_R(G,E)\) while we apply our graded uniqueness theorem, Theorem 4.2. This gives us an easier proof for Theorem 5.5, even in the 1-graph case. Finally, in Sect. 6, we investigate the ideal structure of \({\textrm{EP}}_R(G,\Lambda )\). Using the Steinberg algebras, we can define a conditional expectation \({\mathcal {E}}\) on \({\textrm{EP}}_R(G,\Lambda )\) and then characterize basic, \({\mathbb {Z}}^k\)-graded, diagonal-invariant ideals of \({\textrm{EP}}_R(G,\Lambda )\) by G-saturated G-hereditary subsets of \(\Lambda ^0\). These ideals are exactly basic, \(Q({\mathbb {N}}^k,G)\rtimes _{{\mathcal {T}}}{\mathbb {Z}}^k\)-graded ideals of \({\textrm{EP}}_R(G,\Lambda )\).

1.1 Notation and Terminology

Let \({\mathbb {N}}=\{0,1,2,\ldots \}\). For \(k\ge 1\), we regard \({\mathbb {N}}^k\) as an additive semigroup with the generators \(e_1,\ldots ,e_k\). We use \(\le \) for the partial order on \({\mathbb {N}}^k\) given by \(m\le n\) if and only if \(m_i\le n_i\) for \(1\le i\le k\). We also write \(m\vee n\) and \(m\wedge n\) for the coordinate-wise maximum and minimum, respectively.

A k-graph is a countable small category \(\Lambda =(\Lambda ^0,\Lambda ,r,s)\) equipped with a degree functor \(d:\Lambda \rightarrow {\mathbb {N}}^k\) satisfying the unique factorization property: for \(\mu \in \Lambda \) and \(m,n\in {\mathbb {N}}^k\) with \(d(\mu )=m+n\), then there exist unique \(\alpha ,\beta \in \Lambda \) such that \(d(\alpha )=m\), \(d(\beta )=n\), and \(\mu =\alpha \beta \). We usually denote \(\mu (0,m):=\alpha \) and \(\mu (m,d(\mu )):=\beta \). We refer to \(\Lambda ^0\) as the vertex set and define \(\Lambda ^n:=\{\mu \in \Lambda : d(\mu )=n\}\) for every \(n\in {\mathbb {N}}^k\). For \(A,B\subseteq \Lambda \), define \(AB=\{\mu \nu :\mu \in A,\nu \in B, ~\textrm{and}~ s(\mu )=r(\nu )\}\). Also, for \(\mu ,\nu \in \Lambda \), define \(\Lambda ^{min}(\mu ,\nu )=\{(\alpha ,\beta )\in \Lambda \times \Lambda : \mu \alpha =\nu \beta , d(\mu \alpha )=d(\mu )\vee d(\nu )\}\).

We say that \(\Lambda \) is row-finite if \(v\Lambda ^n\) is finite for all \(n\in {\mathbb {N}}^k\) and \(v\in \Lambda ^0\). A source in \(\Lambda \) is a vertex \(v\in \Lambda ^0\) such that \(v\Lambda ^{e_i}=\emptyset \) for some \(1\le i\le k\).

Standing assumption. Throughout the article, we work only with row-finite k-graphs without sources.

Let \(\Omega _k:=\{(m,n)\in {\mathbb {N}}^k\times {\mathbb {N}}^k:m\le n\}\). By defining \((m,n).(n,l):=(m,l)\), \(r(m,n):=(m,m)\), and \(s(m,n):=(n,n)\), then \(\Omega _k\) is a row-finite k-graph without sources. A graph homomorphism \(x:\Omega _k\rightarrow \Lambda \) is called an infinite path of \(\Lambda \) with the range \(r(x)=x(0,0)\), and we write \(\Lambda ^\infty \) for the set of all infinite paths of \(\Lambda \).

2 Exel-Pardo Algebras of Self-Similar k-Graphs

In this section, we associate a \(*\)-algebra to a self-similar k-graph as the algebraic analogue of [16, Definition 3.9]. Let us first review some definitions and notations.

Following [9], we consider \(*\)-algebras over \(*\)-rings. Let R be a unital commutative \(*\)-ring. Recall that a \(*\)-algebra over R is an algebra A equipped with an involution such that \((a^*)^*=a\), \((ab)^*=b^* a^*\), and \((ra+b)^*=r^* a^*+b^*\) for all \(a,b\in A\) and \(r\in R\). Then \(p\in A\) is called a projection if \(p^2=p=p^*\), and \(s\in A\) a partial isometry if \(s=ss^*s\).

Definition 2.1

([2, Definition 3.1]) Let \(\Lambda \) be a row-finite k-graph without sources. A Kumjian-Pask \(\Lambda \)-family is a collection \(\{s_{\mu }:\mu \in \Lambda \}\) of partial isometries in a \(*\)-algebra A such that

  1. (KP1)

    \(\{s_v:v\in \Lambda ^0\}\) is a family of pairwise orthogonal projections;

  2. (KP2)

    \(s_{\mu \nu }=s_\mu s_\nu \) for all \(\mu ,\nu \in \Lambda \) with \(s(\mu )=r(\nu )\);

  3. (KP3)

    \(s_\mu ^* s_\mu =s_{s(\mu )}\) for all \(\mu \in \Lambda \); and

  4. (KP4)

    \(s_v=\sum _{\mu \in v\Lambda ^n}s_\mu s_\mu ^*\) for all \(v\in \Lambda ^0\) and \(n\in {\mathbb {N}}^k\).

2.1 Self-Similar k-Graphs and Their Algebras

Let \(\Lambda \) be a row-finite k-graph without sources. An automorphism of \(\Lambda \) is a bijection \(\psi :\Lambda \rightarrow \Lambda \) such that \(\psi (\Lambda ^n)\subseteq \Lambda ^n\) for all \(n\in {\mathbb {N}}^k\) with the properties \(s\circ \psi =\psi \circ s\) and \(r\circ \psi =\psi \circ r\). We denote by \(\textrm{Aut}(\Lambda )\) the group of automorphisms on \(\Lambda \). Furthermore, if G is a countable discrete group, an action of G on \(\Lambda \) is a group homomorphism \(g\mapsto \psi _g\) from G into \(\textrm{Aut}(\Lambda )\).

Definition 2.2

([15]) Let \(\Lambda \) be a row-finite k-graph without sources and G a discrete group with identity \(e_G\). We say that a triple \((G,\Lambda ,\varphi )\) is a self-similar k-graph whenever the following properties hold:

  1. (1)

    G acts on \(\Lambda \) by a group homomorphism \(g\mapsto \psi _g\). We prefer to write \(g\cdot \mu \) for \(\psi _g(\mu )\) to ease the notation.

  2. (2)

    \(\varphi :G\times \Lambda \rightarrow G\) is a 1-cocycle for the action \(G\curvearrowright \Lambda \) such that for every \(g\in G\), \(\mu ,\nu \in \Lambda \) and \(v\in \Lambda ^0\) we have

    1. (a)

      \(\varphi (gh,\mu )=\varphi (g,h \cdot \mu )\varphi (h,\mu ) \) (the 1-cocycle property),

    2. (b)

      \(g\cdot (\mu \nu )=(g\cdot \mu )(\varphi (g,\mu ) \cdot \nu ) \) (the self-similar equation),

    3. (c)

      \(\varphi (g,\mu \nu )=\varphi (\varphi (g,\mu ),\nu )\), and

    4. (d)

      \(\varphi (g,v)=g\).

For convenience, we usually write \((G,\Lambda )\) instead of \((G,\Lambda ,\varphi )\).

Remark 2.3

In [15], the authors used the notation \(g|_\mu \) for \(\varphi (g,\mu )\). However, we would prefer to follow [7,8,9] for writing \(\varphi (g,\mu )\).

Remark 2.4

If in equation (2)(a) of Definition 2.2, we set \(g=h=e_G\), then we get \(\varphi (e_G,\mu )=e_G\) for every \(\mu \in \Lambda \). Moreover, [15, Lemma 3.5(ii)] shows that \(\varphi (g,\mu )\cdot v=g\cdot v\) for all \(g\in G\), \(v\in \Lambda ^0\), and \(\mu \in \Lambda \).

Now we generalize the definition of Exel-Pardo \(*\)-algebras [9] to the k-graph case.

Definition 2.5

Let \((G,\Lambda )\) be a self-similar k-graph. An Exel-Pardo \((G,\Lambda )\)-family (or briefly \((G,\Lambda )\)-family) is a set

$$\begin{aligned} \{s_\mu :\mu \in \Lambda \}\cup \{u_{v,g}:v\in \Lambda ^0,g\in G\} \end{aligned}$$

in a \(*\)-algebra satisfying

  1. (1)

    \(\{s_\mu :\mu \in \Lambda \}\) is a Kumjian-Pask \(\Lambda \)-family,

  2. (2)

    \(u_{v,e_G}=s_v\) for all \(v\in \Lambda ^0\),

  3. (3)

    \(u_{v,g}^*=u_{g^{-1} \cdot v,g^{-1}}\) for all \(v\in \Lambda ^0\) and \(g\in G\),

  4. (4)

    \(u_{v,g}s_\mu =\delta _{v,g\cdot r(\mu )}s_{g\cdot \mu }u_{g\cdot s(\mu ),\varphi (g,\mu )}\) for all \(v\in \Lambda ^0\), \(\mu \in \Lambda \), and \(g\in G\),

  5. (5)

    \(u_{v,g}u_{w,h}=\delta _{v,g\cdot w}u_{v,gh}\) for all \(v,w\in \Lambda ^0\) and \(g,h\in G\).

Then the Exel-Pardo algebra \({\textrm{EP}}_R(G,\Lambda )\) is the universal \(*\)-algebra over R generated by a \((G,\Lambda )\)-family \(\{s_\mu ,u_{v,g}\}\).

Recall that the universality of \({\textrm{EP}}_R(G,\Lambda )\) means that for every \((G,\Lambda )\)-family \(\{S_\mu ,U_{v,g}\}\) in a \(*\)-algebra A, there exists a \(*\)-homomorphism \(\phi :{\textrm{EP}}_R(G,\Lambda )\rightarrow A\) such that \(\phi (s_\mu )=S_\mu \) and \(\phi (u_{v,g})=U_{v,g}\) for all \(v\in \Lambda ^0\), \(\mu \in \Lambda \), and \(g\in G\). (See Sect. 2.2 for the construction of \({\textrm{EP}}_R(G,\Lambda )\).) Throughout the paper, we will denote by \(\{s_\mu ,u_{v,g}\}\) the \((G,\Lambda )\)-family generating \({\textrm{EP}}_R(G,\Lambda )\).

2.2 The Construction of \({\textrm{EP}}_R(G,\Lambda )\)

Let \((G,\Lambda )\) be a self-similar k-graph as in Definition 2.2. The following is a standard construction of a universal algebra \({\textrm{EP}}_R(G,\Lambda )\) subject to desired relations. Consider the set of formal symbols

$$\begin{aligned} S=\left\{ S_\mu ,S_\mu ^*,U_{v,g},U_{v,g}^*:\mu \in \Lambda , v\in \Lambda ^0, g\in G\right\} . \end{aligned}$$

Let \(X=w(S)\) be the collection of finite words in S. We equip the free R-module

$$\begin{aligned} {\mathbb {F}}_R(X):=\bigg \{\sum _{i=1}^l r_ix_i: l\ge 1, r_i\in R, x_i\in X\bigg \} \end{aligned}$$

with the multiplication

$$\begin{aligned} \bigg (\sum _{i=1}^l r_i x_i\bigg )\bigg (\sum _{j=1}^{l'} s_j y_j\bigg ):=\sum _{i,j} r_is_j x_iy_j, \end{aligned}$$

and the involution

$$\begin{aligned} \bigg (\sum r_i x_i\bigg )^*:=\bigg (\sum r_i^* x_i^*\bigg ) \end{aligned}$$

where \(x^*=s_l^*\ldots s_1^*\) for each \(x=s_1\ldots s_l\). Then \({\mathbb {F}}_R(X)\) is a \(*\)-algebra over R. If I is the (two-sided and self-adjoint) ideal of \({\mathbb {F}}_R(X)\) containing the roots of relations (1)-(5) in Definition 2.5, then the quotient \({\mathbb {F}}_R(X)/I\) is the Exel-Pardo algebra \({\textrm{EP}}_R(G,\Lambda )\) with the desired universal property. Let us define \(s_\mu :=S_\mu +I\) and \(u_{v,g}:=U_{v,g}+I\) for every \(\mu \in \Lambda \), \(v\in \Lambda ^0\), and \(g\in G\). In case \((G,\Lambda )\) is pseudo-free (Definition 5.1), Theorem 4.2 insures that all generators \(\{s_\mu ,u_{v,g}\}\) of \({\textrm{EP}}_R(G,\Lambda )\) are nonzero.

Proposition 2.7 describes the elements of \({\textrm{EP}}_R(G,\Lambda )\). First, see a simple lemma.

Lemma 2.6

Let \((G,\Lambda )\) be a self-similar graph (as in Definition 2.2) and \(\{S,U\}\) a \((G,\Lambda )\)-family. If \(S_\mu U_{v,g} S_\nu ^*\ne 0\) where \(\mu ,\nu \in \Lambda \), \(v\in \Lambda ^0\) and \(g\in G\), then \(s(\mu )=v=g\cdot s(\nu )\).

Proof

If \(a=S_\mu U_{v,g} S_\nu ^*\) is nonzero, then by Definition 2.5 we can write

$$\begin{aligned} S_\mu U_{v,g}S_\nu ^*&= S_\mu (S_{s(\mu )}U_{v,g} ) S_\nu ^*\\&= S_\mu (U_{s(\mu ),e_G} U_{v,g}) S_\nu ^*\\&=S_\mu (\delta _{s(\mu ),e_G\cdot v} U_{s(\mu ),g}) S_\nu ^*. \end{aligned}$$

Now, the hypothesis \(a\ne 0\) forces \(s(\mu )=v\). On the other hand, a similar computation gives

$$\begin{aligned} a&= S_\mu (U_{v,g} S_{s(\nu )}) S_\nu ^*\\&= S_\mu ( U_{v,g} U_{s(\nu ),e_G}) S_\nu ^*\\&=S_\mu (\delta _{v, g\cdot s(\nu )} U_{v,g}) S_\nu ^*, \end{aligned}$$

and thus \(v=g\cdot s(\nu )\). \(\square \)

Proposition 2.7

Let \((G,\Lambda )\) be a self-similar graph. Then

$$\begin{aligned} {\textrm{EP}}_R(G,\Lambda )=\textrm{span}_R\{s_\mu u_{s(\mu ),g} s_\nu ^*: g\in G,~ \mu ,\nu \in \Lambda ,~ \textrm{and}~ s(\mu )=g\cdot s(\nu )\}. \end{aligned}$$
(2.1)

Proof

Define \(M:=\textrm{span}_R\{s_\mu u_{s(\mu ),g} s_\nu ^*: g\in G,~ \mu ,\nu \in \Lambda \}\). For every \(g,h\in G\) and \(\mu ,\nu ,\alpha ,\beta \in \Lambda \) with \(\alpha =\nu \alpha '\) for some \(\alpha '\in \Lambda \), the relations of Definition 2.5 imply that

$$\begin{aligned} \bigg (s_\mu u_{s(\mu ),g} s_\nu ^* \bigg ) \bigg (s_\alpha u_{s(\alpha ),h} s_\beta ^* \bigg )&= s_\mu u_{s(\mu ),g}(s_\nu ^* s_\alpha ) u_{s(\alpha ),h} s_\beta ^*\\&= s_\mu u_{s(\mu ),g}(s_{\alpha '}) u_{s(\alpha ),h} s_\beta ^* \\&= s_\mu \bigg ( \delta _{s(\mu ),g\cdot r(\alpha ')} s_{g\cdot \alpha '} u_{g\cdot s(\alpha '), \varphi (g,\alpha ')} u_{s(\alpha ),h}\bigg ) s_\beta ^* \\&= \delta _{s(\mu ),g\cdot s(\nu )} s_{\mu (g\cdot \alpha ')}\\&\quad \bigg (\delta _{g\cdot s(\alpha '),\varphi (g,\alpha ') \cdot s(\alpha )} u_{g\cdot s(\alpha '), \varphi (g,\alpha ') h} \bigg ) s_\beta ^*\\&\qquad (\textrm{as}~~ r(\alpha ')=s(\nu )). \end{aligned}$$

In the case \(\nu =\alpha \nu '\) for some \(\nu '\in \Lambda \), the above multiplication may be computed similarly, and otherwise is zero. Hence, M is closed under multiplication. Also, we have

$$\begin{aligned} \bigg (s_\mu u_{s(\mu ),g}s_\nu ^* \bigg )^*= s_\nu u_{g^{-1} \cdot s(\mu ), g^{-1}} s_\mu ^*, \end{aligned}$$

so \(M^*\subseteq M\). Since

$$\begin{aligned} s_\mu =s_\mu u_{s(\mu ),e_G} s_{s(\mu )}^* \textrm{ and } u_{v,g}=s_v u_{v,g} s_{g\cdot v}^* \end{aligned}$$

for all \(g\in G\), \(v\in \Lambda ^0\), and \(\mu \in \Lambda \), it follows that M is a \(*\)-subalgebra of \({\textrm{EP}}_R(G,\Lambda )\) containing the generators of \({\textrm{EP}}_R(G,\Lambda )\). In light of Lemma 2.6, this concludes the identification (2.1). \(\square \)

2.3 The Unital Case

In case \(\Lambda \) is a k-graph with finite \(\Lambda ^0\), we may give a better description for Definition 2.5. Note that this case covers all unital Exel-Pardo algebras \({\textrm{EP}}_R(G,\Lambda )\).

Lemma 2.8

Let \((G,\Lambda )\) be a self-similar k-graph and let \(s_v\) be nonzero in \({\textrm{EP}}_R(G,\Lambda )\) for every \(v\in \Lambda ^0\). Then \({\textrm{EP}}_R(G,\Lambda )\) is a unital algebra if and only if the vertex set \(\Lambda ^0\) is finite.

Proof

If \(\Lambda ^0=\{v_1,\ldots , v_l\}\) is finite, then using identification (2.1), \(P=\sum _{i=1}^l s_{v_i}\) is the unit of \({\textrm{EP}}_R(G,\Lambda )\). Conversely, if \(\Lambda ^0\) is infinite, then the set \(\{s_v:v\in \Lambda ^0\}\subseteq {\textrm{EP}}_R(G,\Lambda )\) contains infinitely many mutually orthogonal projections. Now again by (2.1), there is no element of \({\textrm{EP}}_R(G,\Lambda )\) which acts as an identity on each element of \(\{s_v:v\in \Lambda ^0\}\). \(\square \)

Note that if \(\{s,u\}\) is a \((G,\Lambda )\)-family in a \(*\)-algebra A, then for each \(g\in G\) we may define \(u_g:=\sum _{v\in \Lambda ^0}u_{v,g}\) as an element of the multiplier algebra \({\mathcal {M}}(A)\) with the property \(s_v u_g=u_{v,g}\) for all \(v\in \Lambda ^0\) (relations (2) and (5) of Definition 2.5 yield \(s_vu_{w,g}=\delta _{v,w} u_{v,g}\)). (See [1] for the definition of multiplier algebras.) Thus relations (3) and (5) of Definition 2.5 imply that \(u:G\rightarrow {\mathcal {M}}(A)\), defined by \(g \mapsto u_g\), is a unitary \(*\)-representation of G on \({\mathcal {M}}(A)\). In particular, in case \(\Lambda ^0\) is finite, \(u_g\)’s lie all in A, and we may describe Definition 2.5 as the following:

Proposition 2.9

Let \((G,\Lambda )\) be a self-similar k-graph. Suppose also that \(\Lambda ^0\) is finite. Then \({\textrm{EP}}_R(G,\Lambda )\) is the universal \(*\)-algebra generated by families \(\{s_\mu :\mu \in \Lambda \}\) of partial isometries and \(\{u_g:g\in G\}\) of unitaries satisfying

  1. (1)

    \(\{s_\mu :\mu \in \Lambda \}\) is a Kumjian-Pask \(\Lambda \)-family;

  2. (2)

    \(u:G\rightarrow {\textrm{EP}}_R(G,\Lambda )\), by \(g\mapsto u_g\), is a unitary \(*\)-representation of G on \({\textrm{EP}}_R(G,\Lambda )\), in the sense that

    1. (a)

      \(u_g u_h=u_{gh}\) for all \(g,h\in G\), and

    2. (b)

      \(u^*_g= u^{-1}_g= u_{g^{-1}}\) for all \(g\in G\);

  3. (3)

    \(u_g s_\mu =s_{g\cdot \mu } u_{\varphi (g,\mu )}\) for all \(g\in G\) and \(\mu \in \Lambda \).

3 An example: The Zappa-Szép Product \(\Lambda \bowtie G\) and its \(*\)-Algebra

Let \((G,\Lambda )\) be a self-similar k-graph such that \(|\Lambda ^0|=1\). The \(C^*\)-algebra and quotient boundary \(C^*\)-algebra associated to the Zappa-Szép product \(\Lambda \bowtie G\) as a semigroup were studied in [4, 14]. In this section, we first define \({\mathcal {Q}}_R^{\textrm{alg}}(S)\) as the algebraic analogue of the quotient boundary \(C^*\)-algebra \({\mathcal {Q}}(S)\) of a cancellative semigroup S. Then we show that \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\) is isomorphic to the Exel-Pardo algebra \({\textrm{EP}}_R(G,\Lambda )\).

Let us recall some terminology from [4, 13]. Let S be a left-cancellative semigroup with an identity. Given \(X\subseteq S\) and \(s\in S\), define \(sX:=\{sx:x\in X\}\) and \(s^{-1}X:=\{r\in S:sr\in X\}\). Also, the set of constructible right ideals in S is defined as

$$\begin{aligned} {\mathcal {J}}(S):=\{s^{-1}_1r_1\ldots s_l^{-1} r_l S: l\ge 1, ~ s_i,r_i\in S\}\cup \{\emptyset \}. \end{aligned}$$

Then, a foundation set in \({\mathcal {J}}(S)\) is a finite subset \(F\subseteq {\mathcal {J}}(S)\) such that for each \(Y\in {\mathcal {J}}(S)\), there exists \(X\in F\) with \(X\cap Y\ne \emptyset \).

The following is the algebraic analogue of [13, Definition 2.2].

Definition 3.1

Let S be a left-cancellative semigroup and R be a unital commutative \(*\)-ring. The boundary quotient \(*\)-algebra of S is the universal unital \(*\)-algebra \({\mathcal {Q}}_R^{\textrm{alg}}(S)\) over R generated by a set of isometries \(\{t_s:s\in S\}\) and a set of projections \(\{q_X:X\in {\mathcal {J}}(S)\}\) satisfying

  1. (1)

    \(t_s t_r=t_{sr}\),

  2. (2)

    \(t_s q_X t_s^*=q_{sX}\),

  3. (3)

    \(q_S=1\) and \(q_\emptyset =0\),

  4. (4)

    \(q_X q_Y=q_{X\cap Y}\), and moreover

  5. (5)

    \(\prod _{X\in F} (1-q_X)=0\)

for all \(s,r\in S\), \(X,Y\in {\mathcal {J}}(S)\), and foundation sets \(F\subseteq {\mathcal {J}}(S)\).

Let \(\Lambda \) be a k-graph such that \(\Lambda ^0=\{v\}\). Then \(\mu \nu \) is composable for all \(\mu ,\nu \in \Lambda \), and hence \(\Lambda \) may be considered as a semigroup with the identity v. Also, the unique factorization property implies that \(\Lambda \) is cancellative.

Definition 3.2

([4, Definition 3.1]) Let \((G,\Lambda )\) is a single-vertex self-similar k-graph. If we consider \(\Lambda \) as a semigroup, then the Zappa-Szép product \(\Lambda \bowtie G\) is the semigroup \(\Lambda \times G\) with the multiplication

$$\begin{aligned} (\mu ,g)(\nu ,h):=(\mu (g\cdot \nu ), \varphi (g,\nu ) \cdot h) \qquad (\mu ,\nu \in \Lambda ~ \textrm{and}~ g,h\in G). \end{aligned}$$

Remark 3.3

If \(\Lambda \) is a single-vertex k-graph, then [14, Lemma 3.2 (iv)] follows that

$$\begin{aligned} {\mathcal {J}}(\Lambda )=\bigg \{\bigcup _{i=1}^l\mu _i \Lambda : l\ge 1, \mu _i\in \Lambda , d(\mu _1)=\cdots =d(\mu _l)\bigg \}. \end{aligned}$$

In order to prove Theorem 3.6, the following lemmas are useful.

Lemma 3.4

Let \((G,\Lambda )\) be a self-similar k-graph with \(\Lambda ^0=\{v\}\). Suppose that for each \(\mu \in \Lambda \), the map \(g\mapsto \varphi (g,\mu )\) is surjective. Then

  1. (1)

    \({\mathcal {J}}(\Lambda \bowtie G)={\mathcal {J}}(\Lambda )\times \{G\}\), where \(\emptyset \times G:=\emptyset \).

  2. (2)

    A finite subset \(F\subseteq {\mathcal {J}}(\Lambda )\) is a foundation set if and only if \(F'=F\times \{G\}\) is a foundation set in \({\mathcal {J}}(\Lambda \bowtie G)\).

Proof

Statement (1) is just [14, Lemma 2.13]. For (2), suppose that \(F\subseteq {\mathcal {J}}(\Lambda )\) is a foundation set, and let \(Y\times G \in {\mathcal {J}}(\Lambda \bowtie G)\). Then there exists \(X\in F\) such that \(X\cap Y\ne \emptyset \). Thus \((X\times G)\cap (Y\times G)\ne \emptyset \), from which we conclude that \(F\times \{G\}\) is a foundation set in \({\mathcal {J}}(\Lambda \bowtie G)\). The converse may be shown analogously. \(\square \)

In the following, for \(\mu \in \Lambda \) and \(E\subseteq \Lambda \) we define

$$\begin{aligned} \textrm{Ext}(\mu ;E):=\{\alpha : (\alpha ,\beta )\in \Lambda ^{\textrm{min}}(\mu ,\nu ) \mathrm {~ for ~ some ~} \nu \in E\}. \end{aligned}$$

Lemma 3.5

Let \((G,\Lambda )\) be a self-similar k-graph with \(\Lambda ^0=\{v\}\). For every \(X=\cup _{i=1}^l\mu _i \Lambda \) and \(Y=\cup _{j=1}^{l'} \nu _j \Lambda \) in \({\mathcal {J}}(\Lambda )\), we have

$$\begin{aligned} X\cap Y=\cup \{\mu _i\alpha \Lambda : 1\le i\le l, \alpha \in \textrm{Ext}(\mu _i;\{\nu _i,\ldots , \nu _{l'}\})\}. \end{aligned}$$

Proof

For any \(\lambda \in X\cap Y\), there are \(\alpha ,\beta \in \Lambda \), \(1\le i\le l\), and \(1\le j \le l'\) such that \(\lambda =\mu _i\alpha =\nu _j\beta \). Define

$$\begin{aligned} \alpha ':=\alpha (0,d(\mu _i)\vee d(\nu _j)-d(\mu _i)) ~~~ \textrm{and} ~~~ \beta ':=\beta (0,d(\mu _i)\vee d(\nu _j)-d(\nu _j)). \end{aligned}$$

Then the factorization property implies that \(\lambda =\mu _i\alpha '\lambda '=\nu _j\beta '\lambda '\) where \(d(\mu _i\alpha ')=d(\nu _j \beta ')=d(\mu _i)\vee d(\nu _j)\) and \(\lambda '=\lambda (d(\mu _i)\vee d(\nu _j),d(\lambda ))\). It follows that \(\lambda \in \mu _i\alpha '\Lambda \) with \(\alpha '\in \textrm{Ext}(\mu _i;\{\nu _j\})\) as desired. The reverse containment is trivial. \(\square \)

The following result is inspired by [14, Theorem 3.3].

Theorem 3.6

Let \((G,\Lambda )\) be a self-similar k-graph with \(\Lambda ^0=\{v\}\) and let \(\{s_\mu ,u_g\}\) be the \((G,\Lambda )\)-family generating \({\textrm{EP}}_R(G,\Lambda )\) as in Proposition 2.9. Suppose that for every \(\mu \in \Lambda \) the map \(g\mapsto \varphi (g,\mu )\) is surjective. If the family \(\{t_{(\mu ,g)}, q_X: (\mu ,g)\in \Lambda \bowtie G, X\in {\mathcal {J}}(\Lambda \bowtie G)\}\) generates \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\), then there exists an R-algebra \(*\)-isomorphism \(\pi :{\textrm{EP}}_R(G,\Lambda )\rightarrow {\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\) such that \(\pi (s_\mu )=t_{(\mu ,e_G)}\) and \(\pi (u_g)=q_{(v,g)}\) for all \(\mu \in \Lambda \) and \(g\in G\).

Proof

For every \(\mu \in \Lambda \) and \(g\in G\), define

$$\begin{aligned} S_\mu :=t_{(\mu ,e_G)} ~~ \textrm{and} ~~ U_g:=t_{(v,g)}. \end{aligned}$$

We will show that \(\{S, U\}\) is a \((G,\Lambda )\)-family in \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\), which is described in Proposition 2.9. First, for each \(g\in G\) we have

$$\begin{aligned} U_g U_g^*=t_{(v,g)} ~ q_{\Lambda \bowtie G} ~ t_{(v,g)}^* = q_{(v,g)\Lambda \bowtie G}= q_{\Lambda \bowtie G}=1_{{\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)}. \end{aligned}$$

So, \(g\mapsto U_g\) is a unitary \(*\)-representation of G into \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\). Moreover, (KP1)-(KP3) can be easily checked, so we verify (KP4) for \(\{s_\mu :\mu \in \Lambda \}\). Fix some \(n\in {\mathbb {N}}^k\). Then \(\{\mu \Lambda :\mu \in \Lambda ^n\}\) is a foundation set in \({\mathcal {J}}(\Lambda )\), and thus so is \(F=\{\mu \Lambda \times G:\mu \in \Lambda ^n\}\) in \({\mathcal {J}}(\Lambda \bowtie G)\) by Lemma 3.4. Hence we have

$$\begin{aligned} 1-\sum _{\mu \in \Lambda ^n} S_\mu S_\mu ^*&=\prod _{\mu \in \Lambda ^n}(1-S_\mu S_\mu ^*) \qquad (\mathrm {because ~} S_\mu S_\mu ^*\mathrm {s ~ are ~ pairwise ~ orthogonal})\\&=\prod _{\mu \in \Lambda ^n}(1-t_{(\mu ,e_G)} ~ q_{\Lambda \bowtie G} ~ t_{(\mu ,e_G)}^*)\\&=\prod _{\mu \in \Lambda ^n}(1-q_{\mu \Lambda \times G}) \qquad (\mathrm {by ~ eq.~ (2) ~ of ~ Definition ~ 3.1}) \\&=\prod _{X\in F} (1-q_X)=0 \qquad (\mathrm {by ~ eq.~ (5) ~ of ~ Definition ~ 3.1}). \end{aligned}$$

Because \(S_v=U_{e_G}=1\), (KP4) is verified, and therefore \(\{S_\mu :\mu \in \Lambda \}\) is a Kumjian-Pask \(\Lambda \)-family. Since for each \(\mu \in \Lambda \) and \(g\in G\),

$$\begin{aligned} U_g S_\mu =t_{(v,g)}t_{(\mu ,e_G)} = t_{(v,g)(\mu ,e_G)}=t_{(g\cdot \mu ,\varphi (g,\mu ))}= S_{g\cdot \mu } U_{\varphi (g,\mu )}, \end{aligned}$$

and so we have shown that \(\{S,U\}\) is a \((G,\Lambda )\)-family in \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\). Now the universality implies that the desired \(*\)-homomorphism \(\pi :{\textrm{EP}}_R(G,\Lambda )\rightarrow {\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\) exists.

Now we prove that \(\pi \) is an isomorphism. In order to do this, it suffices to find a homomorphism \(\rho :{\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\rightarrow {\textrm{EP}}_R(G,\Lambda )\) such that \(\rho \circ \pi =\textrm{id}_{{\textrm{EP}}_R(G,\Lambda )}\) and \(\pi \circ \rho =\textrm{id}_{{\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)}\). For any \((\mu ,g)\in \Lambda \bowtie G\) and \(X=(\cup _{i=1}^l \mu _i\Lambda ) \times G\in {\mathcal {J}}(\Lambda \bowtie G)\), we define

$$\begin{aligned} T_{(\mu ,g)}:=s_\mu u_g \textrm{and} Q_X:=\sum _{i=1}^l s_{\mu _i}s_{\mu _i}^*. \end{aligned}$$

We will show that the family \(\{T,Q\}\) satisfies the properties of Definitions 3.1. Relations (1)-(3) easily hold by the \((G,\Lambda )\)-relations for \(\{s,u\}\). Also, for every \(X=(\cup _{i=1}^l \mu _i\Lambda ) \times G\) and \(Y=(\cup _{j=1}^{l'}\nu _j\Lambda ) \times G\) in \({\mathcal {J}}(\Lambda \bowtie G)\), we have

$$\begin{aligned} Q_X Q_Y&=\bigg (\sum _{i=1}^l s_{\mu _i}s_{\mu _i}^* \bigg )\bigg (\sum _{j=1}^{l'} s_{\nu _j}s_{\nu _j}^* \bigg )\\&=\sum _{i,j} s_{\mu _i}(s_{\mu _i}^* s_{\nu _j})s_{\nu _j}^*\\&=\sum _{i,j} s_{\mu _i}\bigg (\sum _{(\alpha ,\beta )\in \Lambda ^{\textrm{min}}(\mu _i,\nu _j)}s_\alpha s_\beta ^* \bigg ) s_{\nu _j}^* \qquad (\mathrm {by ~ [1, Lemma~ 3.3]})\\&=\sum _{i,j} \sum _{\begin{array}{c} \mu _i\alpha =\nu _j \beta \\ d(\mu _i\alpha )=d(\mu _i)\vee d(\nu _j) \end{array}} s_{\mu _i\alpha } s_{\nu _j \beta }^*\\&=Q_{X\cap Y} \qquad (\mathrm {by ~ Lemma} ~ 3.5). \end{aligned}$$

For eq. (5) of Definition 3.1, let \(F=\{X_i\times G:=\cup _{j=1}^{t_i}\mu _{ij}\Lambda \times G: 1\le i\le l\}\) be a foundation set in \({\mathcal {J}}(\Lambda \bowtie G)\). Then \(F'=\{X_i=\cup _{j=1}^{t_i}\mu _{ij}\Lambda \}_{i=1}^l\) is a foundation set in \({\mathcal {J}}(\Lambda )\) by Lemma 3.4(2). Defining \(n:=\bigvee _{i,j}d(\mu _{ij})\), we claim that the set \(M=\{\mu _{ij}\alpha : \alpha \in \Lambda ^{n-d(\mu _{ij})}, 1\le i\le l, 1\le j\le t_i\}\) coincides with \(\Lambda ^n\). Indeed, if on the contrary there exists some \(\lambda \in \Lambda ^n \setminus M\), then \(\Lambda ^{\textrm{min}}(\lambda ,\mu _{ij})=\emptyset \), and hence \(\lambda \Lambda \cap \mu _{ij}\Lambda =\emptyset \) for all i and j. This yields that \(\lambda \Lambda \cap X_i=\emptyset \) for every \(X_i\in F'\), contradicting that \(F'\) is a foundation set in \({\mathcal {J}}(\Lambda )\).

Now one may compute

$$\begin{aligned} \prod _{X_i\times G\in F} (1-Q_{X_i\times G})&= \prod _{i=1}^l (1-\sum _{j=1}^{t_i} s_{\mu _{ij}} s_{\mu _{ij}}^*)\\&= \prod _{i=1}^l \bigg (1- \sum _{j=1}^{t_i} s_{\mu _{ij}} (\sum _{\alpha \in \Lambda ^{n-d(\mu _{ij})}} s_\alpha s_\alpha ^*) s_{\mu _{ij}}^* \bigg )\\&=\prod _{i=1}^l \bigg (1-\sum _{j=1}^{t_i} \sum _{\alpha \in \Lambda ^{n-d(\mu _{ij})}} s_{\mu _{ij}\alpha } s_{\mu _{ij}\alpha }^* \bigg ) \qquad (\star ). \end{aligned}$$

Observe that the projections \(s_{\mu _{ij}\alpha }s_{\mu _{ij}\alpha }^*\) are pairwise orthogonal because \(d(\mu _{ij}\alpha )=n\) for all ij (see [2, Remark 3.2(c)]). Hence, using the above claim, expression \((\star )\) equals to

$$\begin{aligned} (\star )=1-\sum _{\lambda \in \Lambda ^n}s_\lambda s_\lambda ^*=0 (\mathrm {by ~ (KP4)}). \end{aligned}$$

Therefore, the family \(\{T,Q\}\) satisfies the relations of Definition 3.1, and by the universality there exists an algebra \(*\)-homomorphism \(\rho :{\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\rightarrow {\textrm{EP}}_R(G,\Lambda )\) such that \(\rho (t_{(\mu ,g)})=T_{(\mu ,g)}\) and \(\rho (q_X)=Q_X\) for \((\mu ,g)\in \Lambda \bowtie G\) and \(X\in {\mathcal {J}}(\Lambda \bowtie G)\). It is clear that \(\rho \circ \pi =\textrm{id}_{{\textrm{EP}}_R(G,\Lambda )}\) and \(\pi \circ \rho =\textrm{id}_{{\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)}\) because they fix the generators of \({\textrm{EP}}_R(G,\Lambda )\) and \({\mathcal {Q}}_R^{\textrm{alg}}(\Lambda \bowtie G)\), respectively. Consequently, \(\pi \) is an isomorphism, completing the proof. \(\square \)

4 A Graded Uniqueness Theorem

In this section, we prove a graded uniqueness theorem for \({\textrm{EP}}_R(G,\Lambda )\) which generalizes and modifies [9, Theorem A] for self-similar k-graphs. This modification, in particular, helps us to prove Theorems 5.5 and 6.8.

Let us first recall some definitions. Let \(\Gamma \) be a group and A be an algebra over a ring R. A is called \(\Gamma \)-graded (or briefly, graded whenever the group is clear) if there is a family of R-submodules \(\{A_\gamma :\gamma \in \Gamma \}\) of A such that \(A=\bigoplus _{\gamma \in \Gamma } A_\gamma \) and \(A_\gamma A_{\gamma '}\subseteq A_{\gamma \gamma '}\) for all \(\gamma ,\gamma '\in \Gamma \). Then each set \(A_\gamma \) is called a \(\gamma \)-homogeneous component of A. In this case, we say an ideal I of A is \(\Gamma \)-graded if \(I=\bigoplus _{\gamma \in \Gamma }(I\cap A_\gamma )\). Note that an ideal I of A is \(\Gamma \)-graded if and only if it is generated by a subset of \(\bigcup _{\gamma \in \Gamma } A_\gamma \), the homogeneous elements of A.

Furthermore, if A and B are two \(\Gamma \)-graded algebras over R, a homomorphism \(\phi :A\rightarrow B\) is said to be a graded homomorphism if \(\phi (A_\gamma )\subseteq B_\gamma \) for all \(\gamma \in \Gamma \). Hence the kernel of a graded homomorphism is always a graded ideal. Also, if I is a graded ideal of A, then there is a natural \(\Gamma \)-grading \((A_\gamma +I)_{\gamma \in \Gamma }\) on the quotient algebra A/I, and thus the quotient map \(A\rightarrow A/I\) is a graded homomorphism.

Lemma 4.1

Let \((G,\Lambda )\) be a self-similar k-graph. If for every \(n\in {\mathbb {Z}}^k\), we define

$$\begin{aligned} {\textrm{EP}}_R(G,\Lambda )_n:=\textrm{span}_R\bigg \{s_\mu u_{s(\mu ),g} s_\nu ^*: g\in G,~ \mu ,\nu \in \Lambda ,~ \textrm{and}~ d(\mu )-d(\nu )=n\bigg \}, \end{aligned}$$

then \(({\textrm{EP}}_R(G,\Lambda )_n)_{n\in {\mathbb {Z}}^k}\) is a \({\mathbb {Z}}^k\)-grading on \({\textrm{EP}}_R(G,\Lambda )\).

Proof

Consider the free \(*\)-algebra \({\mathbb {F}}_R(X)\) and its ideal I as in Sect. 2.2. If we define

$$\begin{aligned} \theta (S_\mu ):=d(\mu ),~ \theta (S_\mu ^*):=-d(\mu ), ~\textrm{and} ~ \theta (U_{v,g}):=0 \end{aligned}$$

for all \(g\in G\), \(v\in \Lambda ^0\) and \(\mu \in \Lambda \), then \(\theta \) induces a \({\mathbb {Z}}^k\)-grading on \({\mathbb {F}}_R(X)\). Also, since the generators of I are all homogenous, I is a graded ideal. Therefore, \({\textrm{EP}}_R(G,\Lambda )\cong {\mathbb {F}}_R(X)/I\) is a \({\mathbb {Z}}^k\)-graded algebra, and Proposition 2.7 concludes the result. \(\square \)

Theorem 4.2

(Graded Uniqueness) Let \((G,\Lambda )\) be a self-similar k-graph. Let \(\phi :{\textrm{EP}}_R(G,\Lambda )\rightarrow B\) be a \({\mathbb {Z}}^k\)-graded R-algebra \(*\)-homomorphism into a \({\mathbb {Z}}^k\)-graded \(*\)-algebra B. If \(\phi (a)\ne 0\) for every nonzero element of the form \(a=\sum _{i=1}^l r_i u_{v,g_i}\) with \(v\in \Lambda ^0\) and \(g_i^{-1} \cdot v =g_j^{-1} \cdot v \) for \(1\le i,j\le l\), then \(\phi \) is injective.

Proof

For convenience, we write \(A={\textrm{EP}}_R(G,\Lambda )\). Since \(A=\bigoplus _{n\in {\mathbb {Z}}^k}A_n\) and \(\phi \) preserves the grading, it suffices to show that \(\phi \) is injective on each \(A_n\). So, fix some \(b\in A_n\), and assume \(\phi (b)=0\). By equation (2.1), we can write

$$\begin{aligned} b=\sum _{i=1}^lr_i s_{\mu _i}u_{w_i,g_i}s_{\nu _i}^* \end{aligned}$$
(4.1)

where \(w_i=s(\mu _i)=g_i \cdot s(\nu _i)\) and \(d(\mu _i)-d(\nu _i)=n\) for \(1\le i\le l\). Define \(n'=\vee _{1\le i\le l}d(\mu _i)\). Then, for each \(i\in \{1\ldots l\}\), (KP4) says that

$$\begin{aligned} s_{w_i}=\sum _{\lambda \in w_i\Lambda ^{n'-d(\mu _i)}}s_\lambda s_\lambda ^*, \end{aligned}$$

and we can write

$$\begin{aligned} s_{\mu _i}u_{w_i,g_i}s_{\nu _i}^*&=s_{\mu _i}(s_{s(\mu _i)})u_{w_i,g_i}s_{\nu _i}^*\\&=\sum s_{\mu _i}(s_\lambda s_\lambda ^*)u_{w_i,g_i}s_{\nu _i}^*\\&=\sum s_{\mu _i \lambda }\bigg (s_{\nu _i} u_{w_i,g_i}^*s_{\lambda }\bigg )^*\\&=\sum s_{\mu _i \lambda }\bigg (s_{\nu _i}u_{g_i^{-1} \cdot w_i,g_i^{-1}} s_\lambda \bigg )^*\\&=\sum s_{\mu _i \lambda }\bigg (s_{\nu _i} s_{g_i^{-1} \cdot \lambda }u_{g_i^{-1} \cdot s(\lambda ),\varphi (g_i^{-1},\lambda )} \bigg )^*\\&=\sum s_{\mu _i\lambda } u_{\varphi (g_i^{-1},\lambda )^{-1}g_i^{-1} \cdot s(\lambda ),\varphi (g_i^{-1},\lambda )^{-1}} s_{\nu _i (g_i^{-1} \cdot \lambda )}^* \end{aligned}$$

where the above summations are on \(\lambda \in w_i\Lambda ^{n'-d(\mu _i)}\). So, in each term of (4.1), we may assume \(d(\mu _i)=n'\) and \(d(\nu _i)=n'-n\). Now, for any \(1\le j\le l\), (KP3) yields that

$$\begin{aligned} s_{\mu _j}^*b s_{\nu _j}=s_{\mu _j}^*\bigg ( \sum _{i=1}^lr_i s_{\mu _i}u_{w_i,g_i}s_{\nu _i}^*\bigg ) s_{\nu _j}= \sum _{i\in [j]}r_i u_{w_i,g_i}, \end{aligned}$$

where \([j]:=\{1\le i\le l:(\mu _i,\nu _i)=(\mu _j,\nu _j)\}\). Thus

$$\begin{aligned} \phi \bigg (\sum _{i\in [j]}r_i u_{w_i,g_i} \bigg )=\phi (s_{\mu _j}^*)\phi (b)\phi (s_{\nu _j})=0 \end{aligned}$$

and hypothesis forces \(\sum _{i\in [j]}r_i u_{w_i,g_i}=0\). Therefore,

$$\begin{aligned} \sum _{i\in [j]}r_i s_{\mu _i}u_{w_i,g_i}s_{\nu _i}^*=s_{\mu _j}\bigg (\sum _{i\in [j]}r_i u_{w_i,g_i} \bigg ) s_{\nu _j}^*=0. \end{aligned}$$

Since the index set \(\{1,\ldots ,l\}\) is a disjoint union of [j]’s, we obtain \(b=0\). It follows that \(\phi \) is injective. \(\square \)

5 \({\textrm{EP}}_R(G,\Lambda )\) as a Steinberg Algebra

In this section, we want to prove an Steinberg algebra model for \({\textrm{EP}}_R(G,\Lambda )\). Although our result will be the k-graph generalization of [9, Theorem B], note that our proof relies on the graded uniqueness theorem, Theorem 4.2, and is completely different from that of [9, Theorem B]. This gives us a much easier and shorter proof.

Let us first review some terminology about groupoids; see [19] for more details. A groupoid is a small category \({\mathcal {G}}\) with inverses. For each \(\alpha \in {\mathcal {G}}\), we may define the range \(r(\alpha ):=\alpha \alpha ^{-1}\) and the source \(s(\alpha ):=\alpha ^{-1}\alpha \) satisfying \(r(\alpha )\alpha =\alpha =\alpha s(\alpha )\). It follows that for every \(\alpha ,\beta \in {\mathcal {G}}\), the composition \(\alpha \beta \) is well defined if and only if \(s(\alpha )=r(\beta )\). The unit space of \({\mathcal {G}}\) is \({\mathcal {G}}^{(0)}:=\{\alpha ^{-1}\alpha : \alpha \in {\mathcal {G}}\}\). Throughout the paper we work with topological groupoids, which are ones equipped with a topology such that the maps r and s are continuous. Then a bisection is a subset \(B\subseteq {\mathcal {G}}\) such that both restrictions \(r|_B\) and \(s|_B\) are homeomorphisms. In case \({\mathcal {G}}\) has a basis of compact open bisections, \({\mathcal {G}}\) is called an ample groupoid.

Let \((G,\Lambda )\) be a self-similar k-graph. We also recall the groupoid \({\mathcal {G}}_{G,\Lambda }\) introduced in [15]. Let \(C({\mathbb {N}}^k,G)\) be the group of all maps form \({\mathbb {N}}^k\) to G with the pointwise multiplication. For \(f,g\in C({\mathbb {N}}^k,G)\), define the equivalence relation \(f\sim g\) in case there exists \(n_0\in {\mathbb {N}}^k\) such that \(f(n)=g(n)\) for all \(n\ge n_0\). Write \(Q({\mathbb {N}}^k,G):=C({\mathbb {N}}^k,G)/\sim \). Also, for each \(z\in {\mathbb {Z}}^k\), let \({\mathcal {T}}_z: C({\mathbb {N}}^k,G)\rightarrow C({\mathbb {N}}^k,G)\) be the automorphism defined by

$$\begin{aligned} {\mathcal {T}}_z(f)(n)=\left\{ \begin{array}{ll} f(n-z) &{} n-z\ge 0 \\ e_G &{} \textrm{otherwise} \end{array} \right. \qquad (f\in C({\mathbb {N}}^k,G),~ n\in {\mathbb {N}}^k). \end{aligned}$$

Then \({\mathcal {T}}_z\) induces an automorphism, denoted again by \({\mathcal {T}}_z\), on \(Q({\mathbb {N}}^k,G)\), which is \({\mathcal {T}}_z([f])=[{\mathcal {T}}_z(f)]\). So, \({\mathcal {T}}:{\mathbb {Z}}^k\rightarrow \textrm{Aut}Q({\mathbb {N}}^k,G)\) is a homomorphism and we consider the semidirect product group \(Q({\mathbb {N}}^k,G)\rtimes _{{\mathcal {T}}} {\mathbb {Z}}^k\).

Note that for every \(g\in G\) and \(x\in \Lambda ^\infty \), one may define \(\varphi (g,x)\in C({\mathbb {N}}^k,G)\) by

$$\begin{aligned} \varphi (g,x)(n):=\varphi (g,x(0,n)) \qquad (n\in {\mathbb {N}}^k). \end{aligned}$$

Moreover, [15, Lemma 3.7] says that there exists a unique action \(G\curvearrowright \Lambda ^\infty \) by defining

$$\begin{aligned} (g\cdot x)(m,n):=\varphi (g,x(0,m)) \cdot x(m,n) \qquad ((m,n)\in \Omega _k) \end{aligned}$$

for every \(g\in G\) and \(x\in \Lambda ^\infty \).

Definition 5.1

A self-similar k-graph \((G,\Lambda )\) is said to be pseudo-free if for any \(g\in G\) and \(\mu \in \Lambda \), \(g\cdot \mu =\mu \) and \(\varphi (g,\mu )=e_G\) imply \(g=e_G\).

According to [15, Lemma 5.6], in case \((G,\Lambda )\) is pseudo-free, then we have

$$\begin{aligned} g\cdot \mu =h \cdot \mu ~~ \textrm{and} ~~ \varphi (g,\mu )=\varphi (h,\mu ) ~~ \Longrightarrow ~~ g=h \end{aligned}$$

for every \(g,h\in G\) and \(\mu \in \Lambda \).

Definition 5.2

Associated to \((G,\Lambda )\) we define the subgroupoid

$$\begin{aligned} {\mathcal {G}}_{G,\Lambda }&:=\left\{ \bigg (\mu (g\cdot x);{\mathcal {T}}_{d(\mu )}([\varphi (g,x)]),d(\mu )-d(\nu );\nu x\bigg ) :g\in G,\right. \\&\qquad \left. ~\mu ,\nu \in \Lambda ,~ s(\mu )=g\cdot s(\nu ) \right\} \end{aligned}$$

of \(\Lambda ^\infty \times \bigg (Q({\mathbb {N}}^k,G)\rtimes _{{\mathcal {T}}} {\mathbb {Z}}^k\bigg ) \times \Lambda ^\infty \) with the range and source maps

$$\begin{aligned} r(x;[f],n-m;y)=x \quad \textrm{ and }\quad s(x;[f],n-m;y)=y. \end{aligned}$$

Note that if we set

$$\begin{aligned} Z(\mu ,g,\nu ):=\left\{ \bigg (\mu (g\cdot x);{\mathcal {T}}_{d(\mu )}([\varphi (g,x)]),d(\mu )-d(\nu );\nu x\bigg ):x\in s(\nu )\Lambda ^\infty \right\} , \end{aligned}$$

then the basis

$$\begin{aligned} {\mathcal {B}}_{G,\Lambda }:=\left\{ Z(\mu ,g,\nu ):\mu ,\nu \in \Lambda ,g\in G,s(\mu )=g\cdot s(\nu )\right\} \end{aligned}$$

induces a topology on \({\mathcal {G}}_{G,\Lambda }\). In case \((G,\Lambda )\) is pseudo-free, [16, Proposition 3.11] shows that \({\mathcal {G}}_{G,\Lambda }\) is a Hausdorff groupoid with compact open base \({\mathcal {B}}_{G,\Lambda }\).

Definition 5.3

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph and R a unital commutative \(*\)-ring. Then the Steinberg algebra associated to \((G,\Lambda )\) is the R-algebra

$$\begin{aligned} A_R({\mathcal {G}}_{G,\Lambda }):=\textrm{span}_R\{1_B: B ~ \mathrm {is ~ a ~ compact ~ open ~ bisection}\} \end{aligned}$$

endowed with the pointwise addition, the multiplication \(fg(\gamma ):=\sum _{\alpha \beta =\gamma }f(\alpha ) g(\beta )\), and the involution \(f^*(\gamma ):=f(\gamma ^{-1})^*\) for all \(\gamma \in {\mathcal {G}}_{G,\Lambda }\).

To prove Theorem 5.5, we need the following lemma.

Lemma 5.4

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph. Let \(v,w\in \Lambda ^0\) and \(g,h\in G\) with \(g\cdot v=h\cdot v=w\). Then \(Z(v,g,w)\cap Z(v,h,w)= \emptyset \) whenever \(g\ne h\).

Proof

Suppose that \((g\cdot x;[\varphi (g,x)],0;x)=(h \cdot y,[\varphi (h,y)],0;y)\in Z(v,g,w)\cap Z(v,h,w)\) where \(x,y\in Z(w)\). Then \(y=x\), \(g\cdot x=h \cdot x\) and \([\varphi (g,x)]=[\varphi (h,x)]\). Since \((G,\Lambda )\) is pseudo-free, [15, Corollary 5.6] implies that \(g=h\). \(\square \)

Theorem 5.5

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph. Then there is a (unique) \(*\)-algebra isomorphism \(\phi :{\textrm{EP}}_R(G,\Lambda )\rightarrow A_R({\mathcal {G}}_{G,\Lambda })\) such that

$$\begin{aligned} \phi (s_\mu )=1_{Z(\mu , e_G, s(\mu ))} \quad \textrm{ and }\quad \phi (u_{v,g})=1_{Z(v,g,g^{-1} \cdot v )} \end{aligned}$$

for every \(\mu \in \Lambda \), \(v\in \Lambda ^0\), and \(g\in G\). In particular, the elements \(rs_\mu \) and \(ru_{v,g}\) with \(r\in R{\setminus }\{0\}\) are all nonzero.

Proof

For each \(v\in \Lambda ^0\), \(\mu \in \Lambda \) and \(g\in G\), define

$$\begin{aligned} S_\mu :=1_{Z(\mu , e_G, s(\mu ))} \quad \textrm{ and }\quad U_{v,g}:=1_{Z(v,g,g^{-1} \cdot v )}. \end{aligned}$$

Since \(S_\mu ^*=1_{Z(\mu ,e_G,s(\mu ))^{-1}}=1_{Z(s(\mu ),e_G,\mu )}\) and \(U_{v,g}^*=1_{Z(v,g,g^{-1} \cdot v )^{-1}}=1_{Z(g^{-1} \cdot v,g^{-1},v)}\), a long but straightforward computation shows that \(\{S_\mu ,U_{v,g}\}\) is a \((G,\Lambda )\)-family in \(A_R({\mathcal {G}}_{G,\Lambda })\). Then, by the universal property, such \(*\)-homomorphism \(\phi \) exists.

[16, Proposition 3.11] says that \({\mathcal {G}}_{G,\Lambda }\) is ample with compact open base \({\mathcal {B}}_{G,\Lambda }\). Since each element \(Z(\mu ,g,\nu )\) of \({\mathcal {B}}_{G,\Lambda }\) can be written as

$$\begin{aligned} Z(\mu ,g,\nu )=Z(\mu ,e_G,s(\mu ))Z(s(\mu ),g,s(\nu ))Z(\nu ,e_G,s(\nu ))^{-1}, \end{aligned}$$

\(\phi \) is surjective.

We will show the injectivity of \(\phi \) by applying the graded uniqueness theorem. Note that the continuous 1-cocycle \(c:{\mathcal {G}}_{G,\Lambda }\rightarrow {\mathbb {Z}}^k\), defined by \(c(\mu (g\cdot x);[f],d(\mu )-d(\nu );\nu x):=d(\mu )-d(\nu )\), induces a \({\mathbb {Z}}^k\)-grading on \(A_R({\mathcal {G}}_{G,\Lambda })\). Also, \(\phi \) preserves the \({\mathbb {Z}}^k\)-grading because it does on the generators. Now, to apply Theorem 4.2, we assume \(\phi (a)=0\) for an element of the form \(a=\sum _{i=1}^l r_i u_{v,g_i}\) with \(g_i^{-1} \cdot v =g_j^{-1} \cdot v \) for \(1\le i,j\le l\). We may also assume that the \(g_i\)’s are distinct (otherwise, combine the terms with same \(g_i\)’s). We then have

$$\begin{aligned} \phi (a)=\sum _{i=1}^l r_i1_{Z(v,g_i,g_i^{-1} \cdot v )}=0. \end{aligned}$$

Lemma 5.4 says that the bisections \(Z(v,g_i,g_i^{-1} \cdot v )\) are pairwise disjoint. Hence, for each i, if we pick some \(\alpha \in Z(v,g_i,g_i^{-1} \cdot v )\), then \(r_i=\phi (a)(\alpha )=0\). Therefore \(a=0\), and Theorem 4.2 concludes that \(\phi \) is injective. We are done. \(\square \)

Combining [20, Theorem 6.7], [15, Theorem 5.9], and Theorem 5.5 gives the next corollary. (Although in [15] it is supposed \(|\Lambda ^0|<\infty \), but [15, Theorem 5.9] holds also for \(\Lambda \) with infinitely many vertices.)

Corollary 5.6

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph over an amenable group G. Then the complex algebra \(\textrm{EP}_{\mathbb {C}}(G,\Lambda )\) is a dense subalgebra of \({\mathcal {O}}_{G,\Lambda }\) introduced in [16].

In the following, we see that the Kumjian-Pask algebra \(\textrm{KP}_R(\Lambda )\) from [2] can be embedded in \({\textrm{EP}}_R(G,\Lambda )\).

Corollary 5.7

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph. Let the Kumjian-Pask algebra \(\textrm{KP}_R(\Lambda )\) be generated by a Kumjian-Pask \(\Lambda \)-family \(\{t_\mu :\mu \in \Lambda \}\). Then the map \(t_\mu \mapsto s_\mu \) embeds \(\textrm{KP}_R(\Lambda )\) into \({\textrm{EP}}_R(G,\Lambda )\) as a \(*\)-subalgebra.

Proof

We know that \(\textrm{KP}_R(\Lambda )\) is \({\mathbb {Z}}^k\)-graded by the homogenous components

$$\begin{aligned} \textrm{KP}_R(\Lambda )_n:=\textrm{span}_R \left\{ t_\mu t_\nu ^*:\mu ,\nu \in \Lambda , d(\mu )-d(\nu )=n\right\} . \end{aligned}$$

for all \(n\in {\mathbb {Z}}^k\). Then, the universal property of Kumjian-Pask algebras gives a graded \(*\)-algebra homomorphism \(\phi :\textrm{KP}_R(\Lambda )\rightarrow {\textrm{EP}}_R(G,\Lambda )\) such that \(\phi (t_\mu ):=s_\mu \) and \(\phi (t_\mu ^*):=s_\mu ^*\) for every \(\mu \in \Lambda \). Moreover, Theorem 5.5 shows that \(\phi (rt_\mu )=rs_\mu \ne 0\) for all \( r\in R{\setminus } \{0\}\) and \(\mu \in \Lambda \). Therefore, the graded uniqueness theorem for Kumjian-Pask algebras [2, Theorem 4.1] implies that \(\phi \) is injective. \(\square \)

Definition 5.8

Let \({\mathcal {G}}\) be a topological groupoid. We say that \({\mathcal {G}}\) is topologically principal if the set of units with trivial isotropy group, that is \(\{u\in {\mathcal {G}}^{(0)}: s^{-1}(u)\cap r^{-1}(u)=\{u\}\}\), is dense in \({\mathcal {G}}^{(0)}\).

The analogue of the topologically principal property for self-similar k-graphs is G-aperiodicity (see [15, Proposition 6.5]).

Definition 5.9

Let \((G,\Lambda )\) be a self-similar k-graph. Then \(\Lambda \) is said to be G-aperiodic if for every \(v\in \Lambda ^0\), there exists \(x\in v\Lambda ^\infty \) with the property that

$$\begin{aligned} x(p,\infty )=g\cdot x(q,\infty ) ~~\Longrightarrow ~~ g=e_G ~~ \textrm{and} ~~ p=q \qquad (\forall g\in G,~ \forall p,q\in {\mathbb {N}}^k). \end{aligned}$$

Theorem 5.10

(The Cuntz–Krieger uniqueness) Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph. Let \((G,\Lambda )\) be also G-aperiodic. Suppose that \(\phi :{\textrm{EP}}_R(G,\Lambda )\rightarrow A\) is a \(*\)-algebra homomorphism from \({\textrm{EP}}_R(G,\Lambda )\) into a \(*\)-algebra A such that \(\phi (rs_v)\ne 0\) for all \(0\ne r\in R\) and \(v\in \Lambda ^0\). Then \(\phi \) is injective.

Proof

First note that \({\mathcal {G}}_{G,\Lambda }\) is a Hausdorff ample groupoid by [16, Proposition 3.11], and that \({\mathcal {B}}_{G,\Lambda }\) is a basis for \({\mathcal {G}}_{G,\Lambda }\) consisting compact open bisections. Also, [16, Lemma 3.12] says that \({\mathcal {G}}_{G,\Lambda }\) is topologically principal (so is effective in particular). So, we may apply [5, Theorem 3.2].

Denote by \(\psi :{\textrm{EP}}_R(G,\Lambda )\rightarrow A_R({\mathcal {G}}_{G,\Lambda })\) the isomorphism of Theorem 5.5. If on the contrary \(\phi \) is not injective, then neither is \({\widetilde{\phi }}:=\phi \circ \psi ^{-1}:A_R({\mathcal {G}}_{G,\Lambda })\rightarrow A\). Thus, by [5, Theorem 3.2], there exists a compact open subset \(K\subseteq {\mathcal {G}}_{G,\Lambda }^{(0)}\) and \(r\ne 0\) such that \({\widetilde{\phi }}(r1_K)=0\). Since K is open, there is a unit \(U=Z(\mu ,e_G,\mu )\in {\mathcal {B}}_{G,\Lambda }\) such that \(U\subseteq K\). So we get

$$\begin{aligned} \phi (rs_\mu s_\mu ^*)={\widetilde{\phi }}(r1_U)={\widetilde{\phi }}(r1_{U\cap K})={\widetilde{\phi }}(r1_K) {\widetilde{\phi }}(1_U)=0, \end{aligned}$$

and hence

$$\begin{aligned} \phi (rs_{s(\mu )})=\phi (s_\mu ^*)\phi (rs_\mu s_\mu ^*)\phi (s_\mu )=0. \end{aligned}$$

This contradicts the hypothesis, and therefore \(\phi \) is injective. \(\square \)

6 Ideal Structure

By an ideal we mean a two-sided and self-adjoint one. In this section, we characterize basic, \({\mathbb {Z}}^k\)-graded and diagonal-invariant ideals of \({\textrm{EP}}_R(G,\Lambda )\), which are exactly all basic \(Q({\mathbb {N}}^k,G)\rtimes _{{\mathcal {T}}}{\mathbb {Z}}^k\)-graded ones.

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph. Since \({\mathcal {G}}_{G,\Lambda }\) is a Hausdorff ample groupoid [15, Theorem 5.8], \({\mathcal {G}}_{G,\Lambda }^{(0)}\) is both open and closed, and for every \(f\in A_R({\mathcal {G}}_{G,\Lambda })\) the restricted function \(f|_{{\mathcal {G}}_{G,\Lambda }^{(0)}}=f\chi _{{\mathcal {G}}_{G,\Lambda }^{(0)}}\) lies again in \(A_R({\mathcal {G}}_{G,\Lambda })\). Then \(A_R({\mathcal {G}}_{G,\Lambda }^{(0)})\) is a \(*\)-subalgebra of \(A_R({\mathcal {G}}_{G,\Lambda })\) and there is a conditional expectation \({\mathcal {E}}:A_R({\mathcal {G}}_{G,\Lambda })\rightarrow A_R({\mathcal {G}}_{G,\Lambda }^{(0)})\) defined by \({\mathcal {E}}(f)=f|_{{\mathcal {G}}_{G,\Lambda }^{(0)}}\) for \(f\in A_R({\mathcal {G}}_{G,\Lambda })\). Let \(D:=\textrm{span}_R\{s_\mu s_\mu ^*:\mu \in \Lambda \}\) be the diagonal of \({\textrm{EP}}_R(G,\Lambda )\). In light of Theorem 5.5, it is easy to check that the expectation is \({\mathcal {E}}:{\textrm{EP}}_R(G,\Lambda )\rightarrow D\) defined by

$$\begin{aligned} {\mathcal {E}}\bigg (s_\mu u_{s(\mu ),g}s_\nu ^*\bigg )=\delta _{\mu ,\nu }\delta _{g,e_G} s_\mu s_\mu ^* \qquad (\mu ,\nu \in \Lambda ,~s(\mu )=g\cdot s(\nu )). \end{aligned}$$

Definition 6.1

An ideal I of \({\textrm{EP}}_R(G,\Lambda )\) is called diagonal-invariant whenever \({\mathcal {E}}(I)\subseteq I\). Also, I is said to be basic if \(rs_v\in I\) implies \(s_v\in I\) for all \(v\in \Lambda ^0\) and \(r\in R{\setminus } \{0\}\).

Definition 6.2

Let \((G,\Lambda )\) be a self-similar k-graph. A subset \(H\subseteq \Lambda ^0\) is called

  1. (1)

    G-hereditary if \(r(\mu )\in H ~~ \Longrightarrow ~~ g\cdot s(\mu )\in H\) for all \(g\in G\) and \(\mu \in \Lambda \);

  2. (2)

    G-saturated if \(v\in \Lambda ^0\) and \(s(v\Lambda ^n) \subseteq H~~ \mathrm {for ~ some ~} n\in {\mathbb {N}}^k ~~ \Longrightarrow ~~ v\in H\).

In the following, given any \(H\subseteq \Lambda ^0\), we denote by \(I_H\) the ideal of \({\textrm{EP}}_R(G,\Lambda )\) generated by \(\{s_v:v\in H\}\). Also, for each ideal I of \({\textrm{EP}}_R(G,\Lambda )\), we define \(H_I:=\{v\in \Lambda ^0: s_v\in I\}\).

To prove Theorem 6.8 we need some structural lemmas about the ideals \(I_H\) and associated quotients \({\textrm{EP}}_R(G,\Lambda )/I_H\).

Lemma 6.3

If I is an ideal of \({\textrm{EP}}_R(G,\Lambda )\), then \(H_I:=\{v\in \Lambda ^0:s_v\in I\}\) is a G-saturated G-hereditary subset of \(\Lambda ^0\).

Proof

The proof is straightforward. \(\square \)

Lemma 6.4

Let H be a G-saturated G-hereditary subset of \(\Lambda ^0\) and \(I_H\) the ideal of \({\textrm{EP}}_R(G,\Lambda )\) generated by \(\{s_v:v\in H\}\). Then we have

$$\begin{aligned} I_H=\textrm{span}_R\left\{ s_\mu u_{s(\mu ),g}s_\nu ^*:g\in G, ~ s(\mu )=g\cdot s(\nu )\in H\right\} , \end{aligned}$$
(6.1)

and \(I_H\) is a \({\mathbb {Z}}^k\)-graded diagonal-invariant ideal.

Proof

Denote by J the right-hand side of (6.1). The identity

$$\begin{aligned} s_\mu u_{s(\mu ),g}s_\nu ^*=s_\mu (s_{s(\mu )})u_{s(\mu ),g}s_\nu ^* \end{aligned}$$

yields \(J\subseteq I_H\). Also, using the description of \({\textrm{EP}}_R(G,\Lambda )\) in Proposition 2.7, it is straightforward to check that J is an ideal of \({\textrm{EP}}_R(G,\Lambda )\). So, by \(s_v=s_vu_{v,e_G}s_v^*\), J contains all generators of \(I_H\), and we have proved (6.1).

Now, (6.1) says that \(I_H\) is spanned by its homogenous elements, hence it is a graded ideal. Moreover, let \(a=\sum _{i=1}^l s_{\mu _i}u_{s(\mu _i),g_i}s_{\nu _i}^*\in I_H\) such that \(s(\mu _i)=g\cdot s(\nu _i)\in H\). Then, in particular, each term of a with \(g_i=e_G\) belongs to \(I_H\). Therefore, \({\mathcal {E}}(a)\in I_H\), and \(I_H\) is diagonal-invariant. \(\square \)

Let H be a G-saturated G-hereditary subset of \(\Lambda ^0\) and consider the k-subgraph \(\Lambda \setminus \Lambda H\). Then the restricted action \(G\curvearrowright \Lambda \setminus \Lambda H\) is well defined, and hence \((G,\Lambda {\setminus } \Lambda H, \varphi |_{G\times \Lambda {\setminus } \Lambda H})\) is also a self-similar k-graph. So we have:

Lemma 6.5

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph. If H is a G-saturated G-hereditary subset of \(\Lambda ^0\), then \((G,\Lambda \setminus \Lambda H)\) is a pseudo-free self-similar k-graph.

Proof

The proof is straightforward. \(\square \)

Lemma 6.6

Let H be a G-saturated G-hereditary subset of \(\Lambda ^0\). For every \(v\in \Lambda ^0\) and \(r\in R{\setminus }\{0\}\), \(rs_v\in I_H\) implies \(v\in H\).

Proof

Let \(\{t_\mu , w_{v,g}\}\) be the generators of \(\textrm{EP}_R(G,\Lambda \setminus \Lambda H)\). If we define

$$\begin{aligned} S_\mu :=\left\{ \begin{array}{ll} t_\mu &{} s(\mu )\notin H \\ 0 &{} \textrm{otherwise} \end{array} \right. \quad \text { and } \quad U_{v,g}:=\left\{ \begin{array}{ll} w_{v,g} &{} v\notin H \\ 0 &{} \textrm{otherwise,} \end{array} \right. \end{aligned}$$

then \(\{S_\mu ,U_{v,g}\}\) is a \((G,\Lambda )\)-family in \(\textrm{EP}_R(G,\Lambda \setminus \Lambda H)\), and by the universality, there is a \(*\)-homomorphism \(\psi :{\textrm{EP}}_R(G,\Lambda )\rightarrow \textrm{EP}_R(G,\Lambda {\setminus }\Lambda H)\) such that \(\psi (s_\mu )=S_\mu \) and \(\psi (u_{v,g})=U_{v,g}\) for all \(\mu \in \Lambda \), \(v\in \Lambda ^0\) and \(g\in G\). Since \(\psi (s_v)=0\) for every \(v\in H\), we have \(I_H\subseteq \ker \psi \). On the other hand, Theorem 5.5 implies that all \(\psi (rs_v)=rt_v\) are nonzero for \(v\in \Lambda ^0{\setminus } H\) and \(r\in R{\setminus }\{0\}\).

Now assume \(rs_v\in I_H\) for some \(v\in \Lambda ^0\) and \(r\in R{\setminus }\{0\}\). If \(v\in \Lambda ^0{\setminus } H\), then \(\psi (rs_v)=rt_v\ne 0\), and we get \(rs_v\notin \ker \psi \supseteq I_H\), a contradiction. \(\square \)

In fact, Lemma 6.6 says that \(I_H\) is a basic ideal with \(H_{I_H}=H\) for every G-saturated G-hereditary subset H of \(\Lambda ^0\).

Proposition 6.7

Let H be a G-saturated G-hereditary subset of \(\Lambda ^0\). Let \(\{t_\mu ,w_{v,g}\}\) be the \((G,\Lambda \setminus \Lambda H)\)-family generating \(\textrm{EP}_R(G,\Lambda \setminus \Lambda H)\). Then the map \(\psi : \textrm{EP}_R(G,\Lambda {\setminus } \Lambda H) \rightarrow {\textrm{EP}}_R(G,\Lambda )/I_H\) defined by

$$\begin{aligned} \psi (t_\mu w_{s(\mu ),g} t_\nu ^*):=s_\mu u_{s(\mu ),g} s_\nu ^* +I_H \qquad (\mu ,\nu \in \Lambda \setminus \Lambda H,~ g\in G) \end{aligned}$$

is an (R-algebra) \(*\)-isomorphism.

Proof

If we set \(T_\mu := s_\mu +I_H\) and \(W_{v,g}:=u_{v,g}+I_H\) for every \(v\in \Lambda ^0,~\mu \in \Lambda \), and \(g\in G\), then \(\{T_\mu ,W_{v,g}\}\) is a \((G,\Lambda \setminus \Lambda H)\)-family in \({\textrm{EP}}_R(G,\Lambda )/I_H\) (the relations of Definition 2.5 for \(\{T_\mu , W_{v,g}\}\) immediately follow from those for \(\{s_\mu ,u_{v,g}\}\)). So, the universality of \(\textrm{EP}_R(G,\Lambda \setminus \Lambda H)\) gives such \(*\)-homomorphism \(\psi \). Note that \(s_\mu \in I_H\) for each \(\mu \in \Lambda H\) by (6.1), which gives the surjectivity of \(\psi \).

To prove the injectivity, we apply the graded uniqueness theorem, Theorem 4.2. First, since \(I_H\) is a \({\mathbb {Z}}^k\)-graded ideal, \({\textrm{EP}}_R(G,\Lambda )/I_H\) has a natural \({\mathbb {Z}}^k\)-grading and \(\psi \) is a graded homomorphism. Thus, we fix an element in \(\textrm{EP}_R(G,\Lambda \setminus \Lambda H)\) of the form \(a=\sum _{i=1}^l r_iw_{v,g_i}\) such that \(v\in \Lambda ^0{\setminus } H\) and \(g_i^{-1} \cdot v =g_j^{-1} \cdot v \) for all \(1\le i,j\le l\). Without loss of generality, we may also suppose that the \(g_i\)’s are distinct. If \(\psi (a)=0\), then \(\psi (a)=\sum _{i=1}^l r_iu_{v,g_i}+I_H=I_H\) and \(\sum _{i=1}^l r_iu_{v,g_i}\in I_H\). Thus, for each \(1\le j\le l\), we have

$$\begin{aligned} \bigg (\sum _{i=1}^l r_iu_{v,g_i}\bigg )u_{g_j^{-1} \cdot v, g_j^{-1}}=\sum _{i=1}^l r_iu_{v,g_ig_j^{-1}}\in I_H \qquad (\mathrm {by ~ eq.~ (5) ~ in~ Definition~ 2.5}) \end{aligned}$$

and since \(I_H\) is diagonal-invariant,

$$\begin{aligned} r_js_v=r_ju_{v,e_G}={\mathcal {E}}\left( \sum _{i=1}^l r_iu_{v,g_ig_j^{-1}}\right) \in I_H. \end{aligned}$$

As \(v\notin H\), Lemma 6.6 forces \(r_j=0\) for each \(1\le j\le l\), hence \(a=0\). Now Theorem 4.2 implies that \(\psi \) is an isomorphism. \(\square \)

Theorem 6.8

Let \((G,\Lambda )\) be pseudo-free self-similar k-graph. Then \(H\mapsto I_H\) is a one-to-one correspondence between G-saturated G-hereditary subsets of \(\Lambda ^0\) and basic, \({\mathbb {Z}}^k\)-graded and diagonal-invariant ideals of \({\textrm{EP}}_R(G,\Lambda )\), with inverse \(I\mapsto H_I\).

Proof

The injectivity of \(H\mapsto I_H\) follows from Lemma 6.6. Indeed, if \(I_H=I_K\) for G-saturated G-hereditary subsets \(H,K\subseteq \Lambda ^0\), then Lemma 6.6 yields that \(H=H_{I_H}=H_{I_K}=K\).

To see the surjectivity, we take a basic, \({\mathbb {Z}}^k\)-graded and diagonal-invariant ideal I of \({\textrm{EP}}_R(G,\Lambda )\), and then prove \(I=I_{H_I}\). Write \(J:=I_{H_I}\) for convenience. By Proposition 6.7 we may consider \(\textrm{EP}_R(G,\Lambda \setminus \Lambda H_I)\cong {\textrm{EP}}_R(G,\Lambda )/J\) as a \(*\)-R-algebra. Let \(\{s_\mu ,u_{v,g}\}\) and \(\{t_\mu ,w_{v,g}\}\) be the generators of \({\textrm{EP}}_R(G,\Lambda )\) and \({\textrm{EP}}_R(G,\Lambda )/J\), respectively. Since \(J\subseteq I\), we may define the quotient map \(q:{\textrm{EP}}_R(G,\Lambda )/J\rightarrow {\textrm{EP}}_R(G,\Lambda )/I\) such that

$$\begin{aligned} q(t_\mu )=s_\mu +I \quad \textrm{ and }\quad q(w_{v,g})=u_{v,g}+I \end{aligned}$$

for all \(\mu \in \Lambda \), \(v\in \Lambda ^0\) and \(g\in G\). Notice that q preserves the grading because I is a \({\mathbb {Z}}^k\)-graded ideal. So, we can apply Theorem 4.2 to show that q is an isomorphism. To do this, fix an element of the form \(a=\sum _{i=1}^l r_iw_{v,g_i}\) with \(v\in \Lambda ^0{\setminus } H_I\) such that \(g_i^{-1} \cdot v =g_j^{-1} \cdot v \) for \(1\le i,j\le l\) and \(q(a)=0\). Then \(\sum _{i=1}^l r_iu_{v,g_i}\in I\). As before, we may also assume that the \(g_i\)’s are distinct. Thus, for each \(1\le j\le l\), we have

$$\begin{aligned} b_j:=\bigg (\sum _{i=1}^l r_iu_{v,g_i}\bigg )\bigg (u_{g_j^{-1} \cdot v,g_j^{-1}}\bigg )=\sum _{i=1}^l r_iu_{v,g_ig_j^{-1}}\in I. \end{aligned}$$

Since I is diagonal-invariant and basic, the case \(r_j\ne 0\) yields \(r_js_v=r_ju_{v,e_G}={\mathcal {E}}(b_j)\in I\), and thus \(s_v\in I\) and \(v\in H_I\), which contradicts the choice of v. It follows that \(r_j\)’s are all zero, and hence \(a=0\). Now Theorem 4.2 implies that q is injective, or equivalently \(I=J=I_{H_I}\) as desired. \(\square \)

In the end, we remark the following about \(Q({\mathbb {N}}^k,G)\rtimes _{{\mathcal {T}}}{\mathbb {Z}}^k\)-graded ideals of \({\textrm{EP}}_R(G,\Lambda )\).

Remark 6.9

Let \((G,\Lambda )\) be a pseudo-free self-similar k-graph and \({\mathcal {G}}_{G,\Lambda }\) be the associated groupoid. Denote by \(\Gamma :=Q({\mathbb {N}}^k,G)\rtimes _{{\mathcal {T}}}{\mathbb {Z}}^k\) the group introduced in Sect. 5. If we define \(c:{\mathcal {G}}_{G,\Lambda }\rightarrow \Gamma \) by \(c(x;\gamma ;y):=\gamma \), then c is a cocycle on \({\mathcal {G}}_{G,\Lambda }\) because \(c(\alpha \beta )=c(\alpha )*_\Gamma c(\beta )\) for all \(\alpha ,\beta \in {\mathcal {G}}_{G,\Lambda }\) with \(s(\alpha )=r(\beta )\). Hence, it induces a \(\Gamma \)-grading on \(A_R({\mathcal {G}}_{G,\Lambda })={\textrm{EP}}_R(G,\Lambda )\) with the homogenous components

$$\begin{aligned} A_\gamma :=\textrm{span}_R\{1_V:V\subseteq c^{-1}(\gamma ) \mathrm {~is~a~compact~open~bisection~}\} \end{aligned}$$

(see [6, Proposition 5.1] for example). By a similar argument as in [6, §6.5] and combining Theorem 5.5 and [6, Theorem 5.3], we may obtain that the ideals of the form \(I_H\), described in Theorem 6.8 above, are precisely the basic, \(\Gamma \)-graded ideals of \({\textrm{EP}}_R(G,\Lambda )\).