1. Introduction
If E is a graph and $L_K(E)$ is its Leavitt path algebra, it is known that every graded ideal I of $L_K(E)$ is uniquely determined by a pair $(H,S)$ of certain subsets of vertices, known as an admissible pair (we review the definition in Section 2.2). For every such admissible pair, one can define a graph, referred to as the generalised hedgehog graph in [Reference Abrams, Ara and Siles Molina1], such that I is isomorphic to the Leavitt path algebra of this graph. The name ‘hedgehog’ comes from the construction in which one attaches new edges (the ‘spines’) to $H\cup S$ (the ‘body’). We modify this construction by replacing the added edges with possibly fewer paths which are then necessarily of length larger than one. Because of the longer spines, we call the resulting graph the porcupine graph of $(H,S).$ In the main result of the paper, Theorem 3.3, we construct a graded $*$ -isomorphism of I and the Leavitt path algebra of the porcupine graph of the admissible pair corresponding to $I.$
We adapt our result to graph $C^*$ -algebras also. In Corollary 3.4, we show that, for every closed gauge-invariant ideal J of a graph $C^*$ -algebra, there is a gauge-invariant (and graded in the $C^*$ -algebra sense) $*$ -isomorphism which maps the graph $C^*$ -algebra of the porcupine graph of J onto $J.$ Just as in the algebraic case, it is known that such an ideal J is $*$ -isomorphic to a graph $C^*$ -algebra, but it was not clear whether there is a gauge-invariant $*$ -isomorphism between J and a graph $C^*$ -algebra. We show that such an isomorphism indeed exists.
2. Prerequisites
2.1. Graded rings and $\ast $ -rings
A ring R (not necessarily unital) is graded by a group $\Gamma $ if $R=\bigoplus _{\gamma \in \Gamma } R_\gamma $ for additive subgroups $R_\gamma $ and $R_\gamma R_\delta \subseteq R_{\gamma \delta }$ for all $\gamma ,\delta \in \Gamma .$ The elements of the set $\bigcup _{\gamma \in \Gamma } R_\gamma $ are said to be homogeneous and $\gamma $ is the degree of any nonzero element of $R_\gamma .$ The grading is trivial if $R_\gamma =0$ for every $\gamma \in \Gamma $ which is not the group identity. We adopt the standard definitions of graded ring homomorphisms, graded algebras and graded ideals as defined in [Reference Hazrat4].
A ring R is an involutive ring, or a $*$ -ring, if there is an antiautomorphism $*:R\to R$ of order two. If R is also a K-algebra for some commutative, involutive ring K, then R is a $*$ -algebra if $(kx)^*=k^*x^*$ for all $k\in K$ and $x\in R.$ If R and S are $*$ -rings, a ring homomorphism $\phi : R\to S$ is a $*$ -homomorphism if $\phi (x^*)=\phi (x)^*$ for every $x\in R.$
2.2. Leavitt path algebras
Let E be a directed graph. Let $E^0$ denote the set of vertices, let $E^1$ denote the set of edges and let $\mathbf {s}$ and $\mathbf {r}$ be the source and range maps of E, respectively. A sink of E is a vertex which emits no edges, and an infinite emitter is a vertex which emits infinitely many edges. A vertex of E is regular if it is neither a sink nor an infinite emitter. A path is a single vertex or a sequence of edges $e_1e_2\ldots e_n$ for some positive integer n such that $\mathbf {r}(e_i)=\mathbf {s}(e_{i+1})$ for $i=1,\ldots , n-1.$
Extend a graph E to the graph which has the same vertices and whose edges are $E^1\cup \{e^*\mid e\in E^1\}$ , where the range and source functions are the same as in E for $e\in E^1$ and $\mathbf {s}(e^*)=\mathbf {r}(e)$ and $\mathbf {r}(e^*)=\mathbf {s}(e)$ for the added edges. If K is any field, the Leavitt path algebra $L_K(E)$ of E over K is a free K-algebra generated by the set ${E^0\cup E^1\cup \{e^\ast \mid e\in E^1\}}$ such that, for all vertices $v,w$ and edges $e,f$ :
-
(V) $vw =0$ if $v\neq w$ and $vv=v$ ;
-
(E1) $\mathbf {s}(e)e=e\mathbf {r}(e)=e$ ;
-
(E2) $\mathbf {r}(e)e^\ast =e^\ast \mathbf {s}(e)=e^\ast $ ;
-
(CK1) $e^\ast f=0$ if $e\neq f$ and $e^\ast e=\mathbf {r}(e)$ ; and
-
(CK2) $v=\sum _{e\in \mathbf {s}^{-1}(v)} ee^\ast $ for each regular vertex v.
By the first four axioms, every element of $L_K(E)$ can be represented as a sum of the form $\sum _{i=1}^n k_ip_iq_i^\ast $ for some n, paths $p_i$ and $q_i$ and elements $k_i\in K,$ for $i=1,\ldots ,n$ , where $v^*=v$ for $v\in E^0$ and $p^*=e_n^*\ldots e_1^*$ for a path $p=e_1\ldots e_n.$ Using this representation, one can make $L_K(E)$ into an involutive ring by defining $(\sum _{i=1}^n k_ip_iq_i^\ast )^*=\sum _{i=1}^n k_i^*q_ip_i^\ast $ , where $k_i\mapsto k_i^*$ is any involution on K. For more details on these basic properties, see [Reference Abrams, Ara and Siles Molina1].
If we consider K to be trivially graded by $\mathbb Z, L_K(E)$ is naturally graded by $\mathbb Z$ so that the n-component $L_K(E)_n$ is the K-linear span of the elements $pq^\ast $ for paths $p, q$ with $|p|-|q|=n$ , where $|p|$ denotes the length of a path $p.$ Although one can grade a Leavitt path algebra by any group (see [Reference Hazrat4, Section 1.6.1]), we consider only the natural grading by $\mathbb Z.$
If a K-algebra R contains elements $a_v, b_e$ and $c_{e^*}$ which satisfy the axioms (V), (E1), (E2), (CK1) and (CK2), where $v\in E^0$ and $e\in E^1$ for some graph E, the Universal Property of $L_K(E)$ ensures that the map $\phi : v\mapsto a_v, e\mapsto b_e, e^*\mapsto c_{e^*}$ has a unique K-algebra homomorphism extension $\phi : L_K(E)\to R$ (see [Reference Abrams, Ara and Siles Molina1, Remark 1.2.5]). If R is $\mathbb Z$ -graded and the elements $\phi (v), \phi (e)$ and $\phi (e^*)$ have degrees 0, 1 and $-1$ , respectively, then such an extension is graded and the Graded Uniqueness Theorem [Reference Abrams, Ara and Siles Molina1, Theorem 2.2.15] states that $\phi $ is injective if and only if $\phi (v)\neq 0$ for every $v\in E^0.$ If R is involutive and if $a_v^*=a_v$ for every $v\in E^0$ and $(b_e)^*=c_{e^*}$ for every $e\in E^1,$ then $\phi $ is a $*$ -homomorphism (see [Reference Vaš7, Lemma 4.7]).
2.3. Graded ideals of a Leavitt path algebra
A subset H of $E^0$ is said to be hereditary if $\mathbf {r}(p)\in H$ for any path p such that $\mathbf {s}(p)\in H.$ The set H is saturated if $v\in H$ for any regular vertex v such that $\mathbf {r}(\mathbf {s}^{-1}(v))\subseteq H.$ We recall a construction from [Reference Tomforde6]. If H is hereditary and saturated, let
An admissible pair is an ordered pair $(H, S)$ , where $H\subseteq E^0$ is hereditary and saturated and $S\subseteq B_H.$ For such a pair, let $I(H,S)$ denote the graded ideal generated by homogeneous elements $H\cup \{v^H \,|\, v\in S \}.$ The ideal $I(H,S)$ is the K-linear span of the elements $pq^*$ for paths $p,q$ with $\mathbf {r}(p)=\mathbf {r}(q)\in H$ and the elements $pv^Hq^*$ for paths $p,q$ with $\mathbf {r}(p)=\mathbf {r}(q)=v\in S$ (see [Reference Tomforde6, Lemma 5.6]). The converse holds as well: for every graded ideal I, the vertices in I form a hereditary and saturated set H and the set of infinite emitters such that $v^H\in I$ is a subset of $B_H$ (see [Reference Tomforde6, Theorem 5.7] and also [Reference Abrams, Ara and Siles Molina1, Theorem 2.5.8]). For an admissible pair $(H, S),$ let $E_{(H,S)}$ be the graph defined by
with $\mathbf {s}$ and $\mathbf {r}$ the same as on $E^1$ for $e\in E^1\cap E^1_{(H,S)}$ and $\mathbf {s}(\overline p)=p, \mathbf {r}(\overline p)=\mathbf {r}(p)$ for $\overline p\in \overline {F_1}(H,S)\cup \overline {F_2}(H,S)$ . By [Reference Ruiz and Tomforde5, Theorem 6.1] (also [Reference Abrams, Ara and Siles Molina1, Theorem 2.5.22]), the ideal $I(H,S)$ is isomorphic to $L_K (E_{(H,S)}).$
In [Reference Abrams, Ara and Siles Molina1], the graph $E_{(H,\emptyset )}$ is referred to as the hedgehog graph and the graph $E_{(H,S)}$ is referred to as the generalised hedgehog graph. The term ‘hedgehog’ is chosen since the added edges ending at $H\cup S$ resemble spines of the hedgehog body $H\cup S.$ The isomorphism $L_K(E_{(H,S)})\to I(H,S)$ is defined so that the edges of $E_{(H,S)}$ are mapped to certain homogeneous elements of $I(H,S)$ of positive degree, not necessarily degree one. Thus, the degrees are not necessarily preserved under this map and so this isomorphism may not be graded.
In [Reference Aranda Pino, Rangaswamy and Siles Molina2], the authors consider a version of $E_{(H,S)}$ , which they denote $_HE_S$ (see [Reference Ruiz and Tomforde5] for details on differences in definitions); [Reference Aranda Pino, Rangaswamy and Siles Molina2, Proposition 3.7] states that $I(H,S)$ and $L_K(_HE_S)$ are graded isomorphic. However, the map $L_K(_HE_S)\to I(H,S)$ in the proof is defined analogously to the map $L_K(E_{(H,S)})\to I(H,S)$ and, as we noted above, this map is not necessarily a graded map. The last section of [Reference Ruiz and Tomforde5] contains some further details on a problem with the definition of $_HE_S$ and [Reference Aranda Pino, Rangaswamy and Siles Molina2, Proposition 3.7].
3. The main result
We modify the construction of the generalised hedgehog graph of an admissible pair $(H,S)$ by making the spines added to the body $H\cup S$ possibly fewer in number but then necessarily longer in length. Because of the longer spines, we call the resulting graph the porcupine graph and denote it by $P_{(H,S)}.$ The modified construction enables us to create a degree-preserving isomorphism $L_K(P_{(H,S)})\to I(H,S).$
Definition 3.1. For an admissible pair $(H,S),$ we keep the definitions of $F_1(H,S)$ and $F_2(H,S).$ For each $e\in (F_1(H,S)\cup F_2(H,S))\cap E^1,$ let $w^e$ be a new vertex and let $f^e$ be a new edge such that $\mathbf {s}(f^e)=w^e$ and $\mathbf {r}(f^e)=\mathbf {r}(e).$ Continue this process inductively as follows. For each path $p=eq$ , where $q\in F_1(H,S)\cup F_2(H,S)$ and $|q|\geq 1,$ add a new vertex $w^p$ and a new edge $f^p$ such that $\mathbf {s}(f^p)=w^p$ and $\mathbf {r}(f^p)=w^q.$
We define the porcupine graph $P_{(H,S)}$ as follows. The set of vertices of $P_{(H,S)}$ is
The set of edges of $P_{(H,S)}$ is
The $\mathbf {s}$ and $\mathbf {r}$ maps of $P_{(H,S)}$ are the same as in E for the common edges and they are defined as above for the new edges.
Before formulating and proving the main result, we present some examples that compare and contrast the generalised hedgehog and the porcupine graphs of an admissible pair.
Example 3.2. Let E be the Toeplitz graph
and let $H=\{v\}.$ The hedgehog graph of $(H, \emptyset )$ is shown first and the porcupine graph is second.
The graded isomorphism of Theorem 3.3 is such that the path $e^{n-1}g$ of length n corresponds to the path $f^{e^{n-1}g}f^{e^{n-2}g}\ldots f^g$ , which is also of length n.
We consider another example with an infinite emitter. Let E be the graph
and consider $H=\{v\}, S=\{w\}$ . In this case, $F_1(H,S)=\{e_3, e_2e_3, e_1e_2e_3\} $ and $F_2(H,S)=\{e_1\}.$ The generalised hedgehog graph of $(H,S)$ is shown first and the porcupine graph is second.
Theorem 3.3. For an admissible pair $(H,S)$ of a graph $E, I(H,S)$ and $L_K (P_{(H,S)})$ are graded $*$ -isomorphic. Thus, every graded ideal of a Leavitt path algebra is graded $*$ -isomorphic to a Leavitt path algebra.
Proof. We define a function $\phi $ which maps the vertices and edges of $P_{(H,S)}$ to elements of $ I(H,S)$ , as follows. For a vertex v of $P_{(H,S)},$ we let
For an edge g of $P_{(H,S)},$ we let
Extend $\phi $ to $g^*$ for $g\in P_{(H,S)}^1$ by $\phi (g^*)=\phi (g)^*$ and use the definitions to check that (V), (E1), (E2) and (CK1) hold for $\phi (v), \phi (g), \phi (g^*), v\in P_{(H,S)}^0, g\in P_{(H,S)}^1.$ We present more details for checking that (CK2) holds. If v is a regular vertex of $P_{(H,S)},$ then v cannot be in $S,$ so $v\in H$ or $v=w^p$ for some $p\in F_1(H,S)\cup F_2(H,S).$ In the first case, the edges v emits are in E since H is hereditary and the relation $v=\sum _{e\in \mathbf {s}^{-1}(v)} ee^*$ holds in $L_K(P_{(H,S)})$ since it holds in $L_K(E).$ In the second case, consider the four possible cases: $p=e\in F_1(H,S)\cap E^1, p=e\in F_2(H,S)\cap E^1, p=eq$ for $e\in E^1, q\in F_1(H,S),$ and $p=eq$ for $e\in E^1, q\in F_2(H,S).$ Note that, in each case, v emits only one edge $f^p$ so it is sufficient to check that $\phi (f^p)\phi ((f^p)^*)=\phi (v)$ in each case.
-
(i) If $p=e\in F_1(H,S)\cap E^1,$ then $\phi (f^e)\phi ((f^e)^*)=ee^*=\phi (v).$
-
(ii) If $p=e\in F_2(H,S)\cap E^1,$ then $\phi (f^e)\phi ((f^e)^*)=e\mathbf {r}(e)^H \mathbf {r}(e)^H e^*=e\mathbf {r}(e)^H e^*=\phi (v).$
-
(iii) If $p=eq$ for $e\in E^1, q\in F_1(H,S),$ then
$$ \begin{align*}\phi(f^p)\phi((f^p)^*)=eqq^*qq^*e^*=eqq^*e^*=pp^*=\phi(v).\end{align*} $$ -
(iv) If $p=eq$ for $e\in E^1, q\in F_2(H,S),$ then
$$ \begin{align*}\phi(f^p)\phi((f^p)^*)=eq\mathbf{r}(p)^Hq^*q\mathbf{r}(p)^Hq^*e^*=eq\mathbf{r}(p)^Hq^*e^*=p\mathbf{r}(p)^Hp^*=\phi(v).\end{align*} $$
This shows that (CK2) holds.
By the Universal Property, $\phi $ has a unique extension to a K-algebra homomorphism $L_K(P_{(H,S)})\to I(H,S).$ Since $\phi $ preserves degrees of $v\in P_{(H,S)}^0$ and of g and $g^*$ for $g\in P_{(H,S)}^1,$ this extension, which we denote also by $\phi ,$ is a graded homomorphism. As $\phi (v)^*=\phi (v)$ for $v\in P_{(H,S)}^0$ and $\phi (g^*)=\phi (g)^*$ for $g\in P_{(H,S)}^1, \phi $ is a $*$ -homomorphism. The map $\phi $ is nonzero on every vertex of $P_{(H,S)},$ so $\phi $ is injective by the Graded Uniqueness Theorem.
Since $\phi $ is a $*$ -homomorphism, to show surjectivity of $\phi ,$ it is sufficient to show (1) and (2), where (1) denotes the condition that p is in the image of $\phi $ for every path such that $\mathbf {r}(p)\in H,$ and (2) denotes the condition that $p\mathbf {r}(p)^H$ is in the image of $\phi $ for every path p such that $\mathbf {r}(p)\in S.$ Both claims hold for paths of zero length since $\phi (v)=v$ if $v\in H$ and $\phi (v)=v^H$ if $v\in S.$ Thus, consider a path $p=e_1\ldots e_n$ for a positive integer $n.$
To show (1), assume that $\mathbf {r}(p)\in H.$ If $\mathbf {s}(p)\in H,$ then $\phi (p)=p$ so the claim holds. If $\mathbf {s}(p)\notin H,$ let $i\in \{1,2,\ldots , n\}$ be the largest such that $\mathbf {s}(e_i)\notin H.$ We consider three cases: (i) $\mathbf {s}(e_i)\notin S,$ (ii) $\mathbf {s}(e_i)\in S$ and $i=1$ and (iii) $\mathbf {s}(e_i)\in S$ and $i>1.$
In case (i), $e_j\ldots e_i\in F_1(H,S)$ for all $j=1,\ldots i$ and
if $i<n$ , and the analogous argument applies to the case $i=n.$
In case (ii), $\phi (e_j)=e_j$ for every $j=1,\ldots , n$ and $\phi (p)=p.$
In case (iii), $e_j\ldots e_{i-1}\in F_2(H,S)$ for every $j=1,\ldots i-1$ and $\mathbf {r}(e_{i-1})^He_i=e_i$ , so
This shows that (1) holds.
To show (2), assume that $\mathbf {r}(p)\in S.$ Then $e_i\ldots e_n$ is in $F_2(H,S)$ for every $i=1,\ldots , n$ and
This shows that $\phi $ is surjective.
The second sentence of the theorem is a direct corollary of the first sentence and [Reference Tomforde6, Theorem 5.7] (also [Reference Abrams, Ara and Siles Molina1, Theorem 2.5.8]).
3.1. Graph ${\textit{C}}^{\boldsymbol{*}}$ -algebras
Theorem 3.3 has its graph $C^*$ -algebra version. If E is a graph, then the graph $C^*$ -algebra of E is the universal $C^*$ -algebra generated by mutually orthogonal projections $\{p_v\mid v\in E^0\}$ and partial isometries with mutually orthogonal ranges $\{s_e\mid e\in E^1\}$ satisfying the analogues of the (CK1) and (CK2) axioms and the axiom (CK3) stating that $s_es_e^*\leq p_{\mathbf {s}(e)}$ for every $e\in E^1.$ The term ‘universal’ in the definition means that the $C^*$ -algebra version of the algebraic Universal Property, mentioned before, holds (see [Reference Abrams, Ara and Siles Molina1, Definition 5.2.5]). By letting $s_{e_1\ldots e_n}$ be $s_{e_1}\ldots s_{e_n}$ and $s_v=p_v$ for $e_1,\ldots ,e_n\in E^1$ and $v\in E^0, s_p$ is defined for every path $p.$
The set $\{p_v, s_e\mid v\in E^0, e\in E^1\}$ is referred to as a Cuntz–Krieger E-family. For such an E-family and an element z of the unit sphere $\mathbb T$ , one defines a map $\gamma ^E_z$ by $\gamma ^E_z(p_v)=p_v$ and $\gamma ^E_z(s_e)=zs_e$ and then uniquely extends this map to a $*$ -automorphism of $C^*(E)$ (we assume a homomorphism of a $C^*$ -algebra to be bounded). The gauge action $\gamma ^E$ on $\mathbb T$ is given by $\gamma ^E(z)=\gamma ^E_z.$ Note that $\gamma ^E_z(s_ps_q^*)=z^{|p|-|q|}s_ps_q^*$ for $z\in \mathbb T$ and paths p and $q.$ The presence of the degree $|p|-|q|$ of $pq^*$ in the previous formula explains the connection between this action and the $\mathbb Z$ -grading of $L_{\mathbb C}(E)$ . If the integral of a continuous function $f:\mathbb T\to \mathbb C$ over $\mathbb T$ is defined by $\int _{\mathbb T}f(z)\,dz=\int _0^1 f(e^{2\pi i t})\,dt,$ then the gauge action on $C^*(E)$ determines a $\mathbb Z$ -grading of $C^*(E)$ (see [Reference Abrams, Ara and Siles Molina1, Proposition 5.2.11]) so that
is the completion of $L_{\mathbb C}(E)_n$ and $C^*(E)$ is the completion of $\bigoplus _{n\,\in\,\mathbb Z} C^*(E)_n.$ We note that this grading is not a grading in the algebraic sense, but in the $C^*$ -algebra sense (see the paragraph following [Reference Abrams, Ara and Siles Molina1, Theorem 5.2.9.]).
A closed ideal J of a graph $C^*$ -algebra $C^*(E)$ is gauge-invariant if $\gamma ^E_z(J)=J$ for every $z\in \mathbb T.$ By [Reference Bates, Hong, Raeburn and Szymański3, Theorem 3.6], every such ideal J is the completion of the linear span of the elements $s_ps_q^*$ for paths $p,q$ with $\mathbf {r}(p)=\mathbf {r}(q)\in H$ and the elements $s_pp_v^Hs_q^*$ for paths $p,q$ with $\mathbf {r}(p)=\mathbf {r}(q)=v\in S$ , where $p_v^H=p_v-\sum _{e\in \mathbf {s}^{-1}(v)\cap \mathbf {r}^{-1}(E^0-H)}s_es_e^*$ for $v\in B_H$ , and where $(H, S)$ is the admissible pair uniquely determined by J and defined analogously as for an ideal of $L_K(E)$ . Also as before, an admissible pair $(H,S)$ uniquely determines a closed gauge-invariant ideal $J(H,S)$ which is graded in the $C^*$ -algebra sense.
If R is a $C^*$ -algebra with an action $\beta : \mathbb T\to $ Aut $(R),$ then we say that a $*$ -homomorphism $\phi : C^*(E)\to R$ is gauge-invariant if $\beta _z\circ \phi =\phi \circ \gamma ^E_z$ for every $z\in \mathbb T.$ Thus, if J is a closed gauge-invariant ideal of $C^*(E), F$ is a graph and $\phi : C^*(F)\to J$ is a $*$ -homomorphism, then $\phi $ is gauge-invariant if $\gamma _z^E|_J\circ \phi =\phi \circ \gamma ^F_z$ for every $z\in \mathbb T.$ It is direct to see that every such gauge-invariant map $\phi $ is graded. Indeed, using the formula (*) and the fact that $\phi $ is bounded, one directly checks that $\phi (C^*(F)_n)\subseteq J_n.$
Corollary 3.4. For an admissible pair $(H,S)$ of a graph $E,$ there is a gauge-invariant (thus graded) $*$ -isomorphism
Thus, for every closed gauge-invariant ideal J of a graph $C^*$ -algebra, there is a gauge-invariant (and thus graded) $*$ -isomorphism mapping the graph $C^*$ -algebra of the porcupine graph corresponding to J onto $J.$
Proof. One defines a map $\phi $ on $\{p_v, s_g\mid v\in P_{(H,S)}^0, g\in P_{(H,S)}^1\}$ analogously as in Theorem 3.3. We claim that the image of this map constitutes a Cuntz–Krieger $P_{(H,S)}$ -family. Indeed, $\{\phi (p_v)\mid v\in P_{(H,S)}^0\}$ is a set of orthogonal projections in $J(H,S)$ and $\{\phi (s_g)\mid g\in P_{(H,S)}^1\}$ is a set of partial isometries in $I(H,S)$ with mutually orthogonal ranges. One checks that (CK1) and (CK2) hold in the same way as in the proof of Theorem 3.3. If $g\in P_{(H,S)}^1$ and $\mathbf {s}(g)$ is regular, then the requirement $\phi (s_g)\phi (s_g)^*\leq \phi (p_{\mathbf {s}(g)})$ follows from (CK2). If $g\in P_{(H,S)}^1$ and $\mathbf {s}(g)$ is an infinite emitter, then the $\mathbf {s}(g)\in H\cup S$ and g is an edge in $E^1.$ Thus, (CK3) holds for $s_g$ in $C^*(P_{(H,S)})$ since it holds for $s_g$ in $C^*(E).$
The map $\phi $ has a unique extension to a $*$ -homomorphism $\phi :C^*(P_{(H,S)})\to J(H,S)$ by the Universal Property. To show that $\phi $ is gauge-invariant, it is sufficient to check that the condition $\gamma _z^E|_I\circ \phi =\phi \circ \gamma ^{P_{(H,S)}}_z$ holds on $\{p_v, s_g\mid v\in P_{(H,S)}^0, g\in P_{(H,S)}^1\}$ , which is done directly using the definitions.
The map $\phi $ is nonzero on every element $p_v$ for $v\in P_{(H,S)}^0,$ so $\phi $ is injective by the Gauge-Invariant Uniqueness Theorem (see [Reference Abrams, Ara and Siles Molina1, Theorem 5.2.12]). The surjectivity of $\phi $ holds by arguments analogous to those in the proof of Theorem 3.3.
Acknowledgement
The author is very grateful to the referee for the suggestion to consider formulating and proving the $C^*$ -algebra version of the main result (now Corollary 3.4).