This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Gauge-Invariant Uniqueness and
Reductions of Ordered Groups

Robert Huben
(March 2021)
Abstract

A reduction φ\varphi of an ordered group (G,P)(G,P) to another ordered group is an order homomorphism which maps each interval [1,p][1,p] bijectively onto [1,φ(p)][1,\varphi(p)]. We show that if (G,P)(G,P) is weakly quasi-lattice ordered and reduces to an amenable ordered group, then there is a gauge-invariant uniqueness theorem for PP-graph algebras. We also consider the class of ordered groups which reduce to an amenable ordered group, and show this class contains all amenable ordered groups and is closed under direct products, free products, and hereditary subgroups.

1 Introduction

Cuntz-Krieger algebras and their generalizations (Exel-Laca algebras, graph algebras, higher-rank graph algebras, etc) are all, broadly speaking, the universal algebras generated by partial isometries whose range and source projections satisfy certain combinatorial relations. But defining these algebras by their universality comes at a cost, since it becomes difficult to check if any particular collection of partial isometries (a “representation”) is universal. Mathematicians responded with uniqueness theorems, conditions on the representation which guarantee that it is universal. A classic example is the gauge-invariant uniqueness theorem for graph algebras, which states that so long as the canonical generators of the algebra are nonzero and there is a gauge action (meaning an action α\alpha of the circle satisfying αz(Tγ)=zlength(γ)Tγ\alpha_{z}(T_{\gamma})=z^{\text{length}(\gamma)}T_{\gamma} for any z𝕋z\in\mathbb{T} and path γ\gamma in the graph), then any other representation of the graph is a quotient of this representation. This construction and uniqueness theorem has been generalized to higher rank graphs by Kumjian and Pask [15], where paths are given lengths in k\mathbb{N}^{k} instead of \mathbb{N}, and an action of 𝕋k=k^\mathbb{T}^{k}=\widehat{\mathbb{Z}^{k}} replaces the gauge action of 𝕋=^\mathbb{T}=\widehat{\mathbb{Z}}.

One might then ask that paths be allowed to have “lengths” in any positive cone PP of a group GG. The first attempt at such a generalization was by Brownlowe, Sims, and Vittadello [3], who studied PP-graphs when (G,P)(G,P) is quasi-lattice ordered, but surprisingly the authors construct a PP-graph algebra which is not universal but co-universal, meaning it is a quotient of any sufficiently large representation.

In Section 4 of this paper, we build on their work to show that if the grading group (G,P)(G,P) is weakly quasi-lattice ordered and has a certain kind of quotient map called a reduction such that the quotient is amenable, this algebra is both universal and co-universal. This allows us to generalize the gauge-invariant uniqueness theorems for graphs and kk-graphs (such as [18, Theorem 2.2] or [15, Theorem 3.4]) to make a gauge-invariant uniqueness theorem for PP-graphs:

{restatable*}

theoremuniquenesstheorem Let (G,P)(G,P) be a weakly quasi-lattice ordered group which reduces to an amenable ordered group, and let Λ\Lambda be a finitely-aligned PP-graph. Then there is exactly one representation (up to canonical isomorphism) of Λ\Lambda which is Λ\Lambda-faithful, tight, and has a gauge coaction of GG. This representation is universal for tight representations and co-universal for representations which are Λ\Lambda-faithful and have a gauge coaction by GG.

This diagram provides a visual summary of the result:

Λ\Lambda-faithful gauge coactingrepresentationstight representationsCmin(Λ)C^{*}_{min}(\Lambda)Ctight(Λ)C^{*}_{tight}(\Lambda)Cmin(Λ)C^{*}_{min}(\Lambda)no gap if (G,P)(G,P) reduces to an amenable groupbigger representationssmaller representations

We believe that the algebra generated by this simultaneously universal and co-universal representation deserves the title of the Cuntz-Krieger algebra of the PP-graph.

The notion of a reduction of ordered groups is introduced in Section 3, culminating in this theorem showing that several natural operations on ordered groups preserve the existence of a reduction onto an amenable group:

{restatable*}

theoremAmenableReductionTheorem The class of ordered groups which have (strong) reductions onto amenable groups contains all amenable ordered groups and is closed under hereditary subgroups, finite direct products, and finite free products.

In Theorem 1, being reducible to an amenable group plays the role of amenability in guaranteeing a unique representation, but by Theorem 1, this condition is, in at least one sense, more robust than amenability. In particular, being reducible to an amenable group is preserved under free products, whereas being amenable is almost always destroyed by a free product (for instance if GG and HH contain copies of \mathbb{Z}, then GHG*H is not amenable).

The notion of being reducible to an amenable group is sufficiently robust that, as corollary to Theorem 1, the important example (2,2)(\mathbb{Z}^{2}*\mathbb{Z},\mathbb{N}^{2}*\mathbb{N}) from the literature ([21, 3]) reduces to an amenable group. In Proposition 4.43, we use this fact in combination with the results of [3] to show that every Kirchberg algebra in the UCT class is stably isomorphic to the simultaneously universal and co-universal algebra of a (2)(\mathbb{N}^{2}*\mathbb{N})-graph.

Besides the results listed in Theorem 1, a great deal is not yet known about how reductions interact with other group constructions. The limits of our understanding are discussed in Section 3.3.

We hope that this new approach will provide a useful tool for the analysis of PP-graphs and their algebras, and possibly have implications for product systems and Fell bundles.

2 Background

2.1 Ordered Groups

Ordered groups have been studied in both a context of group theory and operator theory, and we largely follow the conventions of [3] and [10, Chapter 32]. The reader should note that other authors, such as [4], use the term “ordered group” to refer to a group with a total order, but we follow the convention that the group has a partial order.

Definition 2.1.

A positive cone in a group GG is a submonoid PP such that PP1={1}P\cap P^{-1}=\{1\}.

A left-invariant partial order on a group GG is a partial order \leq such that for all a,b,cGa,b,c\in G, aba\leq b if and only if cacbca\leq cb.

Every left-invariant partial order on a group arises naturally from a positive cone PP by saying that aba\leq b if and only if a1bPa^{-1}b\in P. We will say that an ordered group (G,P)(G,P) is a group along with a positive cone, with the implication that the group takes on a left-invariant partial order in this way.

A group is totally ordered if its partial order is a total order. This is equivalent to the condition that PP1=GP\cup P^{-1}=G.

Although our theory of reductions does not require any additional structure on our ordered groups, the relators for a graph algebra will ask for an additional property on our orderings from [10, Chapter 32]:

Definition 2.2.

We say (G,P)(G,P) is weakly quasi-lattice ordered (WQLO) if whenever x,yPx,y\in P have a common upper bound in PP, they have a supremum (least common upper bound), denoted xyx\vee y. Note that any upper bound of a positive element is itself positive by transitivity.

While many mathematicians (including[3, 20, 22]) worked in the context of quasi-lattice ordered groups, we have found that the slightly more general condition of weak quasi-lattice order suffices. The condition of weak quasi-lattice order is also friendlier than quasi-lattice order since it is entirely a property of the submonoid PP, and therefore one can “forget” the ambient group. The same is not true for quasi-lattice order. For a more thorough introduction to quasi-lattice order, weak quasi-lattice order, and their connections, the reader is directed to [10, Chapter 32].

Notation 2.3.

In this work, we will make use of the following notation:

  • ={0,1,2,3}\mathbb{N}=\{0,1,2,3...\} includes 0.

  • Given a group GG and subset SGS\subseteq G, S\langle S\rangle will denote the group generated by the elements of SS, and SS^{*} will denote the monoid generated by the elements of SS.

  • The identity element in a multiplicative group GG will be denoted by 1G1_{G}, or 11 if the group is clear from context. Less commonly we will denote the identity element by ee.

Example 2.4.

Let G=kG=\mathbb{Z}^{k}, and let P=kP=\mathbb{N}^{k}. Then (G,P)(G,P) is WQLO. Every pair of elements p=(p1,,pk)p=(p_{1},...,p_{k}) and q=(q1,,qk)q=(q_{1},...,q_{k}) have a supremum pq=(max(p1,q1),,max(pk,qk))p\vee q=(\max(p_{1},q_{1}),...,\max(p_{k},q_{k})). (In fact, this ordering is a “lattice order”, meaning any two elements have both a supremum and an infimum.)

Let G=Fk=a1,,akG=F_{k}=\langle a_{1},...,a_{k}\rangle be the free group on kk generators, and let P={a1,,ak}P=\{a_{1},...,a_{k}\}^{*} be the free monoid generated by the generators of GG. Then (G,P)(G,P) is WQLO, since any p,qPp,q\in P have a common upper bound if and only if pqp\leq q or qpq\leq p, in which case pq=max{p,q}p\vee q=\max\{p,q\}.

Notation 2.5.

Let (G,P)(G,P) be an ordered group. We use the following standard notation for intervals: for a,bGa,b\in G, write

[a,b]:={gG|agb}.[a,b]:=\{g\in G|a\leq g\leq b\}.

It is immediate that such an interval is nonempty if and only if aba\leq b. We will most often be interested in intervals of the form [1,p][1,p] for some pPp\in P.

Definition 2.6.

Let (G,P)(G,P) and (H,Q)(H,Q) be ordered groups. We say that φ:GH\varphi:G\rightarrow H is an order homomorphism if it is a group homomorphism with φ(P)Q\varphi(P)\subseteq Q. We will write φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) to denote an order homomorphism.

Order homomorphisms preserve both the group structure and the order structure, as the following lemma shows:

Lemma 2.7.

Let (G,P)(G,P) and (H,Q)(H,Q) be ordered groups, and φ:GH\varphi:G\rightarrow H a group homomorphism. Then φ\varphi is an order homomorphism if and only if xyx\leq y implies φ(x)φ(y)\varphi(x)\leq\varphi(y) for all x,yGx,y\in G.

In particular, if φ\varphi is an order homomorphism and pPp\in P, then φ([1,p])[1,φ(p)]\varphi([1,p])\subseteq[1,\varphi(p)].

Proof.

If φ\varphi is an order homomorphism and xyx\leq y, then x1yPx^{-1}y\in P, so φ(x1y)=φ(x)1φ(y)Q\varphi(x^{-1}y)=\varphi(x)^{-1}\varphi(y)\in Q. Thus φ(x)φ(y)\varphi(x)\leq\varphi(y).

Conversely, if φ\varphi is not an order homomorphism, then there is pPp\in P such that φ(p)Q\varphi(p)\not\in Q. Then 1Gp1_{G}\leq p, and 1H=φ(1G)φ(p)1_{H}=\varphi(1_{G})\not\leq\varphi(p), as desired.

For the “in particular”, if s[1,p]s\in[1,p], then 1sp1\leq s\leq p, so 1=φ(1)φ(s)φ(p)1=\varphi(1)\leq\varphi(s)\leq\varphi(p), so φ(s)[1,φ(p)]\varphi(s)\in[1,\varphi(p)] as desired. ∎

2.2 CC^{*}-algebras

2.2.1 The Factors Through Theorem

The following result is an elementary notion from algebra, which we state here for completeness since we will make frequent use of it:

Lemma 2.8 (Factors Through Theorem).

Let A,BA,B, and CC be sets, let f:ABf:A\rightarrow B and g:ACg:A\rightarrow C be functions, and suppose that ff is surjective.

A{A}B{B}C{C}f\scriptstyle{f}g\scriptstyle{g}h\scriptstyle{\exists h}
  1. 1.

    Suppose that whenever f(a)=f(a)f(a)=f(a^{\prime}), then g(a)=g(a)g(a)=g(a^{\prime}). Then there is a “bonding map” h:BCh:B\rightarrow C such that hf=gh\circ f=g. In particular, hh is (well-)defined by h(f(a))=g(a)h(f(a))=g(a) for all f(a)Bf(a)\in B.

  2. 2.

    If A,B,A,B, and CC are groups, ff and gg are group homomorphisms, and kerfkerg\ker f\subseteq\ker g, then hypothesis of (1) that f(a)=f(a)f(a)=f(a^{\prime}) implies g(a)=g(a)g(a)=g(a^{\prime}) is satisfied. Therefore, the conclusion of (1) is true.

  3. 3.

    If A,B,CA,B,C have a binary operation \cdot (respectively a unary operation ) which is preserved by ff and gg, then \cdot (respectively ) is also preserved by hh.

Proof.

For (1), for all bBb\in B, since ff is surjective, there is some aAa\in A such that f(a)=bf(a)=b. Now, define h(b)=g(a)h(b)=g(a), which is well-defined since if there were some other aAa^{\prime}\in A with f(a)=bf(a^{\prime})=b, then g(a)=g(a)g(a)=g(a^{\prime}) by hypothesis. That is, we have defined hh by h(f(a))=g(a)h(f(a))=g(a), so hf=gh\circ f=g.

For (2), if f(a)=f(a)f(a)=f(a^{\prime}), then a1akerfkerga^{-1}a^{\prime}\in\ker f\subseteq\ker g, so g(a)=g(a)g(a)=g(a^{\prime}), as desired.

For (3), let b1,b2Bb_{1},b_{2}\in B. Since ff is surjective, there is some a1,a2Aa_{1},a_{2}\in A such that f(a1)=b1,f(a2)=b2f(a_{1})=b_{1},f(a_{2})=b_{2}. Then

h(b1b2)\displaystyle h(b_{1}\cdot b_{2}) =\displaystyle= h(f(a1)f(a2))=h(f(a1a2))=g(a1a2)\displaystyle h(f(a_{1})\cdot f(a_{2}))=h(f(a_{1}\cdot a_{2}))=g(a_{1}\cdot a_{2})
=\displaystyle= g(a1)g(a2)=h(f(a1))h(f(a2))=h(b1)h(b2)\displaystyle g(a_{1})\cdot g(a_{2})=h(f(a_{1}))\cdot h(f(a_{2}))=h(b_{1})\cdot h(b_{2})

and

h(b1)=h(f(a1))=h(f(a1))=g(a1)=g(a1)=h(f(a1))=h(b1)h(b_{1}^{*})=h(f(a_{1})^{*})=h(f(a_{1}^{*}))=g(a_{1}^{*})=g(a_{1})^{*}=h(f(a_{1}))^{*}=h(b_{1})^{*}

as desired.

2.2.2 Universal Algebras

The following construction of a “universal algebra” is a slight simplification of the one outlined by Blackadar in [2]. We state it here for completeness.

Lemma 2.9 (Construction of a Universal Algebra, [2]).

Let 𝒢={xα}αI\mathcal{G}=\{x_{\alpha}\}_{\alpha\in I} denote a set of formal symbols, 𝒢={xα}αI\mathcal{G}^{*}=\{x_{\alpha}^{*}\}_{\alpha\in I} another set of formal symbols over the same indexing set, and \mathcal{R} a set of relators of the form p(xα1,,xαn,xα1,,xαn)=0p(x_{\alpha_{1}},...,x_{\alpha_{n}},x_{\alpha_{1}}^{*},...,x_{\alpha_{n}}^{*})=0 where pp is some polynomial in 2n2n non-commuting variables and complex coefficients.

We define a representation of (𝒢,)(\mathcal{G},\mathcal{R}) to be a collection of elements {yα}αI\{y_{\alpha}\}_{\alpha\in I} in a CC^{*}-algebra that satisfy p(yα1,,yαn,yα1,,yαn)=0p(y_{\alpha_{1}},...,y_{\alpha_{n}},y_{\alpha_{1}}^{*},...,y_{\alpha_{n}}^{*})=0 for each relator in \mathcal{R}.

Suppose that the relators on (𝒢,)(\mathcal{G},\mathcal{R}) imply that in any representation, the images of the generators are partial isometries. Then there is a representation {zα}αI\{z_{\alpha}\}_{\alpha\in I} such that for any other representation {yα}αI\{y_{\alpha}\}_{\alpha\in I} there is a surjective *-homomorphism π:C({zα}αI)C({yα}αI)\pi:C^{*}(\{z_{\alpha}\}_{\alpha\in I})\rightarrow C^{*}(\{y_{\alpha}\}_{\alpha\in I}) satisfying zαyαz_{\alpha}\mapsto y_{\alpha}.

The existence of this surjective *-homomorphism is called the universal property of C({zα}αI)C^{*}(\{z_{\alpha}\}_{\alpha\in I}). We call this representation the universal representation and say that C({zα}αI)C^{*}(\{z_{\alpha}\}_{\alpha\in I}) is the universal algebra for (𝒢,)(\mathcal{G},\mathcal{R}).

Proof.

Let (𝒢)\mathcal{F}(\mathcal{G}) denote the free *-algebra over \mathbb{C} generated by \mathbb{C} and 𝒢𝒢\mathcal{G}\cup\mathcal{G}^{*} where the *-map is given by (xα)=xα(x_{\alpha})^{*}=x_{\alpha}^{*} and (xα)=xα(x_{\alpha}^{*})^{*}=x_{\alpha} and extended to the entire algebra via conjugate-linearity and (ab)=ba(ab)^{*}=b^{*}a^{*}. Note that any representation φ:xαyα\varphi:x_{\alpha}\mapsto y_{\alpha} of (𝒢,)(\mathcal{G},\mathcal{R}) extends uniquely to a *-algebra homomorphism φ¯:(𝒢)C({yα}αI)\bar{\varphi}:\mathcal{F}(\mathcal{G})\rightarrow C^{*}(\{y_{\alpha}\}_{\alpha\in I}).

For X(𝒢)X\in\mathcal{F}(\mathcal{G}), let |||X|||=sup{φ¯(X):φ a rep. of (𝒢,)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=\sup\{\left\|{\bar{\varphi}(X)}\right\|:\varphi\text{ a rep. of }(\mathcal{G},\mathcal{R})\}. Note that since (𝒢,)(\mathcal{G},\mathcal{R}) can be represented by the zero map, this supremum is bounded below by 0, and since each generator must be represented by a partial isometry, the image of each generator has norm 1, and thus the supremum is bounded above by the sum of the absolute values of its \mathbb{C}-coefficients. Thus |X|{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} is a well-defined, non-negative finite number.

Let J={X(𝒢):|X|=0}J=\{X\in\mathcal{F}(\mathcal{G}):{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=0\}, and note that J=φ a rep.kerφ¯J=\displaystyle\bigcap_{\varphi\text{ a rep.}}\ker\bar{\varphi} where the intersection is taken over all representations. Thus JJ is the intersection of *-ideals, so it is a *-ideal. Now let ψ:(𝒢)(𝒢)/J\psi:\mathcal{F}(\mathcal{G})\rightarrow\mathcal{F}(\mathcal{G})/J denote the quotient map, and for all α\alpha, let zα=ψ(xα)z_{\alpha}=\psi(x_{\alpha}). For any relator p(xα1,,xαn,xα1,,xαn)=0p(x_{\alpha_{1}},...,x_{\alpha_{n}},x_{\alpha_{1}}^{*},...,x_{\alpha_{n}}^{*})=0\in\mathcal{R}, since this relator is satisfied for every representation, then |p(xα1,,xαn,xα1,,xαn)|=0{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|p(x_{\alpha_{1}},...,x_{\alpha_{n}},x_{\alpha_{1}}^{*},...,x_{\alpha_{n}}^{*})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=0, and thus p(xα1,,xαn,xα1,,xαn)Jp(x_{\alpha_{1}},...,x_{\alpha_{n}},x_{\alpha_{1}}^{*},...,x_{\alpha_{n}}^{*})\in J, so p(zα1,,zαn,zα1,,zαn)=0p(z_{\alpha_{1}},...,z_{\alpha_{n}},z_{\alpha_{1}}^{*},...,z_{\alpha_{n}}^{*})=0. That is, the {zα}αI\{z_{\alpha}\}_{\alpha\in I} satisfy the relators.

Now note that ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} is a CC^{*}-seminorm, so the norm X+J:=|X|\left\|{X+J}\right\|:={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} is a well-defined CC^{*}-norm on (𝒢)/J\mathcal{F}(\mathcal{G})/J, and thus the completion of (𝒢)/J\mathcal{F}(\mathcal{G})/J with respect to this norm is a CC^{*}-algebra. Thus ψ:xαzα\psi:x_{\alpha}\rightarrow z_{\alpha} is a representation of (𝒢,)(\mathcal{G},\mathcal{R}).

Now, given any other presentation φ:xαyα\varphi:x_{\alpha}\mapsto y_{\alpha}, note that kerφ¯={X(𝒢):φ¯(X)=0}{X(𝒢):|X|=0}=kerψ¯\ker\bar{\varphi}=\{X\in\mathcal{F}(\mathcal{G}):\left\|{\bar{\varphi}(X)}\right\|=0\}\subseteq\{X\in\mathcal{F}(\mathcal{G}):{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|X\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=0\}=\ker\bar{\psi}, so by Lemma 2.8(2), there is a function on the *-algebras generated by yy and zz given by yαzαy_{\alpha}\mapsto z_{\alpha}, and by Lemma 2.8(3) applied to +,×,+,\times, and , this function is a *-homomorphism, so it extends to a *-homomorphism π:C({zα}αI)C({yα}αI)\pi:C^{*}(\{z_{\alpha}\}_{\alpha\in I})\rightarrow C^{*}(\{y_{\alpha}\}_{\alpha\in I}) as desired. ∎

2.2.3 Conditional Expectations

We will also make frequent use of conditional expectations, whose definition we include here. For a more thorough treatment of conditional expectations, the reader is directed to [23, Chapter III.3], where they are called projections of norm one.

Definition 2.10.

Let CC be a CC^{*}-algebra, and DCD\subseteq C a CC^{*}-subalgebra. We say a conditional expectation or projection of norm one is a linear map Φ:CD\Phi:C\rightarrow D such that:

  • Φ\Phi is contractive, meaning Φ(c)c\left\|{\Phi(c)}\right\|\leq\left\|{c}\right\| for all cCc\in C (and is thus continuous).

  • Φ\Phi is idempotent, meaning ΦΦ=Φ\Phi\circ\Phi=\Phi.

  • Φ(d)=d\Phi(d)=d for all dDd\in D.

If additionally c>0c>0 implies Φ(c)>0\Phi(c)>0, we say Φ\Phi is faithful.

Example 2.11.

In a matrix algebra MnM_{n}, there is a faithful conditional expectation Φ\Phi given by “restricting to the diagonal”, meaning Φ([aij])=[bij]\Phi([a_{ij}])=[b_{ij}] where

bij={aij if i=j0 if ij.b_{ij}=\begin{cases}a_{ij}&\text{ if }i=j\\ 0&\text{ if }i\neq j\end{cases}.

The following result is [24, Theorem 1] and is known as Tomiyama’s Theorem. It says that a conditional expectation has some additional properties “for free”. Some authors will define conditional expectations as requiring these properties, and some will define them to only require the shorter list of conditions (as we have done here).

Theorem 2.12 (Tomiyama’s Theorem).

If Φ:CD\Phi:C\rightarrow D is a conditional expectation, then Φ\Phi is positive, meaning that x0x\geq 0 implies Φ(x)0\Phi(x)\geq 0. Furthermore, Φ\Phi is a DD-bimodule map, meaning Φ(d1cd2)=d1Φ(c)d2\Phi(d_{1}cd_{2})=d_{1}\Phi(c)d_{2} for all d1,d2Dd_{1},d_{2}\in D and cCc\in C.

2.2.4 Tensor Products

Here we will give a very brief introduction to the minimal tensor product. A more thorough introduction can be found in [23, Chapter 4]. Throughout this paper, as in much of the coaction literature, we will use unadorned \otimes for the minimal tensor product of CC^{*}-algebras.

Definition 2.13.

Given two vector spaces VV and WW, we let VWV\odot W denote the algebraic tensor product of VV and WW, meaning the \mathbb{C}-vector space spanned by formal symbols {vw:vV,wW}\{v\otimes w:v\in V,w\in W\} with the relations of linearity over either coordinate.

Let AA and BB be CC^{*}-algebras. Then one can give ABA\odot B a *-algebra structure by

(ab)(ab)=(aa)(bb) and (ab)=ab.(a\otimes b)(a^{\prime}\otimes b^{\prime})=(aa^{\prime})\otimes(bb^{\prime})\text{ and }(a\otimes b)^{*}=a^{*}\otimes b^{*}.

If π:A(H)\pi:A\rightarrow\mathcal{B}(H) and ρ:B(𝒦)\rho:B\rightarrow\mathcal{B}(\mathcal{K}) are representations of AA and BB on Hilbert spaces \mathcal{H} and 𝒦\mathcal{K} respectively, then there exists a *-algebra representation πρ:AB(𝒦)\pi\otimes\rho:A\odot B\rightarrow\mathcal{B}(\mathcal{H}\otimes\mathcal{K}) satisfying [(πρ)(ab)](hk)=π(a)hρ(b)k\left[(\pi\otimes\rho)(a\otimes b)\right](h\otimes k)=\pi(a)h\otimes\rho(b)k for all aA,bB,h,k𝒦a\in A,b\in B,h\in\mathcal{H},k\in\mathcal{K}. The minimal norm on ABA\odot B is defined by

i=1naibimin=sup{i=1nπ(ai)ρ(bi)}\left|\left|\displaystyle\sum_{i=1}^{n}a_{i}\otimes b_{i}\right|\right|_{min}=\sup\left\{\left|\left|\displaystyle\sum_{i=1}^{n}\pi(a_{i})\otimes\rho(b_{i})\right|\right|\right\}

where the supremum is taken over all representations π\pi and ρ\rho of AA and BB, respectively (and this is indeed a norm which is minimal in an appropriate sense). Note that abmin=ab\left\|{a\otimes b}\right\|_{min}=\left\|{a}\right\|\cdot\left\|{b}\right\| (that is, min\left\|{\cdot}\right\|_{min} is a “CC^{*}-cross norm”). The minimal tensor product AminBA\otimes_{min}B is the CC^{*}-algebra created by completing ABA\odot B with respect to this norm. Since all of our tensor products will be minimal, we will henceforth write \left\|{\cdot}\right\| for min\left\|{\cdot}\right\|_{min} and ABA\otimes B for AminBA\otimes_{min}B.

The following result will be used occasionally, and it is a minor variant of [6, Lemma A.1]:

Lemma 2.14.

Let φ:AC\varphi:A\rightarrow C and ψ:BD\psi:B\rightarrow D be *-homomorphisms of CC^{*}-algebras. Then there is a homomorphism φψ:ABCD\varphi\otimes\psi:A\otimes B\rightarrow C\otimes D satisfying (φψ)(ab)=φ(a)ψ(b)(\varphi\otimes\psi)(a\otimes b)=\varphi(a)\otimes\psi(b). If φ\varphi and ψ\psi are nondegenerate (respectively, faithful), then so is φψ\varphi\otimes\psi.

Proof.

Consider CC and DD as subalgebras (respectively) of their multiplier algebras M(C)M(C) and M(D)M(D), and therefore consider φ\varphi and ψ\psi as mapping into M(C)M(C) and M(D)M(D), respectively. By [6, Lemma A.1], there is a *-homomorphism φψ:ABM(CD)\varphi\otimes\psi:A\otimes B\rightarrow M(C\otimes D) satisfying (φψ)(ab)=φ(a)ψ(b)(\varphi\otimes\psi)(a\otimes b)=\varphi(a)\otimes\psi(b), and if φ\varphi and ψ\psi are nondegenerate (respectively, faithful), then so is φψ\varphi\otimes\psi.

It then suffices to show that φψ\varphi\otimes\psi maps into CDC\otimes D properly, instead of mapping into M(CD)M(C\otimes D). But for all aA,bBa\in A,b\in B, we have φ(a)C\varphi(a)\in C and φ(b)D\varphi(b)\in D, so (φψ)(ab)CD(\varphi\otimes\psi)(a\otimes b)\in C\otimes D for all aA,bBa\in A,b\in B. Taking closed spans, we get that for any xABx\in A\otimes B, we have (φψ)(x)C×D(\varphi\otimes\psi)(x)\in C\times D, as desired. ∎

2.3 Connections between Groups and CC^{*}-Algebras

In this section we will remind the reader of some of the many connections between groups and CC^{*}-algebras, including group CC^{*}-algebras, amenability, Fell bundles, gradings, coactions, and actions. Our context is relatively simple since we are limiting our attention to discrete groups.

2.3.1 Group CC^{*}-algebras and Amenable Groups

The following discussion of group CC^{*}-algebras and amenability is adapted from [5, Chapter VII], to which the reader is directed for a more thorough treatment.

Definition 2.15.

Let GG be a discrete group. Then a representation of GG in a CC^{*}-algebra is a collection of unitary operators {ug}gG\{u_{g}\}_{g\in G} such that uguh=ughu_{g}u_{h}=u_{gh} and ug=ug1u_{g}^{*}=u_{g^{-1}} for all g,hGg,h\in G.

Since these relators are all polynomials, then by Lemma 2.9 there is a representation of GG which is universal for all representations. We denote it by {Ug}gG\{U_{g}\}_{g\in G}, and call C(G):=C({Ug}gG)C^{*}(G):=C^{*}(\{U_{g}\}_{g\in G}) the full group CC^{*}-algebra.

A particularly nice representation of a group GG, called the left-regular representation of a group, comes from the natural action of GG multiplying on itself. In particular, let {eg}gG\{e_{g}\}_{g\in G} denote the standard orthonormal basis for 2(G)\ell^{2}(G), and define an operator LgL_{g} by Lgeg=eggL_{g}e_{g^{\prime}}=e_{gg^{\prime}} for g,gGg,g^{\prime}\in G. Then {Lg}gG\{L_{g}\}_{g\in G} is a representation of GG in (2(G))\mathcal{B}(\ell^{2}(G)). We call Cr(G):=C({Lg}gG)C_{r}^{*}(G):=C^{*}(\{L_{g}\}_{g\in G}) the reduced group CC^{*}-algebra.

The left-regular representation is more commonly denoted by λ\lambda instead of LL, but our choice of notation will help avoid ambiguity with paths λΛ\lambda\in\Lambda later in the text.

Note that C(G)C^{*}(G) and Cr(G)C^{*}_{r}(G) are the closed spans of {Ug}gG\{U_{g}\}_{g\in G} and {Lg}gG\{L_{g}\}_{g\in G}, respectively.

Famously, there are many equivalent definitions of amenability, but for our purposes, this one is the most convenient:

Definition 2.16.

By the universal property of C(G)C^{*}(G), there is a surjective *-homomorphism πLU:C(G)Cr(G)\pi^{U}_{L}:C^{*}(G)\rightarrow C^{*}_{r}(G) given by UgLgU_{g}\mapsto L_{g}. We say that GG is amenable if and only if πLU\pi^{U}_{L} is injective (and hence an isomorphism).

Readers more familiar with another definition of amenability may wish to read [5, Theorem VII.2.5] which proves the equivalence of this definition with a more common one.

The following remark summarizes some of the well-known results about amenability of groups.

Remark 2.17.

By [5, Proposition VII.2.3], for discrete groups amenability is preserved under:

  1. 1.

    Subgroups

  2. 2.

    Quotients

  3. 3.

    Direct Limits

  4. 4.

    Extensions (meaning that if 1NGH11\rightarrow N\rightarrow G\rightarrow H\rightarrow 1 is a short exact sequence of groups, with NN and HH amenable, then GG is amenable).

  5. 5.

    Finite direct products (which is immediate from being closed under extensions).

Furthermore, by [5, Proposition VII.2.2], every abelian group is amenable. Every finite group is amenable since C(G)C^{*}(G) and Cr(G)C^{*}_{r}(G) have the same finite dimension |G||G|, so by the rank-nullity theorem the surjective map πLU\pi^{U}_{L} is injective.

Notably, the free group on two generators is not amenable by [5, Example VII.2.4]. By the subgroup property, any group containing a free group on two (or more) generators is also not amenable.

2.3.2 Fell Bundles, Topologically Graded CC^{*}-algebras, and Coactions

In this section, we will give a short introduction to Fell bundles, topologically graded CC^{*}-algebras, and coactions, and show some of the ways these closely-related structures overlap. For a more detailed treatment, the reader is directed to [10, 9] for Fell bundles, [10] for topological gradings, and [6, Appendix A] for coactions.

Throughout this section, all of our groups will be discrete, which simplifies some definitions. The following is from [10, Definition 16.1]:

Definition 2.18.

Let GG be a discrete group. Let ={Bg}gG\mathcal{B}=\{B_{g}\}_{g\in G} be a collection of Banach spaces, and write \mathscr{B} for the disjoint union of the {Bg}gG\{B_{g}\}_{g\in G}, called the total space. Suppose \mathscr{B} has a binary operation \cdot called multiplication, and an involution * which satisfy the following properties for all g,hGg,h\in G and b,cb,c\in\mathscr{B}:

  1. a.

    BgBhBghB_{g}B_{h}\subseteq B_{gh},

  2. b.

    Multiplication is bilinear from Bg×BhB_{g}\times B_{h} to BghB_{gh},

  3. c.

    Multiplication on \mathscr{B} is associative,

  4. d.

    bcbc\left\|{bc}\right\|\leq\left\|{b}\right\|\cdot\left\|{c}\right\|,

  5. e.

    (Bg)Bg1(B_{g})^{*}\subseteq B_{g^{-1}},

  6. f.

    Involution is conjugate-linear from BgB_{g} to Bg1B_{g^{-1}},

  7. g.

    (bc)=cb(bc)^{*}=c^{*}b^{*},

  8. h.

    b=bb^{**}=b,

  9. i.

    b=b\left\|{b^{*}}\right\|=\left\|{b}\right\|,

  10. j.

    bb=b2\left\|{b^{*}b}\right\|=\left\|{b}\right\|^{2},

  11. k.

    bb0b^{*}b\geq 0 in B1B_{1}.

Then we say that \mathcal{B} is a Fell bundle over GG. We call each BgB_{g} a fiber.

The following is from [10, Definitions 16.2 and 19.2]:

Definition 2.19.

Let AA be a CC^{*}-algebra, and let GG be a (discrete) group. We say that a (CC^{*}-) grading for AA is a collection {Ag}gG\{A_{g}\}_{g\in G} of linearly independent closed subspaces such that gGAg\bigoplus_{g\in G}A_{g} is dense in AA, AgAhAghA_{g}A_{h}\subseteq A_{gh}, and AgAg1A_{g}^{*}\subseteq A_{g^{-1}} for g,hGg,h\in G. Each AgA_{g} is called a graded subspace or graded component.

If there is also a conditional expectation Φ:AAe\Phi:A\rightarrow A_{e} satisfying

Φ(a)={a if aAe0 if aAg for ge,\Phi(a)=\begin{cases}a&\text{ if }a\in A_{e}\\ 0&\text{ if }a\in A_{g}\text{ for }g\neq e\end{cases},

then we say that {Ag}gG\{A_{g}\}_{g\in G} is a topological grading.

Remark 2.20.

It is straightforward to verify that if AA is a topologically graded CC^{*}-algebra, then its graded components form a Fell bundle. Most of the Fell bundles we will use arise in this way.

Remark 2.21.

Given a Fell bundle \mathcal{B}, one can construct a “reduced” and “full” cross sectional algebra representing this bundle, respectively denoted Cr()C^{*}_{r}(\mathcal{B}) and C()C^{*}(\mathcal{B}) (see Definition 2.3 of [9] and the comment following it).

By [9, Theorem 3.3], if BB is a topologically graded CC^{*}-algebra, and \mathcal{B} denotes its associated Fell bundle, then there is a *-homomorphism L:BCr()L:B\rightarrow C^{*}_{r}(\mathcal{B}) called the left-regular representation of the Fell bundle. Combining this result with their Proposition 3.7, LL is an isomorphism if and only if the conditional expectation from the topological grading is faithful.

In a Fell bundle \mathcal{B}, the full cross sectional algebra C()C^{*}(\mathcal{B}) has a topological grading, and the bundle is called amenable if its left-regular representation L:C()Cr()L:C^{*}(\mathcal{B})\rightarrow C^{*}_{r}(\mathcal{B}) is injective (and hence an isomorphism). The reader may notice the parallel with our definition of amenability for a group (Definition 2.16).

The following two results are respectively [9, Theorem 4.7] and [9, Proposition 4.2] :

Lemma 2.22.

Let GG be a discrete amenable group. Then every Fell bundle over GG is amenable.

Lemma 2.23.

If \mathcal{B} is an amenable Fell bundle, then all topologically graded CC^{*}-algebras whose associated Fell bundles coincide with \mathcal{B} are isomorphic to each other.

Finally, we will give a brief introduction to coactions. The following is both the simplest example of a coaction, and a necessary component of its definition.

Example 2.24.

If GG is a discrete group, then the operators {UgUg}gGC(G)C(G)\{U_{g}\otimes U_{g}\}_{g\in G}\subset C^{*}(G)\otimes C^{*}(G) are a representation of GG. Therefore, by the universal property of C(G)C^{*}(G), there is a *-homomorphism δG:C(G)C(G)C(G)\delta_{G}:C^{*}(G)\rightarrow C^{*}(G)\otimes C^{*}(G) given by UgUgUgU_{g}\mapsto U_{g}\otimes U_{g}.

Note that letting idG:=idC(G)\operatorname{id}_{G}:=\operatorname{id}_{C^{*}(G)}, then (δGidG)δG=(idGδG)δG(\delta_{G}\otimes\operatorname{id}_{G})\circ\delta_{G}=(\operatorname{id}_{G}\otimes\delta_{G})\circ\delta_{G}, which can be easily verified by checking that both send UgUgUgUgC(G)C(G)C(G)U_{g}\mapsto U_{g}\otimes U_{g}\otimes U_{g}\in C^{*}(G)\otimes C^{*}(G)\otimes C^{*}(G).

The following definition of coactions for discrete groups is taken from [7, Section 2]:

Definition 2.25.

Let GG be a discrete group. A coaction of GG on a CC^{*}-algebra AA is an injective, nondegenerate homomorphism δ:AAC(G)\delta:A\rightarrow A\otimes C^{*}(G) satisfying the coaction identity that (δidG)δ=(idAδG)δ(\delta\otimes\operatorname{id}_{G})\circ\delta=(\operatorname{id}_{A}\otimes\delta_{G})\circ\delta as maps from AA into AC(G)C(G)A\otimes C^{*}(G)\otimes C^{*}(G), summarized in this diagram:

A{A}AC(G){A\otimes C^{*}(G)}AC(G){A\otimes C^{*}(G)}AC(G)C(G){A\otimes C^{*}(G)\otimes C^{*}(G)}δ\scriptstyle{\delta}δ\scriptstyle{\delta}δidG\scriptstyle{\delta\otimes\operatorname{id}_{G}}idAδG\scriptstyle{\operatorname{id}_{A}\otimes\delta_{G}}

In this context, nondegeneracy means that span¯[δ(A)(AC(G))]=AC(G)\overline{\operatorname{span}}\left[\delta(A)(A\otimes C^{*}(G))\right]=A\otimes C^{*}(G).

We call the triple (A,G,δ)(A,G,\delta) a cosystem.

If GG is discrete, a cosystem has a nice topological grading as described in [20, Proposition A.3]:

Lemma 2.26.

Let (A,G,δ)(A,G,\delta) be a discrete cosystem, and for gGg\in G let

Ag={aA|δ(a)=aUg}.A_{g}=\{a\in A|\delta(a)=a\otimes U_{g}\}.

Then the collection {Ag}gG\{A_{g}\}_{g\in G} is a topological grading of AA. We will write ΦA:AAe\Phi_{A}:A\rightarrow A_{e} for the conditional expectation.

Some coactions are particularly nice, having a condition called normality. There are many definitions of normality, but we will find a definition depending on this conditional expectation to be the most convenient:

Definition 2.27.

We say a discrete coaction (A,G,δ)(A,G,\delta) is normal if the conditional expectation ΦA\Phi_{A} is faithful.

A more common definition of normality of a coaction (such as [6, Definition A.50] or the comments preceding Definition 1.1 of [17]) is that the coaction is normal if and only if the map jA:=(idAπLU)δj_{A}:=(\operatorname{id}_{A}\otimes\pi^{U}_{L})\circ\delta is injective. Our definition is proved equivalent to the more common one in [17, Lemma 1.4].

For a coaction over a discrete amenable group, this property is automatic:

Lemma 2.28.

Let (A,G,δ)(A,G,\delta) be a cosystem. If GG is amenable and discrete, then δ\delta is normal.

Proof.

For a cosystem (A,G,δ)(A,G,\delta) over an amenable group, by Lemma 2.26 there is a topological grading {Ag}gG\{A_{g}\}_{g\in G} of AA. By Remark 2.20, these graded components form a Fell bundle \mathcal{B}, and this Fell bundle is amenable by Lemma 2.22.

Since \mathcal{B} is amenable, by Lemma 2.23, ACr()A\cong C^{*}_{r}(\mathcal{B}). But for any Fell bundle, Cr()C^{*}_{r}(\mathcal{B}) has a conditional expectation by [9, Proposition 2.9], which is faithful by [9, Proposition 2.12]. ∎

Finally, we will show that coactions, topological gradings, and Fell bundles are equivalent over amenable groups.

Lemma 2.29.

Let GG be an amenable group, and let AA be a CC^{*}-algebra. Then the following are equivalent:

  1. 1.

    AA has a coaction by GG.

  2. 2.

    AA has a topological grading {Ag}gG\{A_{g}\}_{g\in G}.

  3. 3.

    There is a Fell bundle ={Bg}gG\mathcal{B}=\{B_{g}\}_{g\in G} whose fibers are linearly independent closed subspaces of AA such that AC()A\cong C^{*}(\mathcal{B}).

And under these conditions, the structures coincide in the sense that the fibers {Bg}gG\{B_{g}\}_{g\in G} of the Fell bundle are equal to the graded components {Ag}gG\{A_{g}\}_{g\in G}, and the graded components arise from the coaction via Ag={aA:δ(a)=aUg}A_{g}=\{a\in A:\delta(a)=a\otimes U_{g}\}.

Proof.

By Lemma 2.26, (1) implies (2).

To show that (2) implies (3), suppose that AA has a topological grading {Ag}gG\{A_{g}\}_{g\in G}. By Remark 2.20, these fibers form a Fell bundle. To see that the fibers are linearly independent, fix a finite sum i=1nagi\displaystyle\sum_{i=1}^{n}a_{g_{i}} where each agiAgia_{g_{i}}\in A_{g_{i}} for some distinct giGg_{i}\in G, and suppose this sum equals 0. Then by [10, Corollary 19.6], there are contractive linear maps Fg:AAgF_{g}:A\rightarrow A_{g} satisfying Fgj(i=1nagi)=agjF_{g_{j}}(\displaystyle\sum_{i=1}^{n}a_{g_{i}})=a_{g_{j}}. But since 0=i=1nagi0=\displaystyle\sum_{i=1}^{n}a_{g_{i}}, then for each gjg_{j}, we have 0=Fgj(i=1nagi)=agj0=F_{g_{j}}(\displaystyle\sum_{i=1}^{n}a_{g_{i}})=a_{g_{j}}, so each summand is 0, so the graded components are linearly independent. Finally, we must show that AC()A\cong C^{*}(\mathcal{B}). By Lemma 2.22, \mathcal{B} is amenable, and by [10, Proposition 19.3], C()C^{*}(\mathcal{B}) is a topologically graded CC^{*}-algebra. Since this amenable bundle coincides with the bundle in AA, then by Lemma 2.23, AC()A\cong C^{*}(\mathcal{B}), as desired.

To see that (3) implies (1), suppose that ={Bg}gG\mathcal{B}=\{B_{g}\}_{g\in G} is a Fell bundle of linearly independent closed subspaces of AA such that AC()A\cong C^{*}(\mathcal{B}). Since GG is amenable, C(G)Cr(G)C^{*}(G)\cong C^{*}_{r}(G), and by Lemma 2.22, \mathcal{B} is an amenable Fell bundle, so C()Cr()C^{*}(\mathcal{B})\cong C^{*}_{r}(\mathcal{B}). Combining these simplifications with [10, Proposition 18.7], there is an injective *-homomorphism δ:C()C()C(G)\delta:C^{*}(\mathcal{B})\rightarrow C^{*}(\mathcal{B})\otimes C^{*}(G) satisfying δ(bg)=bgUg\delta(b_{g})=b_{g}\otimes U_{g} for all gGg\in G and bgBgb_{g}\in B_{g}. It is routine to verify that this is a coaction, and that the graded parts of this coaction coincide with the original Fell bundle \mathcal{B}.

2.3.3 Coactions as Actions of the Dual Group

In this section we will show that if GG is discrete and abelian, then a coaction by GG is equivalent to an action by its dual group G^\widehat{G}. For a non-abelian group, this correspondence breaks down because there is no dual group.

First, we will remind the reader of actions by groups and dual groups. The following definition is well-known, and can be found for instance in the remarks preceding Proposition 2.1 in [18]:

Definition 2.30.

Given a locally compact group GG, an action of GG on a CC^{*}-algebra AA is a strongly continuous homomorphism α:GAut(A)\alpha:G\rightarrow\operatorname{Aut}(A), the group of automorphisms of AA. (Here, strong continuity means that for all aAa\in A, the map gαg(a)g\mapsto\alpha_{g}(a) is continuous as a function from GG to AA.) The trio (A,G,α)(A,G,\alpha) is called a (CC^{*}-dynamical) system.

We will now define the dual group. A more detailed treatment of dual groups can be found in [12, Chapter 1.7].

Definition 2.31.

Given a abelian locally compact group GG, we say that a character of GG is a continuous homomorphism χ:G𝕋\chi:G\rightarrow\mathbb{T}, where 𝕋={z:|z|=1}\mathbb{T}=\{z\in\mathbb{C}:|z|=1\} denotes the unit circle in the complex numbers. The set of characters G^\widehat{G} forms an abelian locally compact group called the dual group under the operation (χ1χ2)(g):=χ1(g)χ2(g)(\chi_{1}*\chi_{2})(g):=\chi_{1}(g)\chi_{2}(g) and the topology of uniform convergence on compacta.

Some examples of dual groups are that ^k𝕋k\widehat{\mathbb{Z}}^{k}\cong\mathbb{T}^{k}, ^kk\widehat{\mathbb{R}}^{k}\cong\mathbb{R}^{k}, and that if GG is finite and abelian then G^G\widehat{G}\cong G.

The following two theorems are well-known results about the duals of abelian locally compact groups. They can be found in [12, Theorem 1.85] and [12, Theorem 1.88] respectively.

Theorem 2.32 (Pontryagin duality theorem).

An abelian locally compact group is naturally isomorphic to its double dual by the map g[χχ(g)]g\mapsto[\chi\mapsto\chi(g)]. That is, GG^^G\cong\widehat{\widehat{G}}.

Theorem 2.33.

If GG is a locally compact abelian group, then GG is discrete if and only if G^\widehat{G} is compact.

We are now ready to prove the equivalence between coactions and actions by a dual group.

Lemma 2.34.

Let GG be a discrete abelian group and let AA be a CC^{*}-algebra. Then for every CC^{*}-dynamical system (A,G^,α)(A,\widehat{G},\alpha), there is a unique cosystem (A,G,δ)(A,G,\delta) which satisfies

Ag={aA:αχ(a)=χ(g)a for all χG^}A_{g}=\{a\in A:\alpha_{\chi}(a)=\chi(g)\cdot a\text{ for all }\chi\in\widehat{G}\}

for all gGg\in G. All cosystems by discrete abelian groups arise in this way.

Proof.

Since GG is abelian, it is amenable. Thus by Lemma 2.29, a cosystem is equivalent to a topological grading whose graded components are the {Ag}gG\{A_{g}\}_{g\in G}. By [19, Theorem 3], such a topological grading is equivalent to the desired group action. ∎

2.4 Graphs and PP-graphs

Definition 2.35.

Let (G,P)(G,P) be an ordered group, and consider PP as a category with one object where the morphisms are the elements of PP under their multiplication structure. A PP-graph is a countable category Λ\Lambda along with a functor d:ΛPd:\Lambda\rightarrow P with the unique factorization property: if λΛ\lambda\in\Lambda and p1,p2Pp_{1},p_{2}\in P with p1p2=d(λ)p_{1}p_{2}=d(\lambda), then there exists unique λ1,λ2Λ\lambda_{1},\lambda_{2}\in\Lambda with λ=λ1λ2\lambda=\lambda_{1}\lambda_{2} and d(λ1)=p1d(\lambda_{1})=p_{1} (and hence necessarily d(λ2)=p2d(\lambda_{2})=p_{2}).

We refer to the morphisms of Λ\Lambda as paths, and identity morphisms in Λ\Lambda as vertices. We let Λ0\Lambda^{0} denote the set of vertices in Λ\Lambda, and for pPp\in P, we write Λp={μΛ:d(μ)=p}\Lambda^{p}=\{\mu\in\Lambda:d(\mu)=p\}. As is common in the literature, we identify the objects in Λ\Lambda with the identity morphism at those objects, so we will refer to an object vv with its identity morphism idv\operatorname{id}_{v} interchangeably. Given a path λΛ\lambda\in\Lambda, let r(λ)r(\lambda) denote its range vertex and s(λ)s(\lambda) its source vertex. We write the composition of paths “working backwards” so that given α,βΛ\alpha,\beta\in\Lambda, the product αβ\alpha\beta is defined if and only if s(α)=r(β)s(\alpha)=r(\beta), in which case s(αβ)=s(β)s(\alpha\beta)=s(\beta) and r(αβ)=r(α)r(\alpha\beta)=r(\alpha). This can be summarized in the following diagram:

r(α){r(\alpha)}s(α)=r(β){s(\alpha)=r(\beta)}s(β){s(\beta)\par}α\scriptstyle{\alpha}β\scriptstyle{\beta}αβ\scriptstyle{\alpha\beta}
Remark 2.36.

The existence of the degree functor and the unique factorization property gives several nice properties to the graph.

First, Λ0={λΛ:d(λ)=1G}\Lambda^{0}=\{\lambda\in\Lambda:d(\lambda)=1_{G}\}, since for a vertex vv, we have v2=vv^{2}=v, so d(v)2=d(v)d(v)^{2}=d(v), so d(v)=1Gd(v)=1_{G}, and conversely if d(v)=1Gd(v)=1_{G}, then r(v)v=v=vs(v)r(v)v=v=vs(v) are two factorizations with the same degrees, so r(v)=v=s(v)r(v)=v=s(v) and hence vv is a vertex.

Second, having αβ=s(β)\alpha\beta=s(\beta) implies α=β=s(β)\alpha=\beta=s(\beta), since then d(α)d(β)=d(s(β))=1d(\alpha)d(\beta)=d(s(\beta))=1, so d(α),d(β)PP1={1}d(\alpha),d(\beta)\in P\cap P^{-1}=\{1\}.

Finally, the category has both left- and right-cancellation. For left cancellation, if αβ=αγ\alpha\beta=\alpha\gamma, then this provides two factorizations, so by the uniqueness of the factorizations, we have β=γ\beta=\gamma, and a nearly identical argument shows right-cancellation.

These properties together imply that a PP-graph is a category of paths in the sense of [22]. Our representation of a PP-graph will be a special case of their representation of a category of paths, and the reader is invited to compare our Definitions 2.41 and 2.44c to Theorems 6.3 and 8.2 of [22], respectively.

Definition 2.37.

Let (G,P)(G,P) be an ordered group, let Λ\Lambda be a PP-graph. For αΛ\alpha\in\Lambda, we write αΛ={αμ:μΛ}\alpha\Lambda=\{\alpha\mu:\mu\in\Lambda\}.

We can give Λ\Lambda a partial order \leq by saying αβ\alpha\leq\beta if there exists some α1Λ\alpha_{1}\in\Lambda such that αα1=β\alpha\alpha_{1}=\beta. That is, αβ\alpha\leq\beta if and only if βαΛ\beta\in\alpha\Lambda.

We say that α\alpha and β\beta have a common extension if there is a μΛ\mu\in\Lambda such that αμ\alpha\leq\mu and βμ\beta\leq\mu.

Given α,βΛ\alpha,\beta\in\Lambda, we say that μΛ\mu\in\Lambda is a minimal common extension of α\alpha and β\beta if μ\mu is a common extension of α\alpha and β\beta, and for all other common extensions ν\nu, νμ\nu\leq\mu implies ν=μ\nu=\mu. We let MCE(α,β)MCE(\alpha,\beta) denote the set of minimal common extensions of α\alpha and β\beta.

For general ordered groups, the ordering on Λ\Lambda may be sufficiently poorly behaved that there are no minimal common extensions, even if there are common extensions. However, the humble hypothesis of weak quasi-lattice ordering on (G,P)(G,P) prevents this catastrophe, and therefore will recur as a hypothesis in almost all results relating to PP-graphs. The following lemma is the main “entry-point” for the hypothesis of weak quasi-lattice ordering into PP-graphs:

Lemma 2.38.

Let (G,P)(G,P) be an ordered group, let Λ\Lambda be a PP-graph, and α,βΛ\alpha,\beta\in\Lambda. If (G,P)(G,P) is weakly quasi-lattice ordered (WQLO), then:

  1. 1.

    Let MDCE(α,β)={μαΛβΛ:d(μ)=d(α)d(β)}MDCE(\alpha,\beta)=\{\mu\in\alpha\Lambda\cap\beta\Lambda:d(\mu)=d(\alpha)\vee d(\beta)\} denote the set of minimal degree common extensions. For every common extension λ\lambda of α\alpha and β\beta, there is a μMDCE(α,β)\mu\in MDCE(\alpha,\beta) with μλ\mu\leq\lambda.

  2. 2.

    MCE(α,β)=MDCE(α,β)MCE(\alpha,\beta)=MDCE(\alpha,\beta).

  3. 3.

    For every common extension λ\lambda of α\alpha and β\beta, there is a μMCE(α,β)\mu\in MCE(\alpha,\beta) with μλ\mu\leq\lambda.

  4. 4.

    αΛβΛ=μMCE(α,β)μΛ\displaystyle\alpha\Lambda\cap\beta\Lambda=\bigsqcup_{\mu\in MCE(\alpha,\beta)}\mu\Lambda, where \bigsqcup denotes a disjoint union.

Proof.

(1) Suppose that λ\lambda is a common extension of α\alpha and β\beta, so λ=αα1=ββ1\lambda=\alpha\alpha_{1}=\beta\beta_{1}.

Note that d(λ)d(α)d(\lambda)\geq d(\alpha) and d(λ)d(β)d(\lambda)\geq d(\beta), so d(λ)d(α)d(β)d(\lambda)\geq d(\alpha)\vee d(\beta). Thus by factorization, we may write λ=μμ1\lambda=\mu\mu_{1} where d(μ)=d(α)d(β)d(\mu)=d(\alpha)\vee d(\beta). Now, d(μ)d(α)d(\mu)\geq d(\alpha), so μ\mu may be additionally factored as μ=αα2\mu=\alpha^{\prime}\alpha_{2}, where d(α)=d(α)d(\alpha^{\prime})=d(\alpha). Then we have λ=αα1=μμ1=αα2μ1\lambda=\alpha\alpha_{1}=\mu\mu_{1}=\alpha^{\prime}\alpha_{2}\mu_{1}, meaning we have two factorizations of λ\lambda whose initial segments are of equal length d(α)=d(α)d(\alpha^{\prime})=d(\alpha). Then by uniqueness of factorizations we have α=α\alpha^{\prime}=\alpha, so μ\mu is an extension of α\alpha. Similarly, μ\mu will be an extension of β\beta, so μ\mu is a common extension of α\alpha and β\beta of length d(α)d(β)d(\alpha)\vee d(\beta), so μMDCE(α,β)\mu\in MDCE(\alpha,\beta) as desired.

(2) If νMCE(α,β)\nu\in MCE(\alpha,\beta), then by (1) there is some μMDCE(α,β)\mu\in MDCE(\alpha,\beta) with μν\mu\leq\nu. Then by minimality of ν\nu, we have μ=ν\mu=\nu, so νMDCE(α,β)\nu\in MDCE(\alpha,\beta) and thus MCE(α,β)MDCE(α,β)MCE(\alpha,\beta)\subseteq MDCE(\alpha,\beta).

Conversely, if μMDCE(α,β)\mu\in MDCE(\alpha,\beta), suppose there were some νΛ\nu\in\Lambda with αν,βν,νμ\alpha\leq\nu,\beta\leq\nu,\nu\leq\mu. Then by (1), there would be some μMDCE(α,β)\mu^{\prime}\in MDCE(\alpha,\beta) with μνμ\mu^{\prime}\leq\nu\leq\mu. Then μμ\mu^{\prime}\leq\mu, so we may write μs(μ)=μ=μμ1\mu s(\mu)=\mu=\mu^{\prime}\mu_{1} for some μ1Λ\mu_{1}\in\Lambda, and since d(μ)=d(α)d(β)=d(μ)d(\mu)=d(\alpha)\vee d(\beta)=d(\mu^{\prime}), then by the uniqueness of the factorization we have μ=μ\mu=\mu^{\prime}. Then μ=ν\mu=\nu, so μ\mu is “minimal” in the appropriate sense such that μMCE(α,β)\mu\in MCE(\alpha,\beta). Thus MDCE(α,β)=MCE(α,β)MDCE(\alpha,\beta)=MCE(\alpha,\beta).

(3) Immediately follows from (1) and (2).

(4) It is immediate that μMCE(α,β)μΛαΛβΛ\bigcup_{\mu\in MCE(\alpha,\beta)}\mu\Lambda\subseteq\alpha\Lambda\cap\beta\Lambda, and by (3) we have αΛβΛμMCE(α,β)μΛ\alpha\Lambda\cap\beta\Lambda\subseteq\bigcup_{\mu\in MCE(\alpha,\beta)}\mu\Lambda, so it suffices to show the union is disjoint. To this end, if μ,νMCE(α,β)\mu,\nu\in MCE(\alpha,\beta) and λμΛνΛ\lambda\in\mu\Lambda\cap\nu\Lambda, then we would have λ=μμ1=νν1\lambda=\mu\mu_{1}=\nu\nu_{1}, but this presents two factorizations with equal degrees d(μ)=d(α)d(β)=d(ν)d(\mu)=d(\alpha)\vee d(\beta)=d(\nu), so we must have μ=ν\mu=\nu. Thus the {μΛ:μMCE(α,β)}\{\mu\Lambda:\mu\in MCE(\alpha,\beta)\} are pairwise disjoint, as desired. ∎

Remark 2.39.

It should be noted that much of the literature (e.g. [3, Definition 2.3]) defines MCE(α,β)MCE(\alpha,\beta) as we have defined MDCE(α,β)MDCE(\alpha,\beta), which the previous lemma shows are equivalent. We have chosen to emphasize the definition arising purely from the category structure because later we will be considering a category as a PP-graph and as a QQ-graph, and this definition clarifies that the category structure does not depend on the choice of the grading (as long as one exists).

2.5 Representations of PP-graphs and PP-graph Algebras

Two additional notions will be needed in our study of PP-graph algebras.

Definition 2.40.

Let (G,P)(G,P) be a WQLO group, and let Λ\Lambda be a PP-graph. We say that Λ\Lambda is finitely-aligned if for all μ,νΛ\mu,\nu\in\Lambda, MCE(μ,ν)MCE(\mu,\nu) is finite.

We say that a set AΛA\subseteq\Lambda is exhaustive for a set BΛB\subseteq\Lambda if for all βB\beta\in B there is an αA\alpha\in A such that α\alpha and β\beta have a common extension.

We are at last ready to define our main object of study, the representations of PP-graphs in a CC^{*}-algebra. The following definition is due to [3], with the slight modification to apply to WQLO groups:

Definition 2.41.

Let (G,P)(G,P) be a WQLO group and Λ\Lambda a finitely-aligned PP-graph. A representation of Λ\Lambda in a CC^{*}-algebra BB is a function t:ΛB,λtλt:\Lambda\rightarrow B,\lambda\mapsto t_{\lambda} such that:

  1. (T1)

    tv=tvt_{v}=t_{v}^{*} and tvtw=δv,wtvt_{v}t_{w}=\delta_{v,w}t_{v} for all v,wΛ0v,w\in\Lambda^{0} (here δ\delta denotes the Kronecker delta). That is, {tv:vΛ0}\{t_{v}:v\in\Lambda^{0}\} is a collection of pairwise orthogonal projections.

  2. (T2)

    tμtν=tμνt_{\mu}t_{\nu}=t_{\mu\nu} whenever s(μ)=r(ν)s(\mu)=r(\nu).

  3. (T3)

    tμtμ=ts(μ)t_{\mu}^{*}t_{\mu}=t_{s(\mu)} for all μΛ\mu\in\Lambda.

  4. (T4)

    tμtμtνtν=λMCE(μ,ν)tλtλt_{\mu}t_{\mu}^{*}t_{\nu}t_{\nu}^{*}=\displaystyle\sum_{\lambda\in MCE(\mu,\nu)}t_{\lambda}t_{\lambda}^{*} for all μ,νΛ\mu,\nu\in\Lambda.

We denote by C(t)C^{*}(t) the CC^{*}-algebra generated by the {tλ}λΛ\{t_{\lambda}\}_{\lambda\in\Lambda}, and C(t)C^{*}(t) is called a PP-graph CC^{*}-algebra or just PP-graph algebra.

We require that our graphs be finitely-aligned so that the expression in the (T4) relator is a finite sum. The fact that an infinite sum of projections does not (usually) converge in the norm topology is a perennial problem in the field of generalizing Cuntz-Krieger algebras, and our solution is to limit our attention to the finitely-aligned case where the issue does not arise.

We will often want to check the properties of C(t)C^{*}(t) on a dense spanning set, so the following lemma is useful. This result is widely known (see for instance [3, Remark 5.2]).

Lemma 2.42.

Let (G,P)(G,P) be a WQLO group, let Λ\Lambda a finitely aligned PP-graph, and let tt be a representation of a PP-graph. Then C(t)=span¯{tαtβ:α,βΛ}C^{*}(t)=\overline{\operatorname{span}}\{t_{\alpha}t_{\beta}^{*}:\alpha,\beta\in\Lambda\}.

Proof.

Fix some α,β,μ,νΛ\alpha,\beta,\mu,\nu\in\Lambda. If s(α)s(β)s(\alpha)\neq s(\beta) or s(μ)s(ν)s(\mu)\neq s(\nu), then by the T1 relator we have tαtβ=0t_{\alpha}t_{\beta}^{*}=0 or tμtν=0t_{\mu}t_{\nu}^{*}=0 respectively, and in either case (tαtβ)(tμtν)=0(t_{\alpha}t_{\beta}^{*})(t_{\mu}t_{\nu}^{*})=0. If instead s(α)=s(β)s(\alpha)=s(\beta) and s(μ)=s(ν)s(\mu)=s(\nu), then

(tαtβ)(tμtν)\displaystyle(t_{\alpha}t_{\beta}^{*})(t_{\mu}t_{\nu}^{*}) =\displaystyle= (tαtβ)(tβtβ)(tμtμ)(tμtν) by T3 and T2\displaystyle(t_{\alpha}t_{\beta}^{*})(t_{\beta}t_{\beta}^{*})(t_{\mu}t_{\mu}^{*})(t_{\mu}t_{\nu}^{*})\text{ by T3 and T2}
=\displaystyle= (tαtβ)(λMCE(β,μ)tλtλ)(tμtν) by T4\displaystyle(t_{\alpha}t_{\beta}^{*})\left(\displaystyle\sum_{\lambda\in MCE(\beta,\mu)}t_{\lambda}t_{\lambda}^{*}\right)(t_{\mu}t_{\nu}^{*})\text{ by T4}
=\displaystyle= λMCE(β,μ)tα(β1λ)tν(μ1λ) by T2 and T3\displaystyle\displaystyle\sum_{\lambda\in MCE(\beta,\mu)}t_{\alpha(\beta^{-1}\lambda)}t_{\nu(\mu^{-1}\lambda)}\text{ by T2 and T3}

where β1λ\beta^{-1}\lambda denotes the unique path for which β(β1λ)=λ\beta(\beta^{-1}\lambda)=\lambda, and similarly μ1λ\mu^{-1}\lambda denotes the unique path for which μ(μ1λ)=λ\mu(\mu^{-1}\lambda)=\lambda.

This calculation shows that span{tαtβ:α,βΛ}\operatorname{span}\{t_{\alpha}t_{\beta}^{*}:\alpha,\beta\in\Lambda\} is closed under multiplication, so it is a *-subalgebra of C(t)C^{*}(t), and it contains each tαt_{\alpha} since tα=tαts(α)t_{\alpha}=t_{\alpha}t_{s(\alpha)}^{*}. Thus span¯{tαtβ:α,βΛ}\overline{\operatorname{span}}\{t_{\alpha}t_{\beta}^{*}:\alpha,\beta\in\Lambda\} is a CC^{*}-subalgebra of C(t)C^{*}(t) containing all of the generators, so C(t)=span¯{tαtβ:α,βΛ}C^{*}(t)=\overline{\operatorname{span}}\{t_{\alpha}t_{\beta}^{*}:\alpha,\beta\in\Lambda\}.

Example 2.43.

As we have already seen with groups and Fell bundles, a simple way of representing an object often comes from its own action on itself, and such an example arises for representations of PP-graphs as well. To define the left-regular representation of a PP-graph Λ\Lambda, let =2(Λ)\mathcal{H}=\ell^{2}(\Lambda), with the typical orthonormal basis {eα}αΛ\{e_{\alpha}\}_{\alpha\in\Lambda}. (Kribs and Power call this the Fock space of Λ\Lambda in their papers [13, 14].) For μΛ\mu\in\Lambda define a “forward shift by μ\mu” operator LμL_{\mu} by

Lμeα={eμα if s(μ)=r(α)0 if s(μ)r(α)L_{\mu}e_{\alpha}=\begin{cases}e_{\mu\alpha}&\text{ if }s(\mu)=r(\alpha)\\ 0&\text{ if }s(\mu)\neq r(\alpha)\end{cases}

and extending this linearly. A short computation on the inner product shows that LμL_{\mu}^{*} is the “backwards shift by μ\mu” operator, which is given by

Lμeβ={eα if there exists α with μα=β0 if no such α existsL_{\mu}^{*}e_{\beta}=\begin{cases}e_{\alpha}&\text{ if there exists }\alpha\text{ with }\mu\alpha=\beta\\ 0&\text{ if no such }\alpha\text{ exists}\end{cases}

(noting that if such an α\alpha does exist, it is unique by left cancellation).

We will now show that {Lμ}μΛ\{L_{\mu}\}_{\mu\in\Lambda} is a representation of Λ\Lambda. The T1 and T2 relations are immediate.

For the T3 relation, observe that

LμLμeα=Lμ({eμα if s(μ)=r(α)0 if s(μ)r(α))={eα if s(μ)=r(α)0 if s(μ)r(α)=Ls(μ)eαL_{\mu}^{*}L_{\mu}e_{\alpha}=L_{\mu}^{*}\left(\begin{cases}e_{\mu\alpha}&\text{ if }s(\mu)=r(\alpha)\\ 0&\text{ if }s(\mu)\neq r(\alpha)\end{cases}\right)=\begin{cases}e_{\alpha}&\text{ if }s(\mu)=r(\alpha)\\ 0&\text{ if }s(\mu)\neq r(\alpha)\end{cases}=L_{s(\mu)}e_{\alpha}

as desired.

For the T4 relation, first note that

LμLμeα={eα if αμΛ0 if αμΛ.L_{\mu}L_{\mu}^{*}e_{\alpha}=\begin{cases}e_{\alpha}&\text{ if }\alpha\in\mu\Lambda\\ 0&\text{ if }\alpha\not\in\mu\Lambda\end{cases}.

Thus

LμLμLνLνeα={eα if αμΛνΛ0 if αμΛνΛ.L_{\mu}L_{\mu}^{*}L_{\nu}L_{\nu}^{*}e_{\alpha}=\begin{cases}e_{\alpha}&\text{ if }\alpha\in\mu\Lambda\cap\nu\Lambda\\ 0&\text{ if }\alpha\not\in\mu\Lambda\cap\nu\Lambda\end{cases}.

But since μΛνΛ=λMCE(μ,ν)λΛ\mu\Lambda\cap\nu\Lambda=\bigsqcup_{\lambda\in MCE(\mu,\nu)}\lambda\Lambda by Lemma 2.38(4), then if αμΛνΛ\alpha\in\mu\Lambda\cap\nu\Lambda there exists a unique λ0MCE(μ,ν)\lambda_{0}\in MCE(\mu,\nu) with αλΛ\alpha\in\lambda\Lambda, so Lλ0Lλ0eα=eαL_{\lambda_{0}}L_{\lambda_{0}}^{*}e_{\alpha}=e_{\alpha}, and LλLλeα=0L_{\lambda}L_{\lambda}^{*}e_{\alpha}=0 for λMCE(μ,ν)λ0\lambda\in MCE(\mu,\nu)\setminus\lambda_{0}. Thus

λMCE(μ,ν)LλLλeα=eα\displaystyle\sum_{\lambda\in MCE(\mu,\nu)}L_{\lambda}L_{\lambda}^{*}e_{\alpha}=e_{\alpha}

when αμΛνΛ\alpha\in\mu\Lambda\cap\nu\Lambda, so LμLμLνLν=λMCE(μ,ν)LλLλL_{\mu}L_{\mu}^{*}L_{\nu}L_{\nu}^{*}=\displaystyle\sum_{\lambda\in MCE(\mu,\nu)}L_{\lambda}L_{\lambda}^{*} as desired.

Thus {Lμ}μΛ\{L_{\mu}\}_{\mu\in\Lambda} is a representation, as desired.

There are several additional properties that we will sometimes ask of our representation. In order, we will ask for our generators to be nonzero, that our representation preserve its knowledge of the PP-graded structure through a certain coaction called a gauge coaction, and that our representation be “tight” in the sense of [8]. More precisely:

Definition 2.44.

Let (G,P)(G,P) be a WQLO group and Λ\Lambda a finitely aligned PP-graph. We say that a representation tt of Λ\Lambda

  1. a.

    is Λ\Lambda-faithful if each tλt_{\lambda} is nonzero.

  2. b.

    has a gauge coaction if there is a GG-coaction δ\delta on C(t)C^{*}(t) such that

    δ(tλ)=tλUd(λ)\delta(t_{\lambda})=t_{\lambda}\otimes U_{d(\lambda)}

    for all λΛ\lambda\in\Lambda.

  3. c.

    is tight if whenever μΛ\mu\in\Lambda and EμΛE\subset\mu\Lambda is finite and exhaustive for μΛ\mu\Lambda, we have

    αE(tμtμtαtα)=0.\displaystyle\prod_{\alpha\in E}(t_{\mu}t_{\mu}^{*}-t_{\alpha}t_{\alpha}^{*})=0.

    (Recall that EE is exhaustive for μΛ\mu\Lambda if for all νμΛ\nu\in\mu\Lambda, there is an αE\alpha\in E such that α\alpha and ν\nu have a common extension.) This terminology is motivated by [8], which defines a tight representation of a semilattice. We show in Appendix 1 that this notion of tight is equivalent to the one in [8].

  4. d.

    canonically covers another representation ss if there is a (necessarily surjective) *-homomorphism from C(t)C^{*}(t) to C(s)C^{*}(s) given by tλsλt_{\lambda}\mapsto s_{\lambda}. In such a case, we will write πst:C(t)C(s)\pi_{s}^{t}:C^{*}(t)\rightarrow C^{*}(s) to denote the *-homomorphism, which we call the canonical covering. The notation πst\pi_{s}^{t} should hopefully be suggestive of tt “covering” ss.

  5. e.

    is canonically isomorphic to another representation ss if there is an isomorphism between C(s)C^{*}(s) and C(t)C^{*}(t) given by tλsλt_{\lambda}\mapsto s_{\lambda} (i.e. if ss and tt canonically cover each other).

It should be noted that the conditions (T1)-(T4) and tightness are all polynomial relations of the form suitable for Lemma 2.9 (finite-alignment being necessary in the case of (T4)). However, Λ\Lambda-faithfulness and the existence of a gauge coaction are not polynomial relators in an obvious way.

Remark 2.45.

The left-regular representation LL from Example 2.43 is Λ\Lambda-faithful.

However, this representation is as far from tight as possible: given μΛ\mu\in\Lambda and EμΛE\subset\mu\Lambda which is finite and exhaustive for μΛ\mu\Lambda, αE(LμLμLαLα)=0\displaystyle\prod_{\alpha\in E}(L_{\mu}L_{\mu}^{*}-L_{\alpha}L_{\alpha}^{*})=0 if and only if μE\mu\in E (i.e. when one of the terms of the product is 0). To see this, note that if αμΛμ\alpha\in\mu\Lambda\setminus\mu, then LαLαeμ=0L_{\alpha}L_{\alpha}^{*}e_{\mu}=0, so (LμLμLαLα)eμ=eμ(L_{\mu}L_{\mu}^{*}-L_{\alpha}L_{\alpha}^{*})e_{\mu}=e_{\mu}, and thus taking the product over αEμΛμ\alpha\in E\subset\mu\Lambda\setminus\mu, we get that αE(LμLμLαLα)eμ=eμ0\displaystyle\prod_{\alpha\in E}(L_{\mu}L_{\mu}^{*}-L_{\alpha}L_{\alpha}^{*})e_{\mu}=e_{\mu}\neq 0.

We will show in Corollary 4.38 that LL has a gauge coaction if (G,P)(G,P) has a strong reduction to an amenable group.

Lemma 2.46.

Let (G,P)(G,P) be a WQLO group, and let Λ\Lambda be a finitely aligned PP-graph. One may apply Lemma 2.9 to the relators (T1)-(T4), so there is a representation of Λ\Lambda which is universal.

Proof.

Observe that (T1) and (T3) together imply that the generators are partial isometries. Also, all of the relators are polynomials in the generators (in the case of (T4), we must note that the expression is finite since the graph is finitely aligned). Thus there is a universal representation of Λ\Lambda. ∎

Definition 2.47.

We call the universal representation of a PP-graph Λ\Lambda the Toeplitz(-Cuntz-Krieger) representation of Λ\Lambda. We will use 𝒯\mathcal{T} to denote this representation, and write 𝒯C(Λ)\mathcal{T}C^{*}(\Lambda) or C(𝒯)C^{*}(\mathcal{T}) for the algebra generated by it which we call the Toeplitz(-Cuntz-Krieger) algebra of the graph.

Lemma 2.48.

Let (G,P)(G,P) be a WQLO group, and let Λ\Lambda be a finitely aligned PP-graph.

  1. 1.

    Let Rep(Λ)Rep(\Lambda) denote the set of all representations of Λ\Lambda up to canonical isomorphism. Then Rep(Λ)Rep(\Lambda) can be given a partial ordering rep\geq_{rep} by saying that srepts\geq_{rep}t if and only if ss canonically covers tt (and tst\cong s if and only if they are canonically isomorphic).

  2. 2.

    There is a bijection Ψ\Psi from Rep(Λ)Rep(\Lambda) to ideals in 𝒯C(Λ)\mathcal{T}C^{*}(\Lambda) given by tkerπt𝒯t\mapsto\ker\pi_{t}^{\mathcal{T}}, which is order-reversing in the sense that trepsΨ(t)Ψ(s)t\leq_{rep}s\iff\Psi(t)\supseteq\Psi(s).

Proof.

For (1), we must check the partial order properties of reflexivity, transitivity, and antisymmetry.

Certainly each representation canonically covers itself by the identity map, so rep\leq_{rep} is reflexive.

If trepst\leq_{rep}s and sreprs\leq_{rep}r, then there are canonical coverings πsr:C(r)C(s)\pi_{s}^{r}:C^{*}(r)\rightarrow C^{*}(s) and πts:C(s)C(t)\pi_{t}^{s}:C^{*}(s)\rightarrow C^{*}(t). Then the composition πtr:=πtsπsr\pi_{t}^{r}:=\pi_{t}^{s}\circ\pi_{s}^{r} is a canonical covering of tt by rr, so treprt\leq_{rep}r, giving transitivity.

If trepst\leq_{rep}s and srepts\leq_{rep}t, then there are canonical coverings πts:C(s)C(t)\pi_{t}^{s}:C^{*}(s)\rightarrow C^{*}(t) and πst:C(t)C(s)\pi_{s}^{t}:C^{*}(t)\rightarrow C^{*}(s). Composing these maps in either direction shows that they fix each generator, and therefore are the identity maps. Therefore πts\pi_{t}^{s} and πst\pi_{s}^{t} are inverses and hence canonical isomorphisms.

Thus rep\leq_{rep} is indeed a partial order, as desired.

For (3), it is certainly true that Ψ\Psi is a well-defined function from Rep(Λ)Rep(\Lambda) to {ideals in 𝒯C(Λ)}\{\text{ideals in }\mathcal{T}C^{*}(\Lambda)\}.

We will first check that if s,tRep(Λ)s,t\in Rep(\Lambda), then trepsΨ(t)Ψ(s)t\leq_{rep}s\iff\Psi(t)\supseteq\Psi(s). If trepst\leq_{rep}s, then there is a canonical covering πts:C(s)C(t)\pi^{s}_{t}:C^{*}(s)\rightarrow C^{*}(t), and since πtsπs𝒯=πt𝒯\pi^{s}_{t}\circ\pi^{\mathcal{T}}_{s}=\pi^{\mathcal{T}}_{t}, then Ψ(t)=kerπt𝒯πs𝒯=Ψ(s)\Psi(t)=\ker\pi^{\mathcal{T}}_{t}\supseteq\pi^{\mathcal{T}}_{s}=\Psi(s). Conversely, if Ψ(t)Ψ(s)\Psi(t)\supseteq\Psi(s), then by Lemma 2.8 applied to πt𝒯\pi^{\mathcal{T}}_{t} and πs𝒯\pi^{\mathcal{T}}_{s}, there is a *-homomorphism πts:C(s)C(t)\pi_{t}^{s}:C^{*}(s)\rightarrow C^{*}(t) with πtsπs𝒯=πt𝒯\pi_{t}^{s}\circ\pi^{\mathcal{T}}_{s}=\pi^{\mathcal{T}}_{t}so πts\pi_{t}^{s} is a canonical covering and trepst\leq_{rep}s. This completes the proof that Ψ\Psi is order-reversing.

Now to show that Ψ\Psi is a bijection, if Ψ(s)=Ψ(t)\Psi(s)=\Psi(t), then by the above srepts\leq_{rep}t and trepst\leq_{rep}s, so sts\cong t and hence Ψ\Psi is injective. For surjectivity, if J𝒯C(Λ)J\lhd\mathcal{T}C^{*}(\Lambda) is an ideal, represent Λ\Lambda by λ𝒯λ+J\lambda\mapsto\mathcal{T}_{\lambda}+J. Then this representation has kernel JJ, so Ψ\Psi is surjective, as desired. ∎

As a consequence of this, representations have the order structure of the family of ideals of an algebra, and in particular have a lattice order (where meet and join are represented by intersection and addition of ideals).

Remark 2.49.

Λ\Lambda-faithfulness is a “largeness” condition in that if sts\leq t and ss is Λ\Lambda-faithful, then so is tt.

Tightness is a “smallness” condition in that if sts\leq t and tt is tight, then so is ss.

2.5.1 Gauge coactions and Coactionization

It may not be obvious what role “has a gauge coaction” plays: is it saying a representation is “large” or “small”?

The following proposition is a tangled web of results, but it shows that the existence of a gauge coaction is properly considered a “largeness” condition: any representation can be “lifted” (1) to have a gauge coaction (4), the lifted version covers the original (2) and is the smallest gauge coacting representation to do so (6), and they are isomorphic if and only if the original had a gauge coaction (5). We also simplify the notion of a gauge coaction by showing that any homomorphism satisfying tλtλUd(λ)t_{\lambda}\mapsto t_{\lambda}\otimes U_{d(\lambda)} is automatically a gauge coaction (3), and show that a representation is Λ\Lambda-faithful or tight if and only if its coactionization is (7).

Proposition 2.50.

Let (G,P)(G,P) be a WQLO group, and let Λ\Lambda be a finitely aligned PP-graph. Let tt be a representation of Λ\Lambda.

  1. 1.

    There is a representation tt^{\prime} of Λ\Lambda given by tλ=tλUd(λ)t^{\prime}_{\lambda}=t_{\lambda}\otimes U_{d(\lambda)}.

  2. 2.

    There is a canonical covering ttt^{\prime}\mapsto t.

  3. 3.

    If there is a *-homomorphism δ:C(t)C(t)C(G)\delta:C^{*}(t)\rightarrow C^{*}(t)\otimes C^{*}(G) satisfying δ(tλ)=tλ\delta(t_{\lambda})=t^{\prime}_{\lambda}, then δ\delta is a coaction, and hence a gauge coaction.

  4. 4.

    tt^{\prime} has a gauge coaction.

  5. 5.

    tt is canonically isomorphic to tt^{\prime} if and only if tt has a gauge coaction.

  6. 6.

    If sts\leq t, then sts^{\prime}\leq t^{\prime}, and in particular for any representation ss, ss^{\prime} is the smallest gauge coacting representation that covers ss.

  7. 7.

    tt^{\prime} is Λ\Lambda-faithful (respectively, tight) if and only if tt is.

Due to these properties, we believe that tt^{\prime} deserves to be called the coactionization of tt.

Proof.

The proof of (1) is a routine confirmation of the T1-T4 relators, which we include here. We have:

T1: tvtwt_{v}^{\prime}t_{w}^{\prime} = (tvUe)(twUe)=tvtwUe=δv,wtvUe=δv,wtv(t_{v}\otimes U_{e})(t_{w}\otimes U_{e})=t_{v}t_{w}\otimes U_{e}=\delta_{v,w}t_{v}\otimes U_{e}=\delta_{v,w}t_{v}^{\prime}
T2: tμtνt_{\mu}^{\prime}t_{\nu}^{\prime} = (tμUd(μ))(tνUd(ν))=(tμtν)Ud(μ)Ud(ν)(t_{\mu}\otimes U_{d(\mu)})(t_{\nu}\otimes U_{d(\nu)})=(t_{\mu}t_{\nu})\otimes U_{d(\mu)}U_{d(\nu)}
= tμνUd(μν)=tμνt_{\mu\nu}\otimes U_{d(\mu\nu)}=t_{\mu\nu}^{\prime}
T3: tμtμ{t_{\mu}^{\prime}}^{*}t_{\mu}^{\prime} = (tμUd(μ))(tμUd(μ))=tμtμUd(μ)Ud(μ)(t_{\mu}\otimes U_{d(\mu)})^{*}(t_{\mu}\otimes U_{d(\mu)})=t_{\mu}^{*}t_{\mu}\otimes U_{d(\mu)}^{*}U_{d(\mu)}
= ts(μ)Ue=ts(μ)t_{s(\mu)}\otimes U_{e}=t_{s(\mu)}^{\prime}
T4: tνtνtμtμt_{\nu}^{\prime}{t_{\nu}^{\prime}}^{*}t_{\mu}^{\prime}t_{\mu}^{\prime*} = (tνUd(ν))(tνUd(ν))(tμUd(μ))(tμUd(μ))(t_{\nu}\otimes U_{d(\nu)})(t_{\nu}\otimes U_{d(\nu)})^{*}(t_{\mu}\otimes U_{d(\mu)})(t_{\mu}\otimes U_{d(\mu)})^{*}
= (tνtνtμtμ)Ue=λMCE(μ,ν)tλtλUe(t_{\nu}t_{\nu}^{*}t_{\mu}t_{\mu}^{*})\otimes U_{e}=\displaystyle\sum_{\lambda\in MCE(\mu,\nu)}t_{\lambda}t_{\lambda}^{*}\otimes U_{e}
= λMCE(μ,ν)tλtλ\displaystyle\sum_{\lambda\in MCE(\mu,\nu)}t_{\lambda}^{\prime}{t_{\lambda}^{\prime}}^{*}

all as desired.

For (2), let VV denote the trivial representation of GG (given by Vg=1V_{g}=1\in\mathbb{C} for all gGg\in G). Then by the universal property of C(G)C^{*}(G), there is a covering map πVU:C(G)\pi^{U}_{V}:C^{*}(G)\rightarrow\mathbb{C} given by πVU(Ug)=Vg=1\pi^{U}_{V}(U_{g})=V_{g}=1.

Now by Lemma 2.14, there is a map idC(t)πVU:C(t)C(G)C(t)\operatorname{id}_{C^{*}(t)}\otimes\pi^{U}_{V}:C^{*}(t)\otimes C^{*}(G)\rightarrow C^{*}(t)\otimes\mathbb{C} satisfying

(idC(t)πVU)(tλ)\displaystyle(\operatorname{id}_{C^{*}(t)}\otimes\pi^{U}_{V})(t^{\prime}_{\lambda}) =\displaystyle= (idC(t)πVU)(tλUd(λ))\displaystyle(\operatorname{id}_{C^{*}(t)}\otimes\pi^{U}_{V})(t_{\lambda}\otimes U_{d(\lambda)})
=\displaystyle= tλ1.\displaystyle t_{\lambda}\otimes 1.

By identifying C(t)C^{*}(t) with C(t)C^{*}(t)\otimes\mathbb{C} in the natural way, this gives the desired canonical covering.

For (3), we must check that δ\delta is injective, nondegenerate, and satisfies the coaction identity of Definition 2.25. To this end, note that if πtt\pi_{t}^{t^{\prime}} is the canonical covering from the previous part, then πttδ=idC(t)\pi_{t}^{t^{\prime}}\circ\delta=\operatorname{id}_{C^{*}(t)}, so δ\delta must be injective. To check nondegeneracy, we must check that

span¯[δ(C(t))(C(t)C(G))]=C(t)C(G).\overline{\operatorname{span}}\left[\delta(C^{*}(t))(C^{*}(t)\otimes C^{*}(G))\right]=C^{*}(t)\otimes C^{*}(G).

To this end, note that for all gGg\in G and μ,νΛ\mu,\nu\in\Lambda,

δ(ts(μ))(tμtνUg)=(ts(μ)U1)(tμtνUg)=tμtνUg.\delta(t_{s(\mu)})(t_{\mu}t_{\nu}^{*}\otimes U_{g})=(t_{s(\mu)}\otimes U_{1})(t_{\mu}t_{\nu}^{*}\otimes U_{g})=t_{\mu}t_{\nu}^{*}\otimes U_{g}.

That is, {tμtνUg:μ,νΛ,gG}δ(C(t))(C(t)C(G))\{t_{\mu}t_{\nu}^{*}\otimes U_{g}:\mu,\nu\in\Lambda,g\in G\}\subseteq\delta(C^{*}(t))(C^{*}(t)\otimes C^{*}(G)). But since C(t)=span¯{tμtν:μ,νΛ}C^{*}(t)=\overline{\operatorname{span}}\{t_{\mu}t_{\nu}^{*}:\mu,\nu\in\Lambda\} and C(G)=span¯{Ug}gGC^{*}(G)=\overline{\operatorname{span}}\{U_{g}\}_{g\in G}, then these simple tensors have dense span C(t)C(G)C^{*}(t)\otimes C^{*}(G). Thus

C(t)C(G)=span¯{tμtνUg}μ,νΛ,gG=span¯[δ(C(t))(C(t)C(G))]C^{*}(t)\otimes C^{*}(G)=\overline{\operatorname{span}}\{t_{\mu}t_{\nu}^{*}\otimes U_{g}\}_{\mu,\nu\in\Lambda,g\in G}=\overline{\operatorname{span}}\left[\delta(C^{*}(t))(C^{*}(t)\otimes C^{*}(G))\right]

as desired.

Finally, it suffices to check the coaction identity on each generator tμt_{\mu}. To this end,

((δidG)δ)(tμ)\displaystyle((\delta\otimes\operatorname{id}_{G})\circ\delta)(t_{\mu}) =\displaystyle= (δidG)(tμUd(μ))\displaystyle(\delta\otimes\operatorname{id}_{G})(t_{\mu}\otimes U_{d(\mu)})
=\displaystyle= tμUd(μ)Ud(μ)\displaystyle t_{\mu}\otimes U_{d(\mu)}\otimes U_{d(\mu)}
=\displaystyle= tμδG(Ud(μ))\displaystyle t_{\mu}\otimes\delta_{G}(U_{d(\mu)})
=\displaystyle= ((idC(t)δG)(tμUd(μ))\displaystyle((\operatorname{id}_{C^{*}(t)}\otimes\delta_{G})(t_{\mu}\otimes U_{d(\mu)})
=\displaystyle= ((idC(t)δG)δ)(tμ)\displaystyle((\operatorname{id}_{C^{*}(t)}\otimes\delta_{G})\circ\delta)(t_{\mu})

as desired.

For (4), recall that there is a coaction δG:C(G)C(G)C(G)\delta_{G}:C^{*}(G)\rightarrow C^{*}(G)\otimes C^{*}(G) given by UgUgUgU_{g}\mapsto U_{g}\otimes U_{g}. By Lemma 2.14 we may define δ¯:C(t)C(G)C(t)C(G)C(G)\bar{\delta}:C^{*}(t)\otimes C^{*}(G)\rightarrow C^{*}(t)\otimes C^{*}(G)\otimes C^{*}(G) by δ¯=idC(t)δG\bar{\delta}=\operatorname{id}_{C^{*}(t)}\otimes\delta_{G}. Let δ=δ¯C(t)\delta=\bar{\delta}\mid_{C^{*}(t^{\prime})}. Then for λΛ\lambda\in\Lambda,

δ(t)=δ(tλUd(λ))=tλδG(Ud(λ))=tλUd(λ)Ud(λ)=tλUd(λ).\delta(t^{\prime})=\delta(t_{\lambda}\otimes U_{d(\lambda)})=t_{\lambda}\otimes\delta_{G}(U_{d(\lambda)})=t_{\lambda}\otimes U_{d(\lambda)}\otimes U_{d(\lambda)}=t_{\lambda}^{\prime}\otimes U_{d(\lambda)}.

Thus by (3), δ\delta is a gauge coaction of C(t)C^{*}(t^{\prime}).

For (5), if tt is canonically isomorphic to tt^{\prime}, then there is a homomorphism between their CC^{*}-algebras given by tλtλ=tλUd(λ)t_{\lambda}\mapsto t^{\prime}_{\lambda}=t_{\lambda}\otimes U_{d(\lambda)}. By (3) this is precisely a gauge coaction.

For (6), if sts\leq t, recall that πst\pi_{s}^{t} denotes the canonical covering. Then by Lemma 2.14 there is a *-homomorphism πst:C(t)C(s)\pi_{s^{\prime}}^{t^{\prime}}:C^{*}(t^{\prime})\rightarrow C^{*}(s^{\prime}) given by πst=πstidC(G)\pi_{s^{\prime}}^{t^{\prime}}=\pi_{s}^{t}\otimes\operatorname{id}_{C^{*}(G)}, and it is immediate that this is a canonical covering. For the “in particular”, if sts\leq t and tt has a gauge coaction, then ttt\cong t^{\prime} by (5), so stts^{\prime}\leq t^{\prime}\cong t, and thus sts^{\prime}\leq t, as desired.

For (7), observe that for a fixed λΛ\lambda\in\Lambda,

tλ2=tλUd(λ)2=tλtλ1=tλtλ=tλ2\left\|{t_{\lambda}^{\prime}}\right\|^{2}=\left\|{t_{\lambda}\otimes U_{d(\lambda)}}\right\|^{2}=\left\|{t_{\lambda}^{*}t_{\lambda}\otimes 1}\right\|=\left\|{t_{\lambda}^{*}t_{\lambda}}\right\|=\left\|{t_{\lambda}}\right\|^{2}

so any tλt_{\lambda}^{\prime} is 0 if and only if tλt_{\lambda} is 0. Thus tt is Λ\Lambda-faithful if and only if tt^{\prime} is Λ\Lambda-faithful.

For tightness, fix a μΛ\mu\in\Lambda and finite EμΛE\subset\mu\Lambda which is exhaustive for μΛ\mu\Lambda. Then

αEtμtμtαtα\displaystyle\left\|{\displaystyle\prod_{\alpha\in E}t_{\mu}^{\prime}{t_{\mu}^{\prime}}^{*}-t_{\alpha}^{\prime}{t_{\alpha}^{\prime}}^{*}}\right\| =\displaystyle= (αEtμtμtαtα)1\displaystyle\left\|{\left(\displaystyle\prod_{\alpha\in E}t_{\mu}{t_{\mu}}^{*}-t_{\alpha}{t_{\alpha}}^{*}\right)\otimes 1}\right\|
=\displaystyle= αEtμtμtαtα\displaystyle\left\|{\displaystyle\prod_{\alpha\in E}t_{\mu}{t_{\mu}}^{*}-t_{\alpha}{t_{\alpha}}^{*}}\right\|

and in particular αEtμtμtαtα=0\displaystyle\prod_{\alpha\in E}t_{\mu}^{\prime}{t_{\mu}^{\prime}}^{*}-t_{\alpha}^{\prime}{t_{\alpha}^{\prime}}^{*}=0 if and only if αEtμtμtαtα=0\displaystyle\prod_{\alpha\in E}t_{\mu}{t_{\mu}}^{*}-t_{\alpha}{t_{\alpha}}^{*}=0, as desired. ∎

Finally, note that property (3) says that the notion of coaction is dramatically simplified when considering just gauge coactions. Indeed, the necessary and sufficient condition to having a gauge coaction is that the map tλtλUd(λ)t_{\lambda}\mapsto t_{\lambda}\otimes U_{d(\lambda)} extends to a *-homomorphism on C(t)C^{*}(t).

2.6 The Classical Gauge-Invariant Uniqueness Theorem

In this section we will introduce the gauge-invariant uniqueness theorem for directed graphs (what we would call \mathbb{N}-graphs), and explain how we hope to generalize it.

The following definitions and theorems are adapted from Chapters 1 and 2 of [18].

Definition 2.51.

A directed graph is a tuple E=(E0,E1,r,s)E=(E^{0},E^{1},r,s) where E0E^{0} and E1E^{1} are countable sets, and r,s:E1E0r,s:E^{1}\rightarrow E^{0} are functions. We think of E0E^{0} as the set of vertices in our directed graph, E1E^{1} as our set of edges, and rr and ss as the range and source maps for E1E^{1}. We say a directed graph is row-finite if for each vVv\in V, |r1(v)|<|r^{-1}(v)|<\infty.

A Cuntz-Krieger EE-family of a row-finite directed graph (E0,E1,r,s)(E^{0},E^{1},r,s) is a collection of operators {Pv}vE0{Se}eE1\{P_{v}\}_{v\in E^{0}}\cup\{S_{e}\}_{e\in E^{1}} in a CC^{*}-algebra such that {Pv}vE0\{P_{v}\}_{v\in E^{0}} are a family of pairwise orthogonal projections, and

  1. (CK1) Ps(e)=SeSeP_{s(e)}=S_{e}^{*}S_{e} for all eE1e\in E^{1} and

  2. (CK2) Pv=eE1r(e)=vSeSeP_{v}=\displaystyle\sum_{\begin{subarray}{c}e\in E^{1}\\ r(e)=v\end{subarray}}S_{e}S_{e}^{*} whenever r1(v)r^{-1}(v) is nonempty.

The reader may wish to compare this definition to the definition of a representation of a PP-graph (Definition 2.41). Having done so, there are a few differences that we will wish to reconcile.

Remark 2.52.

Given a directed graph (E0,E1,r,s)(E^{0},E^{1},r,s), for n>1n>1 we let En={(e1,e2,,en):s(ei)=r(ei+1) for 1i<n}E^{n}=\{(e_{1},e_{2},...,e_{n}):s(e_{i})=r(e_{i+1})\text{ for }1\leq i<n\}. Then E=n0EnE^{*}=\bigcup_{n\geq 0}E^{n} forms an \mathbb{N}-graph in a natural way, where the operation is concatenation of tuples, and d(e1,,en)=nd(e_{1},...,e_{n})=n. This \mathbb{N}-graph is always finitely aligned, since for any \mathbb{N}-graph,

MCE(α,β)={{α} if βα{β} if αβ otherwiseMCE(\alpha,\beta)=\begin{cases}\{\alpha\}&\text{ if }\beta\leq\alpha\\ \{\beta\}&\text{ if }\alpha\leq\beta\\ \emptyset&\text{ otherwise}\end{cases}

so |MCE(α,β)|1<|MCE(\alpha,\beta)|\leq 1<\infty.

Any Cuntz-Krieger EE-family extends to a representation of EE^{*} by S(e1,e2,,en)=Se1Se2SenS_{(e_{1},e_{2},...,e_{n})}=S_{e_{1}}S_{e_{2}}...S_{e_{n}}. In proving this, the (T1) and (T2) relators are immediate, the (T3) relator follows rapidly from (CK1), and the (T4) relator follows from the expression above for MCE(α,β)MCE(\alpha,\beta).

One should note that the (CK2) condition is analogous to tightness, since for any vertex vv, r1(v)r^{-1}(v) is finite and exhaustive for vΛv\Lambda, and since {SeSe}er1(v)\{S_{e}S_{e}^{*}\}_{e\in r^{-1}(v)} are pairwise orthogonal projections, then

er1(v)(PvPvSeSe)=0 if and only if Pv=eE1r(e)=vSeSe.\displaystyle\prod_{e\in r^{-1}(v)}(P_{v}P_{v}^{*}-S_{e}S_{e}^{*})=0\text{ if and only if }P_{v}=\displaystyle\sum_{\begin{subarray}{c}e\in E^{1}\\ r(e)=v\end{subarray}}S_{e}S_{e}^{*}.

In other words, the (CK2) condition is saying that the representation is tight in the case that μΛ0\mu\in\Lambda^{0} and EΛ1E\subset\Lambda^{1}. Since every path in an \mathbb{N}-graph can be uniquely factored into paths of length 1, tightness in all cases is equivalent to this.

Due to this last point, readers should be aware that in the literature representations of directed graphs (and higher rank graphs) are what we would call tight representations.

In the notation of [18], C(E)C^{*}(E) denotes the universal tight representation of EE, and if {Tv}vE0{Qe}eE1\{T_{v}\}_{v\in E^{0}}\cup\{Q_{e}\}_{e\in E^{1}} is a Cuntz-Krieger EE-family, then πT,Q:C(E)C(T,Q)\pi_{T,Q}:C^{*}(E)\rightarrow C^{*}(T,Q) denotes the canonical covering.

The foundational result we wish to generalize is this, as stated in [18, Theorem 2.2]:

Theorem 2.53 (Gauge Invariant Uniqueness Theorem for Graphs).

Let E=(E0,E1,r,s)E=(E^{0},E^{1},r,s) be a row-finite directed graph, and suppose that {Tv}vE0{Qe}eE1\{T_{v}\}_{v\in E^{0}}\cup\{Q_{e}\}_{e\in E^{1}} is a Cuntz-Krieger EE-family in a CC^{*}-algebra BB with each Qv0Q_{v}\neq 0. If there is a continuous action β:𝕋AutB\beta:\mathbb{T}\rightarrow\operatorname{Aut}B such that βz(Te)=zTe\beta_{z}(T_{e})=zT_{e} for every eE1e\in E^{1} and βz(Qv)=Qv\beta_{z}(Q_{v})=Q_{v} for every vE0v\in E^{0}, then πT,Q\pi_{T,Q} is an isomorphism of C(E)C^{*}(E) onto C(T,Q)C^{*}(T,Q).

Remark 2.54.

The reader may recognize many of the hypotheses of Theorem 2.53 from our earlier list of terminology for representations (Definition 2.44). The statement Qv0Q_{v}\neq 0 for vE0v\in E^{0} implies that Tμ0T_{\mu}\neq 0 for all μE\mu\in E^{*}, so it is equivalent to our Λ\Lambda-faithfulness. The continuous action β\beta is called a gauge action and is equivalent by Lemma 2.34 to a gauge coaction. Also recall that tightness is a built-in hypothesis in the representations considered in [18].

Therefore, we may restate Theorem 2.53 like so: for any \mathbb{N}-graph, there is exactly one Λ\Lambda-faithful, tight, gauge coacting representation up to canonical isomorphism.

What we mean by a gauge invariant uniqueness theorem for PP-graphs is a similar statement: for any PP-graph, there is exactly one Λ\Lambda-faithful, tight, gauge coacting representation up to canonical isomorphism.

We have already seen that in generalizing from \mathbb{N}-graphs to PP graphs, we needed an additional hypothesis (finite alignment). We will see in Lemma 4.24 that such a gauge invariant uniqueness theorem need not be true in general, so another hypothesis is needed. We spend Chapter 3 developing this hypothesis on (G,P)(G,P), and then in Chapter 4 prove a gauge invariant uniqueness theorem for PP-graphs that satisfy this new hypothesis.

3 Reductions of Ordered Groups

In this section we will develop the notion of a reduction of an ordered group. Roughly speaking, a reduction occurs when a positive cone PP in a group GG does not carry “enough” information about the group, and one is able to replace (G,P)(G,P) with some (H,Q)(H,Q) while preserving the essential properties that will be required for a PP-graph (namely, the structure of intervals [1,p][1,p]). A special case of a reduction (called a “strong reduction”) is when one can embed the positive cone PP in another group HH such that this embedding extends to an order homomorphism from (G,P)(G,P) to (H,P)(H,P).

3.1 Definition and Basic Properties

Definition 3.1.

An order homomorphism φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) is a reduction if for all pPp\in P, φ\varphi is a bijection between the interval [1,p]:={xP:1xp}[1,p]:=\{x\in P:1\leq x\leq p\} and the interval [1,φ(p)]:={yQ:1yφ(p)}[1,\varphi(p)]:=\{y\in Q:1\leq y\leq\varphi(p)\}.

We say that φ\varphi is a strong reduction if φP:PQ\varphi\mid_{P}:P\rightarrow Q is a bijection between PP and QQ.

We say (G,P)(G,P) has a (strong) reduction to an amenable group or (strongly) reduces to an amenable group if there is an amenable group HH, a positive cone QHQ\subset H, and a (strong) reduction φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q).

Intuitively, a reduction is a map that preserves the order structure of (G,P)(G,P) “locally”, meaning on every interval. For strong reductions, the reader should think of them as arising from embedding the positive cone PP in two distinct groups (G,P)(G,P) and (H,P)(H,P) such that there is a homomorphism from one to the other taking PP to itself bijectively.

Example 3.2.

Let G=a,bG=\langle a,b\rangle, the free group on two generators. If P={a,b}P=\{a,b\}^{*}, then define φ:(G,P)(,)\varphi:(G,P)\rightarrow(\mathbb{Z},\mathbb{N}) by φ(a)=φ(b)=1\varphi(a)=\varphi(b)=1, which is an order homomorphism.

Then φ\varphi is a reduction since, for example, it maps the interval [e,abba]={e,a,ab,abb,abba}[e,abba]=\{e,a,ab,abb,abba\} bijectively onto the interval [φ(e),φ(abba)]=[0,4]={0,1,2,3,4}[\varphi(e),\varphi(abba)]=[0,4]=\{0,1,2,3,4\}. Since \mathbb{Z} is amenable, then this shows that (G,P)(G,P) reduces to an amenable group.

However, φ\varphi is not injective even when restricted to PP, as φ(a)=1=φ(b)\varphi(a)=1=\varphi(b), so φ\varphi is not a strong reduction.

It may not be immediately obvious from the definition that a strong reduction is a reduction, so we will prove this as a short lemma:

Lemma 3.3.

Let φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) be a strong reduction. Then it is a reduction.

Proof.

We must show that if φ\varphi is bijective as a map from PP to QQ, then it is bijective from [1,p][1,p] to [1,φ(p)][1,\varphi(p)] for each pPp\in P. Fixing a pPp\in P, since [1,p]P[1,p]\subset P, then φ\varphi is injective on [1,p][1,p], so it suffices to show it is surjective onto [1,φ(p)][1,\varphi(p)].

To this end, fix some q1[1,φ(p)]q_{1}\in[1,\varphi(p)]. Since q1φ(p)q_{1}\leq\varphi(p), there is a q2Qq_{2}\in Q such that q1q2=φ(p)q_{1}q_{2}=\varphi(p). Now since q1,q2Qq_{1},q_{2}\in Q and φ\varphi is a bijection from PP to QQ, there is a p1,p2Pp_{1},p_{2}\in P such that φ(p1)=q1\varphi(p_{1})=q_{1} and φ(p2)=q2\varphi(p_{2})=q_{2}. Then φ(p)=q1q2=φ(p1)φ(p2)=φ(p1p2)\varphi(p)=q_{1}q_{2}=\varphi(p_{1})\varphi(p_{2})=\varphi(p_{1}p_{2}). Since p1p2Pp_{1}p_{2}\in P and φ\varphi is injective on PP, then p1p2=pp_{1}p_{2}=p, so p1[1,p]p_{1}\in[1,p], and in particular there is a p1[1,p]p_{1}\in[1,p] such that φ(p1)=q1\varphi(p_{1})=q_{1}. That is, φ\varphi is a bijection from [1,p][1,p] onto [1,φ(p)][1,\varphi(p)], as desired. ∎

We’ll now provide several alternative characterizations of being a reduction and a strong reduction.

Proposition 3.4.

Let φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) be an order homomorphism. Then the following are equivalent:

  1. 1.

    φ\varphi is a reduction in the sense of Definition 3.1.

  2. 2.

    dQP:=φP:PQd^{P}_{Q}:=\varphi\mid_{P}:P\rightarrow Q is a functor that makes PP into a QQ-graph.

  3. 3.

    For all qQ,pPq\in Q,p\in P such that qφ(p)q\leq\varphi(p), there exist unique p1,p2Pp_{1},p_{2}\in P such that p1p2=pp_{1}p_{2}=p and φ(p1)=q\varphi(p_{1})=q.

  4. 4.

    For each nonempty interval [x,y]G[x,y]\subset G, φ\varphi is an order isomorphism between [x,y][x,y] and [φ(x),φ(y)][\varphi(x),\varphi(y)].

  5. 5.

    The following two statements together:

    1. (a)

      Skerφ={1}S\cap\ker\varphi=\{1\} where S=pP[1,p][1,p]1S=\bigcup_{p\in P}[1,p][1,p]^{-1}.

    2. (b)

      For all pPp\in P, φ([1,p])=[1,φ(p)]\varphi([1,p])=[1,\varphi(p)].

Proof.

Fix φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) an order homomorphism. We will prove that (2)(3)(1)(4)(2)(2)\Rightarrow(3)\Rightarrow(1)\Rightarrow(4)\Rightarrow(2) and (1)(5)(1)\iff(5).

(23)(2\Rightarrow 3) Fix qQq\in Q, pPp\in P with qφ(p)=dQP(p)q\leq\varphi(p)=d_{Q}^{P}(p). Then since PP is a QQ-graph, we can uniquely factorize pp as p=p1p2p=p_{1}p_{2} where q=dQP(p1)=φ(p1)q=d^{P}_{Q}(p_{1})=\varphi(p_{1}), as desired.

(31)(3\Rightarrow 1) Fix pPp\in P. By Lemma 2.7, φ([1,p])[1,φ(p)]\varphi([1,p])\subset[1,\varphi(p)]. It then suffices to show that if q[1,φ(p)]q\in[1,\varphi(p)], then there exists a unique s[1,p]s\in[1,p] with φ(s)=q\varphi(s)=q. To this end, if q[1,φ(p)]q\in[1,\varphi(p)], then by (3), there exists p1,p2Pp_{1},p_{2}\in P such that p1p2=pp_{1}p_{2}=p and with φ(p1)=q\varphi(p_{1})=q. Then p1[1,p]p_{1}\in[1,p], so p1p_{1} is such an ss. If there were some p1[1,p]p_{1}^{\prime}\in[1,p] with φ(p1)=φ(p1)\varphi(p_{1})=\varphi(p_{1}^{\prime}), then by the former, there exists p2Pp_{2}^{\prime}\in P with p1p2=pp_{1}^{\prime}p_{2}^{\prime}=p, so by the uniqueness condition of (2), we have p1=p1p_{1}^{\prime}=p_{1} and p2=p2p_{2}^{\prime}=p_{2}, and thus our p1p_{1} is unique, as desired.

(14)(1\Rightarrow 4) Fix x,yGx,y\in G, and suppose the interval [x,y]={zG:xzy}[x,y]=\{z\in G:x\leq z\leq y\} is nonempty. Since this interval is nonempty, then xyx\leq y, so 1x1y1\leq x^{-1}y, and we will write p=x1yPp=x^{-1}y\in P. By left-invariance, xzyx\leq z\leq y if and only if 1x1zx1y1\leq x^{-1}z\leq x^{-1}y, so [x,y]=x[1,x1y]=x[1,p][x,y]=x[1,x^{-1}y]=x[1,p]. By left-invariance, x[1,p]x[1,p] is order-isomorphic to [1,p][1,p], and φ(x)[1,φ(p)]\varphi(x)[1,\varphi(p)] is order-isomorphic to [1,φ(p)][1,\varphi(p)], so it suffices to show that [1,p][1,p] is order-isomorphic to [1,φ(p)][1,\varphi(p)]. By (1), φ\varphi is a bijection between [1,p][1,p] and [1,φ(p)][1,\varphi(p)], so it suffices to show that stφ(s)φ(t)s\leq t\iff\varphi(s)\leq\varphi(t) for s,t[1,p]s,t\in[1,p]. The \Rightarrow direction is immediate from Lemma 2.7. For the other direction, suppose s,t[1,p]s,t\in[1,p] and φ(s)φ(t)\varphi(s)\leq\varphi(t). Then φ(s)[1,φ(t)]\varphi(s)\in[1,\varphi(t)], and by (3), φ\varphi maps [1,t][1,t] surjectively onto [1,φ(t)][1,\varphi(t)], so there exists some s[1,t]s^{\prime}\in[1,t] such that φ(s)=φ(s)\varphi(s^{\prime})=\varphi(s). But s[1,t][1,p]s^{\prime}\in[1,t]\subseteq[1,p], and φ\varphi is injective on [1,p][1,p], so s=ss^{\prime}=s. Thus s[1,t]s\in[1,t], so sts\leq t.

(42)(4\Rightarrow 2) Certainly dQP:=φPd^{P}_{Q}:=\varphi\mid_{P} is a functor from PP to QQ. It then suffices to check unique factorization. To this end, suppose pPp\in P and q1,q2Qq_{1},q_{2}\in Q with dQP(p)=q1q2d_{Q}^{P}(p)=q_{1}q_{2}. Then q1[1,dQP(p)]=[1,φ(p)]q_{1}\in[1,d_{Q}^{P}(p)]=[1,\varphi(p)], so by (4) there is a unique p1[1,p]p_{1}\in[1,p] with q1=φ(p1)=dQP(p1)q_{1}=\varphi(p_{1})=d_{Q}^{P}(p_{1}). Since p1pp_{1}\leq p, then p2:=p11pPp_{2}:=p_{1}^{-1}p\in P. Thus pp is uniquely factorized as p=p1p2p=p_{1}p_{2} with dQP(p1)=q1d_{Q}^{P}(p_{1})=q_{1}.

(15a)(1\Rightarrow 5a) Suppose that sSkerφs\in S\cap\ker\varphi. Then s=qr1s=qr^{-1} where q,r[1,p]q,r\in[1,p] for some pPp\in P. Then φ(r)=φ(s)φ(r)=φ(sr)=φ(q)\varphi(r)=\varphi(s)\varphi(r)=\varphi(sr)=\varphi(q). By (1), φ[1,p]\varphi\mid_{[1,p]} is injective, so q=rq=r and thus s=qr1=1s=qr^{-1}=1, as desired. .

(15b)(1\Rightarrow 5b) is immediate from the fact that φ\varphi is a bijection between the intervals [1,p][1,p] and [1,φ(p)][1,\varphi(p)].

(51)(5\Rightarrow 1) Fix pPp\in P. By Lemma 2.7, φ([1,p])[1,φ(p)]\varphi([1,p])\subseteq[1,\varphi(p)]. It then suffices to show that if q[1,φ(p)]q\in[1,\varphi(p)], then there exists a unique s[1,p]s\in[1,p] with φ(s)=q\varphi(s)=q. By (5b), there exists at least one s[1,p]s\in[1,p] with φ(s)=q\varphi(s)=q. Assume for the sake of contradiction that there were distinct s1,s2[1,p]s_{1},s_{2}\in[1,p] with φ(s1)=q=φ(s2)\varphi(s_{1})=q=\varphi(s_{2}). Then s1s21kerφs_{1}s_{2}^{-1}\in\ker\varphi. But since s1s21[1,p][1,p]1Ss_{1}s_{2}^{-1}\in[1,p][1,p]^{-1}\subseteq S, then by (5a) we have s1s21=1s_{1}s_{2}^{-1}=1, so s1=s2s_{1}=s_{2}, a contradiction of distinctness. Thus φ\varphi is bijective as a map from [1,p][1,p] to [1,φ(p)][1,\varphi(p)].

Lemma 3.5.

Let (G,P)(G,P), (H,Q)(H,Q) be ordered groups, and φ:GH\varphi:G\rightarrow H a group homomorphism. The following are equivalent:

  1. 1.

    φ\varphi is a strong reduction in the sense of Definition 3.1.

  2. 2.

    φP\varphi\mid_{P} is an order isomorphism from PP to QQ.

Proof.

(21)(2\Rightarrow 1) is immediate.

(12)(1\Rightarrow 2) Since φP\varphi\mid_{P} is already a bijection, we must show that for p1,p2Pp_{1},p_{2}\in P, p1p2φ(p1)φ(p2)p_{1}\leq p_{2}\iff\varphi(p_{1})\leq\varphi(p_{2}). The \Rightarrow direction is immediate from Lemma 2.7. For the other direction, suppose that φ(p1)φ(p2)\varphi(p_{1})\leq\varphi(p_{2}). Then there is a qQq\in Q such that φ(p1)q=φ(p2)\varphi(p_{1})q=\varphi(p_{2}). Since φ\varphi is a surjection from PP to QQ, there exists a pPp\in P such that φ(p)=q\varphi(p)=q, and hence φ(p1p)=φ(p2)\varphi(p_{1}p)=\varphi(p_{2}). Since φ\varphi is injective on PP, then p1p=p2p_{1}p=p_{2}, and thus p1p2p_{1}\leq p_{2}, as desired. ∎

While the theory of reductions of ordered groups can be entirely severed from the amenability of the groups involved, it is perhaps unsurprising that a reduction to an amenable group will allow us to do more interesting analysis of the CC^{*}-algebras of PP-graphs that will appear in later sections.

The following example shows that the existence of a reduction to an amenable group (or even a non-injective reduction!) depends on not just the group GG, but also the positive cone PP.

Example 3.6.

Let G=a,bG=\langle a,b\rangle, the free group on two generators. If P={a,b}P=\{a,b\}^{*}, then we’ve seen that (G,P)(G,P) is an ordered group which reduces to an amenable group via the map φ:(G,P)(,)\varphi:(G,P)\rightarrow(\mathbb{Z},\mathbb{N}) given by φ(a)=φ(b)=1\varphi(a)=\varphi(b)=1.

Suppose that (G,R)(G,R) is a total ordering on a group (meaning that RR1=GR\cup R^{-1}=G in addition to RR1={1}R\cap R^{-1}=\{1\}). Then for any reduction φ\varphi, by Proposition 3.4(5), since S=rR[1,r][1,r]1=GS=\bigcup_{r\in R}[1,r][1,r]^{-1}=G, then kerφ=kerφG={1}\ker\varphi=\ker\varphi\cap G=\{1\}. That is, every reduction φ\varphi is injective.

In particular, taking G=F2G=F_{2} and RR a total ordering on GG (such as the ordering arising from the Magnus expansion given in [4, Section 3.2]), (F2,R)(F_{2},R) cannot reduce to an amenable group since the range group of a reduction will always contain a copy of F2F_{2}.

Example 3.7.

For any group (G,P)(G,P), idG\operatorname{id}_{G} is a reduction, so if GG is amenable, idG\operatorname{id}_{G} is a reduction to an amenable group.

Example 3.8.

Let G=a,bG=\langle a,b\rangle, the free group on two generators and P={a,b}P=\{a,b\}^{*}. One can give a strong reduction of (G,P)(G,P) onto the amenable group H=BS(1,2)=c,t|tc=c2tH=BS(1,2)=\langle c,t|tc=c^{2}t\rangle. Let Q={t,ct}Q=\{t,ct\}^{*}, which is a positive cone in BS(1,2)BS(1,2), and let φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) by φ(a)=ct,φ(b)=t\varphi(a)=ct,\varphi(b)=t. Certainly φ\varphi is an order homomorphism. Then one may verify that φ\varphi is bijective from PP to QQ in this way: if φ(p1)=φ(p2)\varphi(p_{1})=\varphi(p_{2}), then write this word in HH as citjc^{i}t^{j}. It must be that jj is the length of p1p_{1} and the length of p2p_{2}, so p1p_{1} and p2p_{2} have equal length. Then, ii will be the number that comes from substituting 11 for aa and 0 for bb in the original word and interpreting the result as a binary number. Two binary numbers with the same number of digits are equal if and only if their digits are equal, so it must be that p1=p2p_{1}=p_{2}. Thus φ\varphi is injective on PP, and hence bijective on PP, meaning φ\varphi is a strong reduction.

The following result shows that reductions preserve weakly quasi-lattice order.

Lemma 3.9.

Let φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) be a reduction. Then:

  1. 1.

    If (H,Q)(H,Q) is WQLO, then so is (G,P)(G,P).

  2. 2.

    If φ(P)=Q\varphi(P)=Q and (G,P)(G,P) is WQLO, then so is (H,Q)(H,Q).

Proof.

For (1), assume for the sake of contradiction that (H,Q)(H,Q) is WQLO but (G,P)(G,P) is not. Then there are x,yPx,y\in P with a common upper bound but no least common upper bound. That is, there are upper bounds b1>b2>b3>b_{1}>b_{2}>b_{3}>... such that x,ybix,y\leq b_{i} for all ii\in\mathbb{N}.

For all ii\in\mathbb{N}, since 1xbi1\leq x\leq b_{i}, then 1bi1\leq b_{i}. In particular, bi[1,b1]b_{i}\in[1,b_{1}] for all ii\in\mathbb{N}. Also note that x,y[1,b1]x,y\in[1,b_{1}].

Now, since φ\varphi is a reduction, φ\varphi is an order isomorphism from [1,b1][1,b_{1}] to [1,φ(b1)][1,\varphi(b_{1})], so in QQ, φ(b1)>φ(b2)>\varphi(b_{1})>\varphi(b_{2})>... is a strictly decreasing sequence of common upper bounds of φ(x)\varphi(x) and φ(y)\varphi(y). This contradicts the fact that (H,Q)(H,Q) is WQLO.

For (2), assume for the sake of contradiction that φ(P)=Q\varphi(P)=Q and (G,P)(G,P) is WQLO but that (H,Q)(H,Q) is not WQLO. Then there are z,wQz,w\in Q with a common upper bound but no least common upper bound. That is, there are upper bounds c1>c2>c3>c_{1}>c_{2}>c_{3}>... such that z,wciz,w\leq c_{i} for all ii\in\mathbb{N}.

For all ii\in\mathbb{N}, since 1zci1\leq z\leq c_{i}, then 1ci1\leq c_{i}. In particular, ci[1,c1]c_{i}\in[1,c_{1}] for all ii\in\mathbb{N}. Also note that z,w[1,c1]z,w\in[1,c_{1}].

Now, since φ(P)=Q\varphi(P)=Q, there is some pPp\in P such that φ(p)=c1\varphi(p)=c_{1}. Since φ\varphi is a reduction, φ\varphi is an order isomorphism from [1,p][1,p] to [1,φ(p)]=[1,c1][1,\varphi(p)]=[1,c_{1}]. Let xx, yy, and bib_{i} denote the unique elements in [1,p][1,p] such that φ(x)=z,φ(y)=w\varphi(x)=z,\varphi(y)=w, and φ(bi)=ci\varphi(b_{i})=c_{i} for all ii\in\mathbb{N}. Then in PP, b1>b2>b3b_{1}>b_{2}>b_{3}... is a strictly decreasing sequence of common upper bounds of xx and yy. This contradicts the fact that (G,P)(G,P) is WQLO.

Remark 3.10.

There is a notion of amenability for semigroups, and (G,P)(G,P) may strongly reduce to an amenable group even if neither GG nor PP is amenable as a group or semigroup (respectively). For example, in (F2,P2)(F_{2},P_{2}), the free group on 2 generators with positive cone P2P_{2} which is the free monoid on 2 generators, we will show that (F2,P2)(F_{2},P_{2}) strongly reduces to (,)(\mathbb{Z}\wr\mathbb{Z},\mathbb{N}\wr\mathbb{N}), which is amenable, but F2F_{2} is famously not amenable, and P2P_{2} is also not amenable as a semigroup.

The fact that P2P_{2} is not amenable but can be embedded in an amenable group has been known for more than half a century. In [11], the author shows an embedding of P2P_{2} into 2\mathbb{Z}\wr\mathbb{Z}^{2}.

3.2 Constructions

In this section, we will show that the notion of “(strongly) reduces to an (amenable) group” behaves well with the usual group theory constructions such as composition, hereditary subgroups, direct products, and free products.

3.2.1 Composition

Lemma 3.11.

A composition of (strong) reductions is a (strong) reduction. That is, if φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) and ψ:(H,Q)(F,R)\psi:(H,Q)\rightarrow(F,R) are two (strong) reductions, then ψφ\psi\circ\varphi is a (strong) reduction.

Proof.

Suppose φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) and ψ:(H,Q)(F,R)\psi:(H,Q)\rightarrow(F,R) are two reductions. Then we will show that ψφ\psi\circ\varphi is a reduction.

Certainly, ψφ\psi\circ\varphi is an order homomorphism, so by Lemma 3.4 (3), we must show that for all pPp\in P, ψφ\psi\circ\varphi bijectively maps [1,p][1,p] onto [1,ψ(φ(p))][1,\psi(\varphi(p))].

To this end, we know that φ:[1,p][1,φ(p)]\varphi:[1,p]\rightarrow[1,\varphi(p)] is bijective, and ψ:[1,φ(p)][1,ψ(φ(p))]\psi:[1,\varphi(p)]\rightarrow[1,\psi(\varphi(p))] are bijections since φ\varphi and ψ\psi are reductions, so therefore their composition is a bijection, as desired.

For the composition of strong reductions, if φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) and ψ:(H,Q)(F,R)\psi:(H,Q)\rightarrow(F,R) are two strong reductions, then φ\varphi maps PP bijectively onto QQ and ψ\psi maps QQ bijectively onto RR, so ψφ\psi\circ\varphi maps PP bijectively onto RR, as desired.

3.2.2 Hereditary Subgroups

In the context of ordered groups, the “correct” notion of a subgroup is often a hereditary subgroups in the sense of [3, Corollary 5.6]. In this section we will remind the reader of the definition and some basic results about the concept.

Definition 3.12.

Given an ordered group (G,P)(G,P), if QPQ\subseteq P is a subsemigroup, we say that QQ is hereditary in PP if p1,p2Pp_{1},p_{2}\in P, p1p2Qp_{1}p_{2}\in Q implies that p1,p2Qp_{1},p_{2}\in Q.

If (G,P)(G,P) is an ordered group, we will say that (H,Q)(H,Q) is a hereditary subgroup if HH is a subgroup of GG, QQ is a subsemigroup of PHP\cap H, and QQ is hereditary in PP.

Lemma 3.13.

Let (H,P)(H,P) be an ordered group, and QPQ\subseteq P a hereditary subsemigroup. For the sake of clarity, let P\leq_{P} and Q\leq_{Q} denote the orderings on GG arising from PP and QQ, respectively. Then for all qQq\in Q and pPp\in P, 1PpPq1\leq_{P}p\leq_{P}q if and only if 1QpQq1\leq_{Q}p\leq_{Q}q.

In particular, for qQq\in Q, the intervals [1,q]P[1,q]_{\leq_{P}} and [1,q]Q[1,q]_{\leq_{Q}} are equal as sets.

Proof.

Fix some qQ,pPq\in Q,p\in P.

Suppose 1PpPq1\leq_{P}p\leq_{P}q. Then pPp\in P and there exists an sPs\in P such that ps=qps=q. Since qQq\in Q and QQ is hereditary, then this implies that p,sQp,s\in Q, so 1QpQq1\leq_{Q}p\leq_{Q}q, as desired.

Suppose 1QpQq1\leq_{Q}p\leq_{Q}q. Then pQp\in Q and there exists an sQs\in Q such that ps=qps=q. Since QPQ\subseteq P, then 1PpPq1\leq_{P}p\leq_{P}q, as desired.

The “in particular” part is immediate.

Lemma 3.14.

Let (G,P),(H,Q)(G,P),(H,Q) be ordered groups, and φ:GH\varphi:G\rightarrow H a homomorphism. Let H=φ(G)H^{\prime}=\varphi(G), Q=φ(P)Q^{\prime}=\varphi(P), and let φ\varphi^{\prime} denote the codomain restriction of φ\varphi to the codomain HH^{\prime}. Then the following are equivalent:

  1. 1.

    φ\varphi is a reduction.

  2. 2.

    (H,Q)(H^{\prime},Q^{\prime}) is a hereditary subgroup of (H,Q)(H,Q) and φ\varphi^{\prime} is a surjective reduction of (G,P)(G,P) onto (H,Q)(H^{\prime},Q^{\prime}).

Proof.

(12)(1\Rightarrow 2) If φ\varphi is a reduction, we will first show that Q=φ(P)Q^{\prime}=\varphi(P) is a hereditary subsemigroup of QQ. If q=φ(p)φ(P)q=\varphi(p)\in\varphi(P), and rr=qrr^{\prime}=q for some r,rQr,r^{\prime}\in Q, then r[1,φ(p)]r\in[1,\varphi(p)], so by Proposition 3.4(3), there exists a unique s[1,p]s\in[1,p] such that φ(s)=r\varphi(s)=r. Then φ(s)r=φ(p)\varphi(s)r^{\prime}=\varphi(p), so r=φ(s1p)r^{\prime}=\varphi(s^{-1}p). Since sps\leq p, then there exists sPs^{\prime}\in P such that ss=pss^{\prime}=p, so r=φ(s)r^{\prime}=\varphi(s^{\prime}). Thus r,rφ(P)r,r^{\prime}\in\varphi(P), so φ(P)\varphi(P) is a hereditary subsemigroup of QQ.

Since additionally H<HH^{\prime}<H and Q<Qφ(G)Q^{\prime}<Q\cap\varphi(G), then (H,Q)(H^{\prime},Q^{\prime}) is a hereditary subgroup of (H,Q)(H,Q).

To show that φ\varphi^{\prime} is a surjective reduction, it is immediately surjective since its codomain was restricted to the image. Now fix a pPp\in P. Since φ\varphi is a reduction, then by Proposition 3.4 (3), φ\varphi maps [1,p][1,p] bijectively onto [1,φ(p)][1,\varphi(p)]. By Lemma 3.13, we have [1,φ(p)]=[1,φ(p)][1,\varphi(p)]=[1,\varphi^{\prime}(p)] as sets, so φ\varphi^{\prime} maps [1,p][1,p] bijectively onto [1,φ(p)][1,\varphi^{\prime}(p)]. Then by Proposition 3.4 (3), φ\varphi^{\prime} is a reduction.

(21)(2\Rightarrow 1) Fix a pPp\in P. Since φ\varphi^{\prime} is a reduction, then by Proposition 3.4 (3), φ\varphi^{\prime} maps [1,p][1,p] bijectively onto [1,φ(p)][1,\varphi^{\prime}(p)]. By Lemma 3.13, we have [1,φ(p)]Q=[1,φ(p)]Q[1,\varphi^{\prime}(p)]_{\leq Q^{\prime}}=[1,\varphi(p)]_{\leq Q} as sets, so φ\varphi maps [1,p][1,p] bijectively onto [1,φ(p)][1,\varphi(p)]. Then by Proposition 3.4 (3), φ\varphi is a reduction.

Lemma 3.15.

Let (G,P)(G,P) be an ordered group and (H,Q)(H,Q) a hereditary subgroup of (G,P)(G,P). Then the inclusion map i:(H,Q)(G,P)i:(H,Q)\rightarrow(G,P) is a reduction.

Proof.

Observe that the inclusion map, restricted to its image, is the identity map, and thus a surjective reduction. Then the inclusion map is a reduction by Lemma 3.14.

We can now conclude that “reduces to an amenable group” is closed under taking hereditary subgroups.

Corollary 3.16.

Let (G,P)(G,P) be an ordered group and (H,Q)(H,Q) a hereditary subgroup of (G,P)(G,P). If (G,P)(G,P) reduces to an amenable group, then (H,Q)(H,Q) reduces to an amenable group.

Proof.

Let i:HGi:H\rightarrow G denote the inclusion map, which by the previous lemma is a reduction.

If φ:(G,P)(F,R)\varphi:(G,P)\rightarrow(F,R) is a reduction to an amenable group, then φi:(H,Q)(F,R)\varphi\circ i:(H,Q)\rightarrow(F,R) is a composition of reductions, and hence a reduction by Lemma 3.11. It has amenable range, so it is a reduction to an amenable group. ∎

3.2.3 Direct Products

Now, we’ll show that reductions respect direct products, which implies that “reduces to an amenable group” is preserved under direct products.

First, let us check that direct products preserve WQLO:

Lemma 3.17.

Let (G,P)(G,P) and (H,Q)(H,Q) be ordered groups. Let P×QP\times Q denote the submonoid of G×HG\times H generated by (the canonical copies of) PP and QQ. Then (G×H,P×Q)(G\times H,P\times Q) is an ordered group, and if (G,P)(G,P) and (H,Q)(H,Q) are WQLO, then so is (G×H,P×Q)(G\times H,P\times Q).

Proof.

By definition, P×QP\times Q is a submonoid of G×HG\times H, and it is immediate that P×Q(P×Q)1=(PP1)×(QQ1)={1G}×{1H}P\times Q\cap(P\times Q)^{-1}=(P\cap P^{-1})\times(Q\cap Q^{-1})=\{1_{G}\}\times\{1_{H}\}, so (G×H,P×Q)(G\times H,P\times Q) is an ordered group.

If (G,P)(G,P) and (H,Q)(H,Q) are WQLO, then one may quickly confirm that given (p1,q1),(p2,q2)P×Q(p_{1},q_{1}),(p_{2},q_{2})\in P\times Q, their least upper bound is (p1p2,q1q2)(p_{1}\vee p_{2},q_{1}\vee q_{2}).

Lemma 3.18.

For i=1,2i=1,2, let (Gi,Pi)(G_{i},P_{i}) and (Hi,Qi)(H_{i},Q_{i}) be ordered groups, φi:GiHi\varphi_{i}:G_{i}\rightarrow H_{i} a homomorphism. Let G=G1×G2G=G_{1}\times G_{2}, and similarly define P,H,QP,H,Q. Then there is a homomorphism φ:GH\varphi:G\rightarrow H given by φ(g1,g2)=(φ1(g1),φ2(g2))\varphi(g_{1},g_{2})=(\varphi_{1}(g_{1}),\varphi_{2}(g_{2})), and φ\varphi is an order homomorphism (respectively, reduction or strong reduction) if φ1\varphi_{1} and φ2\varphi_{2} both are.

Proof.

The existence of such a homomorphism is an elementary fact of group theory.

If φ1,φ2\varphi_{1},\varphi_{2} are both order homomorphisms, then φ1(P1)Q1\varphi_{1}(P_{1})\subseteq Q_{1} and φ2(P2)Q2\varphi_{2}(P_{2})\subseteq Q_{2}, so φ(P1×P2)φ1(P1)×φ2(P2)Q1×Q2\varphi(P_{1}\times P_{2})\subseteq\varphi_{1}(P_{1})\times\varphi_{2}(P_{2})\subseteq Q_{1}\times Q_{2}, as desired.

If φ1,φ2\varphi_{1},\varphi_{2} are both reductions, then we will use Proposition 3.4 (3) as our notion of a reduction, so it suffices to show that for all p1P1,p2P2p_{1}\in P_{1},p_{2}\in P_{2}, [(1,1),(p1,p2)][(1,1),(p_{1},p_{2})] is mapped bijectively onto [(1,1),(φ1(p1),φ2(p2)][(1,1),(\varphi_{1}(p_{1}),\varphi_{2}(p_{2})]. But in (G1×G2,P1×P2)(G_{1}\times G_{2},P_{1}\times P_{2}), we have that [(1,1),(x1,x2)]=[1,x1]×[1,x2][(1,1),(x_{1},x_{2})]=[1,x_{1}]\times[1,x_{2}], so we know that φ=φ1×φ2\varphi=\varphi_{1}\times\varphi_{2} bijectively carries [(1,1),(p1,p2)]=[1,p1]×[1,p2][(1,1),(p_{1},p_{2})]=[1,p_{1}]\times[1,p_{2}] onto [(1,1),(φ1(p1),φ2(p2))]=[1,φ1(p1)]×[1,φ2(p2)][(1,1),(\varphi_{1}(p_{1}),\varphi_{2}(p_{2}))]=[1,\varphi_{1}(p_{1})]\times[1,\varphi_{2}(p_{2})], as desired.

If φ1,φ2\varphi_{1},\varphi_{2} are both strong reductions, then φ1\varphi_{1} and φ2\varphi_{2} map P1P_{1} and P2P_{2} bijectively onto Q1Q_{1} and Q2Q_{2} respectively, so for all q=(q1,q2)Qq=(q_{1},q_{2})\in Q there is a unique p1P1,p2P2p_{1}\in P_{1},p_{2}\in P_{2} such that φ1(p1)=q1,φ2(p2)=q2\varphi_{1}(p_{1})=q_{1},\varphi_{2}(p_{2})=q_{2}, and hence p=(p1,p2)p=(p_{1},p_{2}) is the unique element of PP with φ(p)=q\varphi(p)=q, so φ\varphi is a bijection from PP to QQ, as desired.

Corollary 3.19.

If (G1,P1)(G_{1},P_{1}), (G2,P2)(G_{2},P_{2}) are ordered groups which reduce to amenable ordered groups, then (G1×G2,P1×P2)(G_{1}\times G_{2},P_{1}\times P_{2}) reduces to an amenable group.

Proof.

For i=1,2i=1,2, let φi:(Gi,Pi)(Hi,Qi)\varphi_{i}:(G_{i},P_{i})\rightarrow(H_{i},Q_{i}) denote the reduction of (Gi,Pi)(G_{i},P_{i}) to an amenable ordered group.

By the previous Lemma, we know that (G1×G2,P1×P2)(G_{1}\times G_{2},P_{1}\times P_{2}) reduces to (H1×H2,Q1×Q2)(H_{1}\times H_{2},Q_{1}\times Q_{2}). Since H1H_{1} and H2H_{2} are amenable, then H1×H2H_{1}\times H_{2} is amenable, so this is a reduction to an amenable group. ∎

3.2.4 Free Products

In this section we will show that the class of groups which have reductions to amenable groups is closed under (finite) free products. That is, we will show that if (G,P)(G,P) and (H,Q)(H,Q) reduce to amenable groups, then (GH,PQ)(G*H,P*Q) also reduces to an amenable group.

This will consist of two steps, analogous to Lemma 3.18 and Corollary 3.19 from the direct product case. A free product analogue of Lemma 3.18 is straightforward. However, in the free product case, an analogue of Corollary 3.19 is more difficult, since the free product of two amenable groups is almost never amenable. Therefore, our second step is longer.

Given ordered groups (G1,P1)(G_{1},P_{1}) and (G2,P2)(G_{2},P_{2}), we denote by P1P2P_{1}*P_{2} the submonoid of G1G2G_{1}*G_{2} generated by the naturally embedded copies of P1P_{1} and P2P_{2}. First, we will prove a lemma characterizing the order on (G1G2,P1P2)(G_{1}*G_{2},P_{1}*P_{2}):

Lemma 3.20.

Let (G1,P1),(G2,P2)(G_{1},P_{1}),(G_{2},P_{2}) be ordered groups, and let (G,P)=(G1G2,P1P2)(G,P)=(G_{1}*G_{2},P_{1}*P_{2}).

  1. 1.

    For pPp\in P, we may write p=p1p2p2np=p_{1}p_{2}...p_{2n} where piP1p_{i}\in P_{1} if ii is odd, and piP2p_{i}\in P_{2} if ii even, and pi1p_{i}\neq 1 for 1<i<2n1<i<2n.

  2. 2.

    This representation of pp is unique. That is, there is exactly one such choice of nn and elements pip_{i} such that piP1p_{i}\in P_{1} if ii is odd, and piP2p_{i}\in P_{2} if ii even, and pi1p_{i}\neq 1 for 1<i<2n1<i<2n.

  3. 3.

    Given such a representation p=p1p2p2np=p_{1}p_{2}...p_{2n}, we have [1,p]=i=12nXi[1,p]=\bigcup_{i=1}^{2n}X_{i}, where Xi=p1pi1[1,pi]X_{i}=p_{1}...p_{i-1}[1,p_{i}].

  4. 4.

    Given such a decomposition [1,p]=i=12nXi[1,p]=\bigcup_{i=1}^{2n}X_{i}, if aXia\in X_{i} and bXjb\in X_{j} for i<ji<j, then aba\leq b.

Proof.

For (1), by definition every element of PP can be written as a product of elements of P1P_{1} and P2P_{2}. If two consecutive elements are from the same PiP_{i}, they can be combined, and if any intermediate term is a 1, it can be removed and the now-consecutive terms combined. Finally, if necessary a term of 1 can be put at the beginning or end of the expression to make the alternating product begin with a term from P1P_{1} and end with a term from P2P_{2}.

For (2), this representation is unique because of the Normal Form Theorem for Free Products ([16, Chapter IV, Theorem 1.2]).

For (3), fix a p=p1p2p2np=p_{1}p_{2}...p_{2n}. It is immediate that each XiX_{i} is a subset of [1,p][1,p], so it suffices to show containment in the other direction. To this end, suppose q[1,p]q\in[1,p], so qPq\in P and p=qrp=qr for some rPr\in P. Then writing q=q1q2mq=q_{1}...q_{2m} and r=r1r2kr=r_{1}...r_{2k}, we have that p=p1p2p2n=q1q2mr1r2kp=p_{1}p_{2}...p_{2n}=q_{1}...q_{2m}r_{1}...r_{2k}. We now have a few cases depending on whether or not q2mq_{2m} and r1r_{1} are the identity.

  • q2m=1,r1=1q_{2m}=1,r_{1}=1: Then p=p1p2p2n=q1q2m1r2r2kp=p_{1}p_{2}...p_{2n}=q_{1}...q_{2m-1}r_{2}...r_{2k} is the unique alternating presentation with nonidentity terms, so p1=q1,p2=q2p_{1}=q_{1},p_{2}=q_{2}, etc. In particular, q=q1q2m1=p1p2m1X2m=p1p2m1[1,p2m]q=q_{1}...q_{2m-1}=p_{1}...p_{2m-1}\in X_{2m}=p_{1}...p_{2m-1}[1,p_{2m}].

  • q2m1,r11q_{2m}\neq 1,r_{1}\neq 1: Then p=p1p2p2n=q1q2mr1r2kp=p_{1}p_{2}...p_{2n}=q_{1}...q_{2m}r_{1}...r_{2k} is the unique alternating presentation with nonidentity terms, so p1=q1,p2=q2p_{1}=q_{1},p_{2}=q_{2}, etc. In particular, q=q1q2m=p1p2mX2m=p1p2m1[1,p2m]q=q_{1}...q_{2m}=p_{1}...p_{2m}\in X_{2m}=p_{1}...p_{2m-1}[1,p_{2m}].

  • q2m1,r1=1q_{2m}\neq 1,r_{1}=1: Then p=p1p2p2n=q1(q2mr2)r3r2kp=p_{1}p_{2}...p_{2n}=q_{1}...(q_{2m}r_{2})r_{3}...r_{2k} is the unique alternating presentation with nonidentity terms, so p1=q1,p2=q2p_{1}=q_{1},p_{2}=q_{2}, etc, ending with p2m=q2mr2p_{2m}=q_{2m}r_{2}. Thus q2m[1,p2m]q_{2m}\in[1,p_{2m}], so q=q1q2m=p1p2p2m1q2mX2m=p1p2m1[1,p2m]q=q_{1}...q_{2m}=p_{1}p_{2}...p_{2m-1}q_{2m}\in X_{2m}=p_{1}...p_{2m-1}[1,p_{2m}].

  • q2m=1,r11q_{2m}=1,r_{1}\neq 1: Then p=p1p2p2n=q1q2m2(q2m1r1)r2r2kp=p_{1}p_{2}...p_{2n}=q_{1}...q_{2m-2}(q_{2m-1}r_{1})r_{2}...r_{2k} is the unique alternating presentation with nonidentity terms, so p1=q1,p2=q2p_{1}=q_{1},p_{2}=q_{2}, etc, ending with p2m1=q2m1r1p_{2m-1}=q_{2m-1}r_{1}. Thus q2m1[1,p2m1]q_{2m-1}\in[1,p_{2m-1}], so q=q1q2m1=p1p2p2m2q2m1X2m1=p1p2m2[1,p2m1]q=q_{1}...q_{2m-1}=p_{1}p_{2}...p_{2m-2}q_{2m-1}\in X_{2m-1}=p_{1}...p_{2m-2}[1,p_{2m-1}].

In all four cases, qq is in some XiX_{i}, so [1,p]=i=12nXi[1,p]=\bigcup_{i=1}^{2n}X_{i}, as desired.

For (4), note that for each yXiy\in X_{i}, we have p1p2pi1yp1p2pip_{1}p_{2}...p_{i-1}\leq y\leq p_{1}p_{2}...p_{i}, so ap1p2pip1p2pj1ba\leq p_{1}p_{2}...p_{i}\leq p_{1}p_{2}...p_{j-1}\leq b, as desired. ∎

We can now confirm that weak quasi-lattice order is preserved under free products.

Lemma 3.21.

If (G1,P1)(G_{1},P_{1}) and (G2,P2)(G_{2},P_{2}) are ordered groups (respectively, WQLO groups) then so is (G1G2,P1P2)(G_{1}*G_{2},P_{1}*P_{2}).

Proof.

If (G1,P1)(G_{1},P_{1}) and (G2,P2)(G_{2},P_{2}) are ordered groups, then by definition P1P2P_{1}*P_{2} is a submonoid of GHG*H. We must now check that (P1P2)(P1P2)1={1}(P_{1}*P_{2})\cap(P_{1}*P_{2})^{-1}=\{1\}, or equivalently that if a,bP1P2a,b\in P_{1}*P_{2} with ab=1ab=1, then a=b=1a=b=1. But if ab=1ab=1, then by writing a=p1p2p2na=p_{1}p_{2}...p_{2n} and b=q1q2q2mb=q_{1}q_{2}...q_{2m} as in the previous theorem, then 1=ab=p1p2p2nq1q2q2m1=ab=p_{1}p_{2}...p_{2n}q_{1}q_{2}...q_{2m}. By the Normal Form Theorem for Free Products ([16, Chapter IV, Theorem 1.2], a reduced sequence in a free product can equal 1 if and only if its length is 1, so all the pip_{i}s and qiq_{i}s in the products must cancel with each other. But since each term of the product is positive, they can only cancel if each term is equal to 1. Thus a=b=1a=b=1, as desired, so (G1G2,P1P2)(G_{1}*G_{2},P_{1}*P_{2}) is an ordered group.

If (G1,P1)(G_{1},P_{1}) and (G2,P2)(G_{2},P_{2}) are WQLO, we will check that (G1G2,P1P2)(G_{1}*G_{2},P_{1}*P_{2}) is WQLO. To this end, fix some y1,y2P1P2y_{1},y_{2}\in P_{1}*P_{2} and suppose y1,y2y_{1},y_{2} have some common upper bound y3P1P2y_{3}\in P_{1}*P_{2}. By part (1) of the previous lemma, we can write y3=p1p2p3p2n1p2ny_{3}=p_{1}p_{2}p_{3}...p_{2n-1}p_{2n}, where each piP1p_{i}\in P_{1} for ii odd and piP2p_{i}\in P_{2} for ii even. By part (3) of that lemma, [1,y3]=i=12nXi[1,y_{3}]=\bigcup_{i=1}^{2n}X_{i} where Xi=p1pi1[1,pi]X_{i}=p_{1}...p_{i-1}[1,p_{i}]. Since y1,y2[1,y3]y_{1},y_{2}\in[1,y_{3}], then y1Xi,y2Xjy_{1}\in X_{i},y_{2}\in X_{j} for some i,j2ni,j\leq 2n. Without loss of generality, suppose iji\leq j.

If i<ji<j, then by part (4) of the previous Lemma, we have y1y2y_{1}\leq y_{2}, so y1y2=y2y_{1}\vee y_{2}=y_{2}.

If i=ji=j, then y1,y2Xi=p1pi1[1,pi]y_{1},y_{2}\in X_{i}=p_{1}...p_{i-1}[1,p_{i}], so there are r1,r2r_{1},r_{2} such that y1=p1pi1r1y_{1}=p_{1}...p_{i-1}r_{1} and y2=p1pi1r2y_{2}=p_{1}...p_{i-1}r_{2} where r1,r2P1r_{1},r_{2}\in P_{1} if ii is odd, and r1,r2P2r_{1},r_{2}\in P_{2} if ii is even. In either case, p1pi1(r1r2)p_{1}...p_{i-1}(r_{1}\vee r_{2}) will be the supremum of y1y_{1} and y2y_{2}.

We will now show that if two groups have a reduction, their free product reduces to the free product of the reductions.

Lemma 3.22.

For i=1,2i=1,2, let φi:(Gi,Pi)(Hi,Qi)\varphi_{i}:(G_{i},P_{i})\rightarrow(H_{i},Q_{i}) be a reduction. Then let

G=G1G2,P=P1P2,H=H1H2, and Q=Q1Q2,G=G_{1}*G_{2},P=P_{1}*P_{2},H=H_{1}*H_{2},\text{ and }Q=Q_{1}*Q_{2},

and let φ:GH\varphi:G\rightarrow H be the homomorphism satisfying φGi=φi\varphi\mid_{G_{i}}=\varphi_{i} for i=1,2i=1,2. Then φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) is a reduction.

Proof.

Certainly, φ\varphi is a homomorphism. By Proposition 3.4 (3), it suffices to show that for all pPp\in P, the interval [1,p][1,p] is mapped bijectively onto [1,φ(p)][1,\varphi(p)].

Fix some pPp\in P, and then by Lemma 3.20, we may write p=p1p2p2np=p_{1}p_{2}...p_{2n} where pip_{i} is in P1P_{1} if ii is odd, and in P2P_{2} if ii is even. Then [1,p]=i=12nXi[1,p]=\bigcup_{i=1}^{2n}X_{i}, where Xi=p1pi1[1,pi]X_{i}=p_{1}...p_{i-1}[1,p_{i}].

Now, for 1i2n1\leq i\leq 2n, let qi=φ1(pi)q_{i}=\varphi_{1}(p_{i}) if ii is odd, and qi=φ2(pi)q_{i}=\varphi_{2}(p_{i}) if ii is even. Thus φ(p)=q1q2n\varphi(p)=q_{1}...q_{2n}. Note that for 1<i<2n1<i<2n, qi1q_{i}\neq 1 since pi1p_{i}\neq 1 and by Proposition 3.4 (4), both φ1\varphi_{1} and φ2\varphi_{2} send strictly positive elements to strictly positive elements.

Thus φ(p)=q1q2n\varphi(p)=q_{1}...q_{2n} is the unique way to write φ(p)Q\varphi(p)\in Q as an alternating product of non-unit elements, so by Lemma 3.20, [1,φ(p)]=i=12nYi[1,\varphi(p)]=\bigcup_{i=1}^{2n}Y_{i} where Yi=q1qi1[1,qi]Y_{i}=q_{1}...q_{i-1}[1,q_{i}]. But since φ1\varphi_{1} and φ2\varphi_{2} are reductions, we know that each XiX_{i} is carried bijectively to its YiY_{i}, so it suffices to explain why bijectivity is preserved when taking the union of the XiX_{i} and YiY_{i}s. It is immediate that the map from Xi\bigcup X_{i} to Yi\bigcup Y_{i} is surjective.

For injectivity, first note that for 1ijn1\leq i\leq j\leq n,

YiYj={j>i+1{q1qi}j=i+1Yii=jY_{i}\cap Y_{j}=\begin{cases}\emptyset&j>i+1\\ \{q_{1}...q_{i}\}&j=i+1\\ Y_{i}&i=j\end{cases}

and similarly,

XiXj={j>i+1{p1pi}j=i+1Xii=jX_{i}\cap X_{j}=\begin{cases}\emptyset&j>i+1\\ \{p_{1}...p_{i}\}&j=i+1\\ X_{i}&i=j\end{cases}

Then, if φ(s1)Yi\varphi(s_{1})\in Y_{i} and φ(s2)Yj\varphi(s_{2})\in Y_{j} for s1,s2[1,p]s_{1},s_{2}\in[1,p] such that φ(s1)=φ(s2)\varphi(s_{1})=\varphi(s_{2}), either i=ji=j so s1,s2Xis_{1},s_{2}\in X_{i} and thus by injectivity on XiX_{i} we have s1=s2s_{1}=s_{2}, or j=i+1j=i+1, so φ(s1)=φ(s2)=q1qi\varphi(s_{1})=\varphi(s_{2})=q_{1}...q_{i}, so s1=s2=p1pis_{1}=s_{2}=p_{1}...p_{i}. In either case, s1=s2s_{1}=s_{2}, so we have injectivity of φ\varphi. Thus φ\varphi is bijective on [1,p][1,p] as desired, so it is a reduction.

This completes the first step, showing that if two groups have reductions to amenable groups, their free product reduces to a free product of amenable groups. Now comes the harder step: showing that a free product of amenable groups reduces to an amenable group!

The key is a construction from group theory called a wreath product. Recall the following definition of a wreath product:

Definition 3.23.

Let GG and HH be groups. Let

GH={f:HG a function|supp(f) is finite}G^{H}=\{f:H\rightarrow G\text{ a function}|\operatorname{supp}(f)\text{ is finite}\}

where supp(f)={hH:f(h)1}\operatorname{supp}(f)=\{h\in H:f(h)\neq 1\}, and give GHG^{H} a group structure by pointwise multiplication. Note that GHG^{H} is isomorphic to a direct sum of |H||H| copies of GG, hence the notation.

Give GHG^{H} an action α:HAut(GH)\alpha:H\rightarrow Aut(G^{H}) by translation: [αh(f)](h)=f(h1h)\left[\alpha_{h}(f)\right](h^{\prime})=f(h^{-1}h^{\prime}).

Then we define the (restricted) wreath product to be the semidirect product GH:=GHαHG\wr H:=G^{H}\rtimes_{\alpha}H, so there is a (split) short exact sequence:

1GHGHH1.1\rightarrow G^{H}\rightarrow G\wr H\rightarrow H\rightarrow 1.

The following remark establishes the basic properties of wreath products, and are all routine to verify.

Remark 3.24.

For each gGg\in G and hHh\in H, let gδhg\delta_{h} denote the function in GHG^{H} given by gδh(h)={gh=h1hhg\delta_{h}(h^{\prime})=\begin{cases}g&h=h^{\prime}\\ 1&h\neq h^{\prime}\end{cases}. Then there is a natural embedding of GG into GHG\wr H by g(gδ1H,1H)g\mapsto(g\delta_{1_{H}},1_{H}) and there is a natural embedding of HH into GHG\wr H by h(1GH,h)h\mapsto(1_{G^{H}},h). In a slight abuse of notation, we will write gg or gδ1g\delta_{1} for the embedded element of GG in GHG\wr H, and write hh for the embedded element of HH in GHG\wr H.

Now note that

[αh2(gδh1)](h3)\displaystyle\left[\alpha_{h_{2}}(g\delta_{h_{1}})\right](h_{3}) =\displaystyle= gδh1(h21h3)\displaystyle g\delta_{h_{1}}(h_{2}^{-1}h_{3})
=\displaystyle= {gh1=h21h31h1h21h3\displaystyle\begin{cases}g&h_{1}=h_{2}^{-1}h_{3}\\ 1&h_{1}\neq h_{2}^{-1}h_{3}\end{cases}
=\displaystyle= {gh2h1=h31h2h1h3\displaystyle\begin{cases}g&h_{2}h_{1}=h_{3}\\ 1&h_{2}h_{1}\neq h_{3}\end{cases}
=\displaystyle= gδh2h1(h3)\displaystyle g\delta_{h_{2}h_{1}}(h_{3})

so unbinding the h3h_{3}, we have that αh2(gδh1)=gδh2h1\alpha_{h_{2}}(g\delta_{h_{1}})=g\delta_{h_{2}h_{1}}. Now,

h2(gδh1)h21\displaystyle h_{2}(g\delta_{h_{1}})h_{2}^{-1} =\displaystyle= (1G,h2)(gδh1,1H)(1G,h21)\displaystyle(1_{G},h_{2})(g\delta_{h_{1}},1_{H})(1_{G},h_{2}^{-1})
=\displaystyle= (1G,h2)(gδh1,h21)\displaystyle(1_{G},h_{2})(g\delta_{h_{1}},h_{2}^{-1})
=\displaystyle= (1G,h2)(gδh1,h21)\displaystyle(1_{G},h_{2})(g\delta_{h_{1}},h_{2}^{-1})
=\displaystyle= (αh2(gδh1),1H)\displaystyle(\alpha_{h_{2}}(g\delta_{h_{1}}),1_{H})
=\displaystyle= gδh2h1\displaystyle g\delta_{h_{2}h_{1}}

Since GHG^{H} is generated by the gδhg\delta_{h}, then together (the embedded copies of) GG and HH generate GHG\wr H.

Given P,QP,Q positive cones in G,HG,H, we let PQP\wr Q denote the monoid generated by the naturally embedded copies of PP and QQ within GHG\wr H.

With these natural embeddings of GG and HH into GHG\wr H, by the universal property of free products of groups, we get a homomorphism φ:GHGH\varphi:G*H\rightarrow G\wr H given by sending (the naturally embedded copies of) GG and HH to (the naturally embedded copies of) GG and HH. We will call this the natural homomorphism of GHG*H into GHG\wr H. Since GHG\wr H is generated by GG and HH, it is immediate that the natural homomorphism is surjective. For the same reason, PQP*Q will be mapped surjectively onto PQP\wr Q. We will show later that φ:(GH,PQ)(GH,PQ)\varphi:(G*H,P*Q)\rightarrow(G\wr H,P\wr Q) is a strong reduction.

Finally, note that every element of GHG^{H} can be written as a (possibly empty) product of a finite number of giδhig_{i}\delta_{h_{i}} where each gig_{i} is nonunital and each hih_{i} is distinct, and this presentation is unique up to a permutation of the giδhig_{i}\delta_{h_{i}}. Note that the element giδhig_{i}\delta_{h_{i}} commutes with the element gjδhjg_{j}\delta_{h_{j}} as long as hihjh_{i}\neq h_{j}. When fGHf\in G^{H} is written as f=g1δh1gnδhnf=g_{1}\delta_{h_{1}}...g_{n}\delta_{h_{n}} where the hih_{i} are distinct, then

f(h)={g1 if h=h1g2 if h=h2gn if h=hn1 otherwise.f(h)=\begin{cases}g_{1}&\text{ if }h=h_{1}\\ g_{2}&\text{ if }h=h_{2}\\ \vdots\\ g_{n}&\text{ if }h=h_{n}\\ 1&\text{ otherwise}\end{cases}.

Finally, we will remind the reader of this fact about amenability of wreath products:

Lemma 3.25.

Let GG and HH be amenable groups. Then GHG\wr H is amenable.

Proof.

Recall that the class of amenable groups are closed under finite direct sums, direct limits, and extensions.

From our definition of the wreath product, there is a short exact sequence

1GHGHH1.1\rightarrow G^{H}\rightarrow G\wr H\rightarrow H\rightarrow 1.

Recall also that GHG^{H} is isomorphic to the direct sum of |H||H| copies of GG. Thus GHG^{H} is a direct limit of {Gn}n\{G^{n}\}_{n\in\mathbb{N}}. Since amenable groups are closed under finite direct sum and GG is amenable by hypothesis, then each GnG^{n} is amenable, and since amenable groups are closed under direct limit, then GHG^{H} is amenable.

Also, HH is amenable by hypothesis, so GHG\wr H is an extension of an amenable group by an amenable group. Thus GHG\wr H is amenable.

Now, we will have two results that establish why φ:(G1G2,P1P2)(G1G2,P1P2)\varphi:(G_{1}*G_{2},P_{1}*P_{2})\rightarrow(G_{1}\wr G_{2},P_{1}\wr P_{2}) is a reduction.

Lemma 3.26.

Let (G1,P1)(G_{1},P_{1}) and (G2,P2)(G_{2},P_{2}) be ordered groups, and let p=p1p2p2nP1P2p=p_{1}p_{2}...p_{2n}\in P_{1}\wr P_{2} be an element such that piP1p_{i}\in P_{1} for ii odd and piP2p_{i}\in P_{2} for ii even and pi1p_{i}\neq 1 for 1<i<n1<i<n. Then, as an element of P1P2G1G2G2P_{1}\wr P_{2}\subset G_{1}^{G_{2}}\rtimes G_{2}, p=(f,p2n)p=(f,p_{\leq 2n}) where p2n=p2p4p2np_{\leq 2n}=p_{2}p_{4}...p_{2n} and

f(h)={p1 if h=1G2p3 if h=p2p5 if h=p2p4p2n1 if h=p2p2n21 otherwisef(h)=\begin{cases}p_{1}&\text{ if }h=1_{G_{2}}\\ p_{3}&\text{ if }h=p_{2}\\ p_{5}&\text{ if }h=p_{2}p_{4}\\ \vdots\\ p_{2n-1}&\text{ if }h=p_{2}...p_{2n-2}\\ 1&\text{ otherwise}\end{cases}
Proof.

By elementary algebra, we may rewrite the product p=p1p2p3p2np=p_{1}p_{2}p_{3}...p_{2n} as:

p=p1p2p3p2n=p1(p2p3p21)(p2p4p2n2p2n1p2n21p41p21)(p2p4p2n)p=p_{1}p_{2}p_{3}...p_{2n}=p_{1}(p_{2}p_{3}p_{2}^{-1})...(p_{2}p_{4}...p_{2n-2}p_{2n-1}p_{2n-2}^{-1}...p_{4}^{-1}p_{2}^{-1})(p_{2}p_{4}...p_{2n})

With the shorthand that p0=1Hp_{\leq 0}=1_{H} and p2i=p2p4p2ip_{\leq 2i}=p_{2}p_{4}...p_{2i} for 1in1\leq i\leq n, we then have that

p=[(p0p1p01)(p2p3p21)(p4p5p41)(p2n2p2n1p2n21)]p2n.p=\left[(p_{\leq 0}p_{1}p_{\leq 0}^{-1})(p_{\leq 2}p_{3}p_{\leq 2}^{-1})(p_{\leq 4}p_{5}p_{\leq 4}^{-1})...(p_{\leq 2n-2}p_{2n-1}p_{\leq 2n-2}^{-1})\right]p_{\leq 2n}.

Let f=(p0p1p01)(p2p3p21)(p4p5p41)(p2n2p2n1p2n21)f=(p_{\leq 0}p_{1}p_{\leq 0}^{-1})(p_{\leq 2}p_{3}p_{\leq 2}^{-1})(p_{\leq 4}p_{5}p_{\leq 4}^{-1})...(p_{\leq 2n-2}p_{2n-1}p_{\leq 2n-2}^{-1}) be the bracketed term, so then pp is represented in G1G2G2G_{1}^{G_{2}}\rtimes G_{2} by (f,p2n)(f,p_{\leq 2n}). It then suffices to show that ff assumes the values as claimed.

Recall that for ii odd, pip_{i} is embedded as piδ1Hp_{i}\delta_{1_{H}}, and for ii even, pip_{i} is embedded as an element of G2G_{2}. By the calculation in the above remark, we then have that p2ip2i+1p2i1=p2i(p2i+1δ1H)p2i1=p2i+1δp2ip_{\leq 2i}p_{2i+1}p_{\leq 2i}^{-1}=p_{\leq 2i}(p_{2i+1}\delta_{1_{H}})p_{\leq 2i}^{-1}=p_{2i+1}\delta_{p_{\leq 2i}}, so f=(p1δp0)(p3δp2)(p2n1δp2n2)f=(p_{1}\delta_{p_{\leq 0}})(p_{3}\delta_{p_{\leq 2}})...(p_{2n-1}\delta_{p_{\leq 2n-2}}).

Note that p2i+2=p2ip2i+2p_{\leq 2i+2}=p_{\leq 2i}p_{2i+2}, and since p2i+2>1p_{2i+2}>1 for 0i<n10\leq i<n-1, we have that p2i<p2i+2p_{\leq 2i}<p_{\leq 2i+2} for 0i<n10\leq i<n-1, and in particular the p2ip_{\leq 2i} are distinct (except possibly that p2n2=p2np_{\leq 2n-2}=p_{\leq 2n}).

Since the {p0,,p2n2}\{p_{\leq 0},...,p_{\leq 2n-2}\} are distinct, then by the above remark, we know that

f=(p1δp0)(p3δp2)(p2n1δp2n2)f=(p_{1}\delta_{p_{\leq 0}})(p_{3}\delta_{p_{\leq 2}})...(p_{2n-1}\delta_{p_{\leq 2n-2}})

is the unique representation of ff as a product of giδhig_{i}\delta_{h_{i}} (except possibly that the p1δp0p_{1}\delta_{p_{\leq 0}} term is equal to 1), so

f(h)={p1 if h=p0p3 if h=p2p2n1 if h=p2n21 otherwisef(h)=\begin{cases}p_{1}&\text{ if }h=p_{\leq 0}\\ p_{3}&\text{ if }h=p_{\leq 2}\\ \vdots\\ p_{2n-1}&\text{ if }h=p_{\leq 2n-2}\\ 1&\text{ otherwise}\end{cases}

as desired.

Proposition 3.27.

Let (G1,P1)(G_{1},P_{1}), (G2,P2)(G_{2},P_{2}) be two ordered groups. Let φ\varphi denote the natural homomorphism from G1G2G_{1}*G_{2} onto G1G2G_{1}\wr G_{2}. If p,qP:=P1P2p,q\in P:=P_{1}*P_{2} and φ(p)=φ(q)\varphi(p)=\varphi(q), then p=qp=q.

In particular, φ\varphi is a strong reduction of G1G2G_{1}*G_{2} onto G1G2G_{1}\wr G_{2}.

Proof.

Since p,qPp,q\in P, then by Lemma 3.20 we may write p=p1p2p2np=p_{1}p_{2}...p_{2n} where pip_{i} is in P1P_{1} if ii is odd, and in P2P_{2} if ii is even, and all pi1p_{i}\neq 1 for 1<i<n1<i<n. Similarly, we may write q=q1q2q2mq=q_{1}q_{2}...q_{2m} where qiq_{i} is in P1P_{1} if ii is odd, and in P2P_{2} if ii is even, and all pi1p_{i}\neq 1 for 1<i<m1<i<m.

Now, by the previous Lemma, we may write φ(p)=(fp,p2n)\varphi(p)=(f_{p},p_{\leq 2n}) and φ(q)=(fq,q2m)\varphi(q)=(f_{q},q_{\leq 2m}) where p2i=p2p4p2ip_{\leq 2i}=p_{2}p_{4}...p_{2i}, q2i=q2q4q2iq_{\leq 2i}=q_{2}q_{4}...q_{2i},

fp(h)={p1 if h=1G2p3 if h=p2p5 if h=p2p4p2n1 if h=p2p2n21 otherwise and fq(h)={q1 if h=1G2q3 if h=q2q5 if h=q2q4q2m1 if h=q2q2m21 otherwisef_{p}(h)=\begin{cases}p_{1}&\text{ if }h=1_{G_{2}}\\ p_{3}&\text{ if }h=p_{2}\\ p_{5}&\text{ if }h=p_{2}p_{4}\\ \vdots\\ p_{2n-1}&\text{ if }h=p_{2}...p_{2n-2}\\ 1&\text{ otherwise}\end{cases}\text{ and }f_{q}(h)=\begin{cases}q_{1}&\text{ if }h=1_{G_{2}}\\ q_{3}&\text{ if }h=q_{2}\\ q_{5}&\text{ if }h=q_{2}q_{4}\\ \vdots\\ q_{2m-1}&\text{ if }h=q_{2}...q_{2m-2}\\ 1&\text{ otherwise}\end{cases}

Since φ(p)=φ(q)\varphi(p)=\varphi(q), then fp=fqf_{p}=f_{q}, so p1=q1p_{1}=q_{1}. Furthermore, {p2,p2p4,,p2p2n2}=supp(fp){1}=supp(fq){1}={q2,q2q4,,q2q2m2}\{p_{2},p_{2}p_{4},...,p_{2}...p_{2n-2}\}=\operatorname{supp}(f_{p})\setminus\{1\}=\operatorname{supp}(f_{q})\setminus\{1\}=\{q_{2},q_{2}q_{4},...,q_{2}...q_{2m-2}\}, and these sets are strictly increasing sequences, so it must be that n=mn=m and p2=q2p_{2}=q_{2}, p2p4=q2q4p_{2}p_{4}=q_{2}q_{4}, and so on. By cancelling common terms, we get that p2i=q2ip_{2i}=q_{2i} for 1i<n1\leq i<n. Returning to our expressions for fpf_{p} and fqf_{q}, we can now compare their values at p2=q2,p2p4=q2q4p_{2}=q_{2},p_{2}p_{4}=q_{2}q_{4}, etc to get that p3=q3,p5=q5p_{3}=q_{3},p_{5}=q_{5}, and so on. Finally, recall that (fp,p2n)=φ(p)=φ(q)=(fq,q2n)(f_{p},p_{\leq 2n})=\varphi(p)=\varphi(q)=(f_{q},q_{\leq 2n}) so p2n=q2np_{\leq 2n}=q_{\leq 2n} and by cancelling the common factor of p2p2n2=q2q2n2p_{2}...p_{2n-2}=q_{2}...q_{2n-2}, we have that p2n=q2np_{2n}=q_{2n}.

Thus for all 1i2n1\leq i\leq 2n, we have that pi=qip_{i}=q_{i}, so p=qp=q, as desired.

For the “in particular”, by Corollary 3.5 it suffices to show that φ\varphi is a bijection of P1P2P_{1}*P_{2} onto P1P2P_{1}\wr P_{2}. It is immediate that φ\varphi is surjective, and by the previous part of the proposition it must be injective. Thus φ\varphi is a bijection of P1P2P_{1}*P_{2} onto P1P2P_{1}\wr P_{2}, so φ\varphi is a strong reduction as desired.

We have done the hard work already, so we may now reap our rewards:

Proposition 3.28.

Let (G,P),(H,Q)(G,P),(H,Q) be ordered groups, and suppose that GG and HH are amenable. Then the natural homomorphism φ:(GH,PQ)(GH,PQ)\varphi:(G*H,P*Q)\rightarrow(G\wr H,P\wr Q) is a strong reduction to an amenable group.

Proof.

By Propositon 3.27, φ\varphi is a strong reduction. By Lemma 3.25, GHG\wr H is amenable, so φ\varphi is a strong reduction to an amenable group.

And combining this with a previous result gives:

Corollary 3.29.

The class of ordered groups which reduce to an amenable group is closed under finite free products.

Proof.

It suffices to check two-term free products. Suppose that (G1,P1)(G_{1},P_{1}) and (G2,P2)(G_{2},P_{2}) are ordered groups with reductions φ1,φ2\varphi_{1},\varphi_{2} to amenable groups (H1,Q1)(H_{1},Q_{1}) and (H2,Q2)(H_{2},Q_{2}) respectively. Let G=G1G2G=G_{1}*G_{2}, P=P1P2P=P_{1}*P_{2}, H=H1H2H=H_{1}*H_{2}, and Q=Q1Q2Q=Q_{1}*Q_{2}.

Now by Lemma 3.22 there is a reduction φ\varphi of (G,P)(G,P) onto (H,Q)(H,Q). By Proposition 3.28, since (H,Q)(H,Q) is the free product of two amenable groups, then there is a strong reduction φ:(H1H2,Q1Q2)(H1H2,Q1Q2)\varphi^{\prime}:(H_{1}*H_{2},Q_{1}*Q_{2})\rightarrow(H_{1}\wr H_{2},Q_{1}\wr Q_{2}) where H1H2H_{1}\wr H_{2} is amenable. Then φφ:(G,P)(F,R)\varphi^{\prime}\varphi:(G,P)\rightarrow(F,R) is a composition of reductions, hence a reduction by Lemma 3.11, and since its range is amenable, (G,P)(G,P) reduces to an amenable group. ∎

3.3 Summary

To summarize our results on reductions to amenable groups:

\AmenableReductionTheorem
Proof.

If (G,P)(G,P) is an ordered group with GG amenable, then idG:(G,P)(G,P)\operatorname{id}_{G}:(G,P)\rightarrow(G,P) is a reduction to an amenable group, so every amenable ordered group contains a reduction to an amenable group.

We’ve seen that this class is closed under hereditary subgroups in Corollary 3.16, direct products in Lemma 3.19, and free products in Corollary 3.29.

Corollary 3.30.

Let G=2G=\mathbb{Z}^{2}*\mathbb{Z} and P=2P=\mathbb{N}^{2}*\mathbb{N}. Then (G,P)(G,P) strongly reduces to an amenable group.

Proof.

By Proposition 3.27, (2,2)(\mathbb{Z}^{2}*\mathbb{Z},\mathbb{N}^{2}*\mathbb{N}) strongly reduces to (2,2)(\mathbb{Z}^{2}\wr\mathbb{Z},\mathbb{N}^{2}\wr\mathbb{N}), which is amenable by Lemma 3.25. ∎

Although the theory we have developed gives a rich class of examples of groups which reduce to amenable groups, there are still many open questions worthy of exploration:

Question 3.31.

Are there any ordered groups that reduce to an amenable group, but do not strongly reduce to an amenable group?

Question 3.32.

Is the class of ordered groups which reduce to amenable groups closed under other group theoretic constructions (direct limits, HNN extensions, amalgamated free products, graph products, etc)?

For the latter question, we would conjecture that the answer is yes, and that this should follow from this class being closed under quotients by “positive relators”:

Conjecture 3.33.

Let (G,P)(G,P) be an ordered group. Let p1,p2Pp_{1},p_{2}\in P be incomparable (meaning p1p2p_{1}\not\leq p_{2} and p1p2p_{1}\not\geq p_{2}), and let NN denote the normal subgroup of GG generated by p1p21p_{1}p_{2}^{-1}. Then:

  1. a.

    If (G,P)(G,P) reduces to an amenable group, then (G/N,P/N)(G/N,P/N) reduces to an amenable group.

  2. b.

    If (G,P)(G,P) strongly reduces to an amenable group, then (G/N,P/N)(G/N,P/N) strongly reduces to an amenable group.

Of these two sub-conjectures, (b) seems more plausible: one could imagine “carrying the quotient forward” into the reduction. However, attempting that in the case of a non-strong reduction of (a) seems more difficult. For instance, we’ve seen that (F2,P2)(F_{2},P_{2}) reduces to (,)(\mathbb{Z},\mathbb{N}), and taking p1=ab,p2=bap_{1}=ab,p_{2}=ba results in the relator ab=baab=ba, so F2/N2F_{2}/N\cong\mathbb{Z}^{2} and P2/N2P_{2}/N\cong\mathbb{N}^{2}, but (2,2)(\mathbb{Z}^{2},\mathbb{N}^{2}) does not reduce to (,)(\mathbb{Z},\mathbb{N}) or any quotient thereof. While (2,2)(\mathbb{Z}^{2},\mathbb{N}^{2}) reduces to a different amenable group (itself), it is not clear in general how one would avoid this type of issue.

4 Gauge-Invariant Uniqueness for PP-graphs

In this section, our goal is to prove a gauge-invariant uniqueness theorem for PP-graphs: there is exactly one representation of a PP-graph which is Λ\Lambda-faithful, tight, and has a gauge coaction. However, we will show in Lemma 4.24 that the gauge-invariant uniqueness theorem can fail in general, so we require an additional hypothesis. That additional hypothesis is that (G,P)(G,P) can reduce to an amenable ordered group, and we will prove a gauge-invariant uniqueness theorem for such PP-graphs in Theorem 1.

4.1 A Co-Universal Algebra

In this section, we will prove a weaker form of a gauge invariant uniqueness theorem which says that there is exactly one representation of a PP-graph which is Λ\Lambda-faithful, tight, and has a normal gauge coaction. This result is a slight generalization of [3, Theorem 5.3] to the context of weakly quasi-lattice ordered groups.

To prove this, we define and study the balanced algebra of a representation, culminating in Theorem 4.15, which classifies the balanced algebras of a graph.

To give a brief summary of the argument:

  1. 1.

    Use the structure of “infinite paths” in Λ\Lambda to construct a tight, Λ\Lambda-faithful representation called the ultrafilter representation ff.

  2. 2.

    For a representation tt, we define the balanced algebra (t)=span¯{tμtν:d(μ)=d(ν)}\mathcal{B}(t)=\overline{\operatorname{span}}\{t_{\mu}t_{\nu}^{*}:d(\mu)=d(\nu)\} (Lemma 4.7). If there is a gauge coaction on tt, there is a conditional expectation Φt:C(t)(t)\Phi_{t}:C^{*}(t)\rightarrow\mathcal{B}(t) which is given by

    Φt(tμtν)={tμtν if d(μ)=d(ν)0 otherwise,\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ otherwise}\end{cases},

    and this conditional expectation is faithful if and only if the gauge coaction is normal (Lemma 4.8). We define a balanced covering to be a map ψst:(t)(s)\psi^{t}_{s}:\mathcal{B}(t)\rightarrow\mathcal{B}(s) given by tμtνsμsνt_{\mu}t_{\nu}^{*}\mapsto s_{\mu}s_{\nu}^{*}.

  3. 3.

    Recalling that 𝒯\mathcal{T} denotes the Toeplitz representation of Λ\Lambda from Definition 2.47, in Theorem 4.15 we show that kerψt𝒯\ker\psi^{\mathcal{T}}_{t} is generated by the “bolts” and range projections it contains. Using this, we show in Lemma 4.16 that any tight representation is balanced covered by any Λ\Lambda-faithful representation.

  4. 4.

    In Lemma 4.21 we show that in the presence of appropriate coactions, we can lift a balanced covering to a covering of the entire algebras.

  5. 5.

    Finally, we show in Theorem 4.22 that there is a unique tight, Λ\Lambda-faithful representation with a normal gauge coaction, and that this representation is co-universal for Λ\Lambda-faithful coacting representations in the sense of [3, Theorem 5.3].

4.1.1 The Ultrafilter Representation

In this subsection, we construct the ultrafilter representation ff of a PP-graph Λ\Lambda, and demonstrate that it is Λ\Lambda-faithful and tight. This construction follows the one in [3, Section 3] and is included here for completeness.

Definition 4.1.

Let (G,P)(G,P) be a WQLO group, and Λ\Lambda a finitely aligned PP-graph. Recall that Λ\Lambda has a partial order given by αβ\alpha\leq\beta if βαΛ\beta\in\alpha\Lambda.

A filter of Λ\Lambda is a nonempty subset UΛU\subseteq\Lambda such that:

  1. (F1)

    If μU\mu\in U and λμ\lambda\leq\mu, then λU\lambda\in U

  2. (F2)

    If μ,νU\mu,\nu\in U, there exists a λU\lambda\in U such that μλ\mu\leq\lambda and νλ\nu\leq\lambda.

Given a filter UU, it is nonempty so it contains some μ\mu. By F1F1, r(μ)Ur(\mu)\in U, and by F2F2 since no distinct vertices have a common extension, then r(μ)r(\mu) is the only vertex in UU. Therefore, for any μ,νU\mu,\nu\in U, r(μ)=r(ν)r(\mu)=r(\nu), so we write r(U)r(U) for this unique vertex.

A ultrafilter of Λ\Lambda is a maximal filter. We denote the set of filters by Λ^\widehat{\Lambda} and the set of ultrafilters by Λ^\widehat{\Lambda}_{\infty}.

Lemma 4.2.

Every filter is contained in an ultrafilter.

Proof.

The proof is a standard Zorn’s Lemma argument: fix a filter UU, and let Λ^U={VΛ^:UV}\widehat{\Lambda}_{U}=\{V\in\widehat{\Lambda}:U\subseteq V\}. We will show that Λ^U\widehat{\Lambda}_{U} contains a maximal element by Zorn’s lemma, and that this maximal element of Λ^U\widehat{\Lambda}_{U} is also maximal in Λ^\widehat{\Lambda}.

To apply Zorn’s Lemma, we must check that Λ^U={VΛ^:UV}\widehat{\Lambda}_{U}=\{V\in\widehat{\Lambda}:U\subseteq V\} is nonempty and that chains in Λ^U\widehat{\Lambda}_{U} have an upper bound in Λ^U\widehat{\Lambda}_{U}. The former is immediate since UΛ^UU\in\widehat{\Lambda}_{U}. If 𝒞\mathcal{C} is a chain in Λ^U\widehat{\Lambda}_{U}, then let W=V𝒞VW=\bigcup_{V\in\mathcal{C}}V, which we will show is a filter. To check F1, if μW\mu\in W and νμ\nu\leq\mu, then μV\mu\in V for some VCV\in C, so νV\nu\in V and thus νW\nu\in W. To check F2, if μ,νW\mu,\nu\in W, then μV1\mu\in V_{1}, νV2\nu\in V_{2} for some V1,V2𝒞V_{1},V_{2}\in\mathcal{C}. Since 𝒞\mathcal{C} is a chain, either V1V2V_{1}\subseteq V_{2} or V2V1V_{2}\subseteq V_{1}. Without loss of generality assuming the former, we have that μ,νV2\mu,\nu\in V_{2}, so there is some λV2W\lambda\in V_{2}\subseteq W with μ,νλ\mu,\nu\leq\lambda as desired. Since WW also contains UU, then WΛ^UW\in\widehat{\Lambda}_{U}, so by Zorn’s Lemma Λ^U\widehat{\Lambda}_{U} contains a maximal element UU_{\infty}.

Finally, we will show that UU_{\infty} is an ultrafilter, meaning that we will check that UU_{\infty} is maximal in Λ^\widehat{\Lambda}. If it were not, there would be some VUUV\supsetneq U_{\infty}\supseteq U, but then we’d have VΛ^UV\in\widehat{\Lambda}_{U}, so UU_{\infty} was not maximal in Λ^U\widehat{\Lambda}_{U}, a contradiction. Thus UU_{\infty} is indeed maximal in Λ^\widehat{\Lambda}, so UΛ^U_{\infty}\in\widehat{\Lambda}_{\infty} as desired.

Lemma 4.3.

Let UU be an ultrafilter, and suppose μU\mu\in U, EμΛE\subset\mu\Lambda, and that EE is exhaustive for μΛ\mu\Lambda. Then EUE\cap U is nonempty.

Proof.

We split into two cases: U is finite and U is infinite. If UU is finite, it has a greatest element ω\omega by F2. Since μω\mu\leq\omega, then ωμΛ\omega\in\mu\Lambda, and since EE is exhaustive for μΛ\mu\Lambda there is some αE\alpha\in E with MCE(α,ω)MCE(\alpha,\omega)\neq\emptyset. That is, we can find some βMCE(α,ω)\beta\in MCE(\alpha,\omega), so define U={ν:νβ}U^{\prime}=\{\nu:\nu\leq\beta\}. Then UU^{\prime} is a filter containing UU, and since UU was an ultrafilter then U=UU^{\prime}=U. Thus βU\beta\in U and by F1, αβU\alpha\leq\beta\in U, as desired.

Suppose instead that UU is infinite. Since Λ\Lambda is countable, so is UU, so let give UU an enumeration U={ν1,ν2,}U=\{\nu_{1},\nu_{2},...\} and without loss of generality suppose μ=ν1\mu=\nu_{1}. By the F2 property, for each nn\in\mathbb{N}, we may choose an ηn\eta_{n} such that ηn1ηn\eta_{n-1}\leq\eta_{n} if n>1n>1 and νiηn\nu_{i}\leq\eta_{n} for all 1in1\leq i\leq n. Then in particular {ηn}n\{\eta_{n}\}_{n\in\mathbb{N}} is a rising sequence of elements of UU such that ηnμΛ\eta_{n}\in\mu\Lambda for all nn, and by the F1 property λU\lambda\in U if and only if ληn\lambda\leq\eta_{n} for some nn. Finally, we may assume that the sequence {ηn}n\{\eta_{n}\}_{n\in\mathbb{N}} is strictly increasing by passing to a subsequence.

Now, for each ηn\eta_{n}, since ηnμΛ\eta_{n}\in\mu\Lambda and EE is exhaustive is μΛ\mu\Lambda, there is an αnE\alpha_{n}\in E such that ηn\eta_{n} has a common extension with αn\alpha_{n}. But EE is finite, so one αE\alpha\in E must appear infinitely many times in the sequence {αn}n\{\alpha_{n}\}_{n\in\mathbb{N}}. Thus there are infinitely many nn with MCE(ηn,α)MCE(\eta_{n},\alpha) nonempty. Our goal now is to show that αU\alpha\in U by constructing a potentially bigger ultrafilter that contains α\alpha, but since UU is already an ultrafilter it must be that α\alpha is already in UU.

To this end, for iji\leq j, if λMCE(ηj,α)\lambda\in MCE(\eta_{j},\alpha), we can factorize λ=ζζ\lambda=\zeta\zeta^{\prime} where d(ζ)=d(ηi)d(α)d(\zeta)=d(\eta_{i})\vee d(\alpha), which implies ζMCE(ηi,α)\zeta\in MCE(\eta_{i},\alpha), and therefore λζΛ\lambda\in\zeta\Lambda for some ζMCE(ηi,α)\zeta\in MCE(\eta_{i},\alpha). Thus

MCE(ηj,α)=ζMCE(ηi,α)[MCE(ηj,α)ζΛ].MCE(\eta_{j},\alpha)=\bigcup_{\zeta\in MCE(\eta_{i},\alpha)}\left[MCE(\eta_{j},\alpha)\cap\zeta\Lambda\right].

Considering ii as fixed but taking the union over all jij\geq i, we have that

jiMCE(ηj,α)\displaystyle\bigcup_{j\geq i}MCE(\eta_{j},\alpha) =\displaystyle= ji(ζMCE(ηi,α)[MCE(ηj,α)ζΛ])\displaystyle\bigcup_{j\geq i}\left(\bigcup_{\zeta\in MCE(\eta_{i},\alpha)}\left[MCE(\eta_{j},\alpha)\cap\zeta\Lambda\right]\right)
=\displaystyle= ζMCE(ηi,α)(jiMCE(ηj,α)ζΛ)\displaystyle\bigcup_{\zeta\in MCE(\eta_{i},\alpha)}\left(\bigcup_{j\geq i}MCE(\eta_{j},\alpha)\cap\zeta\Lambda\right)

Since infinitely many MCE(ηj,α)MCE(\eta_{j},\alpha) are nonempty and the sequence {ηn}n\{\eta_{n}\}_{n\in\mathbb{N}} is strictly increasing, the lefthand side is infinite, and thus the righthand side must have one of the terms jiMCE(ηj,α)ζΛ\bigcup_{j\geq i}MCE(\eta_{j},\alpha)\cap\zeta\Lambda being infinite. In particular, when i=1i=1, there is some ζ1\zeta_{1} such that MCE(ηj,α)ζ1ΛMCE(\eta_{j},\alpha)\cap\zeta_{1}\Lambda is infinite. Then we may repeat a version of this argument: we have

j2MCE(ηj,α)ζ1Λ=ζMCE(η2,α)(j2MCE(ηj,α)ζΛζ1Λ)\bigcup_{j\geq 2}MCE(\eta_{j},\alpha)\cap\zeta_{1}\Lambda=\bigcup_{\zeta\in MCE(\eta_{2},\alpha)}\left(\bigcup_{j\geq 2}MCE(\eta_{j},\alpha)\cap\zeta\Lambda\cap\zeta_{1}\Lambda\right)

so one of the terms j2MCE(ηj,α)ζΛζ1\bigcup_{j\geq 2}MCE(\eta_{j},\alpha)\cap\zeta\Lambda\cap\zeta_{1} must be infinite, so we take ζ2\zeta_{2} to be the element of MCE(η2,α)MCE(\eta_{2},\alpha) giving that infinite term. Next, since ζ1ζ2\zeta_{1}\leq\zeta_{2} then ζ2Λζ1=ζ2Λ\zeta_{2}\Lambda\cap\zeta_{1}=\zeta_{2}\Lambda, and we may repeat this process inductively.

In this way we create a sequence {ζn}n\{\zeta_{n}\}_{n\in\mathbb{N}} with ζnMCE(ηn,α)\zeta_{n}\in MCE(\eta_{n},\alpha) and ζn1ζn\zeta_{n-1}\leq\zeta_{n} for all nn. Finally, let U={λ:λζn for some n}U_{\infty}=\{\lambda:\lambda\leq\zeta_{n}\text{ for some }n\in\mathbb{N}\}. Then it is immediate to verify that UU_{\infty} is a filter. Furthermore, if λU\lambda\in U, then ληn\lambda\leq\eta_{n} for some nn, so ληnζn\lambda\leq\eta_{n}\leq\zeta_{n}, so λU\lambda\in U_{\infty}. That is, UUU\subseteq U_{\infty}, but UU was an ultrafilter, so it must be that U=UU=U_{\infty}. Since αζ1\alpha\leq\zeta_{1}, then αU=U\alpha\in U_{\infty}=U, as desired.

In either case, we have shown that there is some αEU\alpha\in E\cap U. ∎

The following result is [3, Lemma 3.4]:

Lemma 4.4.

Let (G,P)(G,P) be a WQLO group, Λ\Lambda a finitely-aligned PP-graph, λΛ\lambda\in\Lambda and let UU and VV be filters. If r(U)=s(λ)r(U)=s(\lambda) and if λV\lambda\in V, we define

λU:=μU{αλμ} and \lambda\cdot U:=\bigcup_{\mu\in U}\{\alpha\leq\lambda\mu\}\text{ and }
λV:={μΛ:λμV}\lambda^{*}\cdot V:=\{\mu\in\Lambda:\lambda\mu\in V\}

Then λU\lambda\cdot U and λV\lambda^{*}\cdot V are filters and if UU and VV are ultrafilters, then so are λU\lambda\cdot U and λV\lambda^{*}\cdot V. Finally, λ(λU)=U\lambda^{*}\cdot(\lambda\cdot U)=U and λ(λV)=V\lambda\cdot(\lambda^{*}\cdot V)=V.

We are now ready to define our Λ\Lambda-faithful tight representation by letting Λ\Lambda act by translation on its set of ultrafilters:

Lemma 4.5.

Let (G,P)(G,P) be a WQLO group, and Λ\Lambda a finitely-aligned PP-graph. For λΛ\lambda\in\Lambda, define an operator fλ(2(Λ^))f_{\lambda}\in\mathcal{B}(\ell^{2}(\widehat{\Lambda}_{\infty})) by

fλeU={eλU if r(U)=s(λ)0 otherwise.f_{\lambda}e_{U}=\begin{cases}e_{\lambda\cdot U}&\text{ if }r(U)=s(\lambda)\\ 0&\text{ otherwise}\end{cases}.

Then ff is a representation of Λ\Lambda which is Λ\Lambda-faithful and tight. We call ff the ultrafilter representation.

Proof.

Defining the operators fλf_{\lambda} as given above, a typical inner product argument shows that fλeV={eλV if λV0 otherwisef_{\lambda}^{*}e_{V}=\begin{cases}e_{\lambda^{*}\cdot V}&\text{ if }\lambda\in V\\ 0&\text{ otherwise}\end{cases}.

To check the T1 operator, it is immediate that if vΛ0v\in\Lambda^{0}, then fvf_{v} is a projection, namely a projection onto the subspace spanned by the set of ultrafilters containing vv. To show that the {fv}vΛ0\{f_{v}\}_{v\in\Lambda^{0}} are orthogonal, first note that this family commutes, and fix distinct v,wΛ0v,w\in\Lambda^{0}. For any ultrafilter UU, r(U)r(U) is the unique vertex contained in UU, so either vUv\not\in U or wUw\not\in U. In the former case, fwfveU=0f_{w}f_{v}e_{U}=0 and in the latter case fvfweU=0f_{v}f_{w}e_{U}=0, so in either case fvfw=fwfv=0f_{v}f_{w}=f_{w}f_{v}=0.

The T2 operator is immediate.

The T3 operator is immediate from the previous lemma.

For the T4 operator, observe that fλfλf_{\lambda}f_{\lambda}^{*} is projection onto the ultrafilters containing λ\lambda. Therefore, fμfμfνfνf_{\mu}f_{\mu}^{*}f_{\nu}f_{\nu}^{*} is projection onto the ultrafilters containing μ\mu and ν\nu. By the F2 and F1 properties, such an ultrafilter would contain some λMCE(μ,ν)\lambda\in MCE(\mu,\nu). In fact, it would contain exactly one such term, since if it contained two distinct λ1,λ2MCE(μ,ν)\lambda_{1},\lambda_{2}\in MCE(\mu,\nu), by the F2 property it would contain a common extension of λ1\lambda_{1} and λ2\lambda_{2}, but no such extension exists by the uniqueness of factorizations. Thus the set of ultrafilters containing μ\mu and ν\nu is precisely the disjoint union of the ultrafilters containing a λMCE(μ,ν)\lambda\in MCE(\mu,\nu). As operators, this is to say that fμfμfνfν=λMCE(μ,ν)fλfλf_{\mu}f_{\mu}^{*}f_{\nu}f_{\nu}^{*}=\displaystyle\sum_{\lambda\in MCE(\mu,\nu)}f_{\lambda}f_{\lambda}^{*}.

Thus ff is indeed a representation.

To show that ff is Λ\Lambda-faithful, fix some λΛ\lambda\in\Lambda. Since {s(λ)}\{s(\lambda)\} is a filter, then by Lemma 4.2, there is an ultrafilter UU containing s(λ)s(\lambda), in which case r(U)=s(λ)r(U)=s(\lambda). Then, fλeU=eλU0f_{\lambda}\cdot e_{U}=e_{\lambda\cdot U}\neq 0, so fλ0f_{\lambda}\neq 0.

To show that ff is tight, fix some μΛ\mu\in\Lambda and EμΛE\subset\mu\Lambda which is finite and exhaustive for μΛ\mu\Lambda. Let B=αE(fμfμfαfα)B=\displaystyle\prod_{\alpha\in E}(f_{\mu}f_{\mu}^{*}-f_{\alpha}f_{\alpha}^{*}) be the corresponding bolt (see Definition 4.10). We’ll now show that for any ultrafilter UU, BeU=0Be_{U}=0, and this implies that B=0B=0 as desired.

First, if UU does not contain μ\mu, then it also does not contain any of the αE\alpha\in E, so fμfμeU=0=fαfαeUf_{\mu}f_{\mu}^{*}e_{U}=0=f_{\alpha}f_{\alpha}^{*}e_{U}, and thus BeU=0Be_{U}=0.

If instead UU does contain μ\mu, by the previous lemma, then UU contains one of the αE\alpha\in E.

Then, since μ,αU\mu,\alpha\in U, fμfμeU=eU=fαfαUf_{\mu}f_{\mu}^{*}e_{U}=e_{U}=f_{\alpha}f_{\alpha}^{*}U, so BeU=αE(fμfμfαfα)eU=0Be_{U}=\displaystyle\prod_{\alpha\in E}(f_{\mu}f_{\mu}^{*}-f_{\alpha}f_{\alpha}^{*})e_{U}=0.

We have now shown that BeU=0Be_{U}=0 for all ultrafilters UU, so B=0B=0 as an operator in (2(Λ^))\mathcal{B}(\ell^{2}(\widehat{\Lambda}_{\infty})). Thus ff is tight.

Since the ultrafilter representation is Λ\Lambda-faithful and tight, it is natural to ask if it also has a gauge coaction. The answer is often no, as we will show in the next example.

Example 4.6.

Let (G,P)(G,P) be a WQLO group, and suppose that PP is directed, meaning that any two elements of PP have a common upper bound. Examples of directed positive cones include k\mathbb{N}^{k} and any total order. Then letting Λ=P\Lambda=P, Λ\Lambda itself is a filter, so Λ\Lambda is the unique ultrafilter. Thus Λ^={Λ}\widehat{\Lambda}_{\infty}=\{\Lambda\}, so the ultrafilter representation is given by fμ=1f_{\mu}=1 for all μΛ\mu\in\Lambda. In particular, the ultrafilter representation fails to have a gauge coaction if there are distinct μ,νΛ=P\mu,\nu\in\Lambda=P, since such a coaction δ\delta cannot send fμ=fνf_{\mu}=f_{\nu} to the distinct elements fμUd(μ)f_{\mu}\otimes U_{d(\mu)} and fμUd(ν)f_{\mu}\otimes U_{d(\nu)}.

4.1.2 Balanced Algebras and Balanced Coverings

We will now discuss a particularly nice AF subalgebra of a PP-graph algebra. Our approach in this section mimics [3, Section 4], although we use slightly different notation.

Lemma 4.7.

Let (G,P)(G,P) be a WQLO group, Λ\Lambda a finitely-aligned PP-graph, and tt a representation of Λ\Lambda. Then (t)=span¯{tμtν:d(μ)=d(ν)}\mathcal{B}(t)=\overline{\operatorname{span}}\{t_{\mu}t_{\nu}^{*}:d(\mu)=d(\nu)\} is a closed *-subalgebra of C(t)C^{*}(t). We call (t)\mathcal{B}(t) the balanced (sub)algebra.

Proof.

It is immediate that (t)\mathcal{B}(t) is a closed *-invariant subspace, so it suffices to check that it is closed under multiplication. To this end, suppose tμtν,tαtβt_{\mu}t_{\nu}^{*},t_{\alpha}t_{\beta}^{*} satisfy d(μ)=d(ν)d(\mu)=d(\nu) and d(α)=d(β)d(\alpha)=d(\beta). Then,

(tμtν)(tαtβ)\displaystyle(t_{\mu}t_{\nu}^{*})(t_{\alpha}t_{\beta}^{*}) =\displaystyle= (tμtν)(tνtνtαtα)(tαtβ)\displaystyle(t_{\mu}t_{\nu}^{*})(t_{\nu}t_{\nu}^{*}t_{\alpha}t_{\alpha}^{*})(t_{\alpha}t_{\beta}^{*})
=\displaystyle= (tμtν)(λMCE(ν,α)tλtλ)(tαtβ)\displaystyle(t_{\mu}t_{\nu}^{*})(\displaystyle\sum_{\lambda\in MCE(\nu,\alpha)}t_{\lambda}t_{\lambda}^{*})(t_{\alpha}t_{\beta}^{*})
=\displaystyle= λMCE(ν,α)tμ(ν1λ)tβ(α1λ)\displaystyle\displaystyle\sum_{\lambda\in MCE(\nu,\alpha)}t_{\mu(\nu^{-1}\lambda)}t_{\beta(\alpha^{-1}\lambda)}^{*}

where ν1λ\nu^{-1}\lambda denotes the unique path such that ν(ν1λ)=λ\nu(\nu^{-1}\lambda)=\lambda and similarly α1λ\alpha^{-1}\lambda denotes the unique path such that α(α1λ)=λ\alpha(\alpha^{-1}\lambda)=\lambda. Then note that

d(μ(ν1λ))=d(μ)d(ν)1d(λ)=d(λ)=d(α)d(β)1d(λ)=d(α(β1λ))d(\mu(\nu^{-1}\lambda))=d(\mu)d(\nu)^{-1}d(\lambda)=d(\lambda)=d(\alpha)d(\beta)^{-1}d(\lambda)=d(\alpha(\beta^{-1}\lambda))

so the product is in (t)\mathcal{B}(t) as desired. ∎

In the following lemma, we rephrase Lemma 2.26 for the context of a gauge coaction on a PP-graph CC^{*}-algebra:

Lemma 4.8.

Let (G,P)(G,P) be a WQLO group, Λ\Lambda a finitely-aligned PP-graph, tt a representation of Λ\Lambda and suppose tt has a gauge coaction δ\delta. Then there is a conditional expectation Φt:C(t)(t)\Phi_{t}:C^{*}(t)\rightarrow\mathcal{B}(t) such that

Φt(tμtν)={tμtν if d(μ)=d(ν)0 otherwise.\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ otherwise}\end{cases}.
Proof.

Let A=C(t)A=C^{*}(t). For any μ,νΛ\mu,\nu\in\Lambda, observe that δ(tμtν)=tμtνUd(μ)d(ν)1\delta(t_{\mu}t_{\nu}^{*})=t_{\mu}t_{\nu}^{*}\otimes U_{d(\mu)d(\nu)^{-1}}, so tμtνAd(μ)d(ν)1t_{\mu}t_{\nu}^{*}\in A_{d(\mu)d(\nu)^{-1}}. Then tμtνAet_{\mu}t_{\nu}^{*}\in A_{e} if and only if d(μ)=d(ν)d(\mu)=d(\nu), so (t)=Ae\mathcal{B}(t)=A_{e}.

Let Φt:C(t)(t)\Phi_{t}:C^{*}(t)\rightarrow\mathcal{B}(t) be the conditional expectation arising from Lemma 2.26. If d(μ)=d(ν)d(\mu)=d(\nu), then tμtνAet_{\mu}t_{\nu}^{*}\in A_{e}, and Φt\Phi_{t} fixes its range space, so Φt(tμtν)=tμtν\Phi_{t}(t_{\mu}t_{\nu}^{*})=t_{\mu}t_{\nu}^{*}. If d(μ)d(ν)d(\mu)\neq d(\nu), then tμtνAgt_{\mu}t_{\nu}^{*}\in A_{g} for g=d(μ)d(ν)1eg=d(\mu)d(\nu)^{-1}\neq e, and by Lemma 2.26, Φt\Phi_{t} vanishes on all AgA_{g} for geg\neq e, so Φt(tμtν)=0\Phi_{t}(t_{\mu}t_{\nu}^{*})=0, as desired.

Recall from Definition 2.27 that Φt\Phi_{t} is a faithful conditional expectation if and only if δ\delta is a normal coaction.

Definition 4.9.

Let (G,P)(G,P) be a WQLO group, Λ\Lambda a PP-graph, and s,ts,t two representations of Λ\Lambda. Let us say that a balanced covering is a (necessarily surjective) *-homomorphism ψst:(t)(s)\psi^{t}_{s}:\mathcal{B}(t)\rightarrow\mathcal{B}(s) given by

ψst(tμtν)=sμsν\psi^{t}_{s}(t_{\mu}t_{\nu}^{*})=s_{\mu}s_{\nu}^{*}

for all μ,νΛ\mu,\nu\in\Lambda with d(μ)=d(ν)d(\mu)=d(\nu).

If such a *-homomorphism exists, we will write tbalst\geq_{bal}s. It is immediate that bal\geq_{bal} is reflexive and transitive, but it may not be a partial ordering since it may not be antisymmetric (that is, there may be two representations whose balanced algebras are isomorphic, but which are not canonically isomorphic as representations). If tbalst\geq_{bal}s and sbalts\geq_{bal}t, we will write tbalst\cong_{bal}s, and say that their balanced algebras are canonically isomorphic.

When there may be ambiguity between a balanced covering a canonical covering, we will write rep\geq_{rep} to clarify that we mean a canonical covering of the full PP-graph CC^{*}-algebra. Note that a canonical covering gives rise to a balanced covering by restricting the canonical covering to the balanced algebra. That is, trepst\geq_{rep}s implies tbalst\geq_{bal}s. The converse can fail, but we prove a partial converse in Lemma 4.21.

4.1.3 The Kernel of ψt𝒯\psi^{\mathcal{T}}_{t}

This section is devoted to proving the following fact which is our Theorem 4.15: letting 𝒯\mathcal{T} denote the Toeplitz representation from Definition 2.47, then for any representation tt, kerψt𝒯\ker\psi^{\mathcal{T}}_{t} is generated (as an ideal) by the “bolts” (see Definition 4.10) and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains. This is our analogue of [3, Theorem 4.9], although slightly generalized to allow the kerψt𝒯\ker\psi^{\mathcal{T}}_{t} to contain 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}.

This result is highly involved in the sense that the proof is many times longer than the statement. The key to the argument is to show that the property “kerψt𝒯A\ker\psi^{\mathcal{T}}_{t}\cap A is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains” is preserved under direct limits, and that this property is true for a dense collection of subalgebras A(𝒯)A\subseteq\mathcal{B}(\mathcal{T}). This will take several lemmas to establish.

Definition 4.10.

Let (G,P)(G,P) be a WQLO group, let Λ\Lambda be a finitely aligned PP-graph and let tt be a representation of Λ\Lambda. We will say that a bolt in C(t)C^{*}(t) is an element of the form

αE(tμtμtαtα)\displaystyle\prod_{\alpha\in E}(t_{\mu}t_{\mu}^{*}-t_{\alpha}t_{\alpha}^{*})

where μΛ\mu\in\Lambda and EμΛE\subset\mu\Lambda is finite and exhaustive for μΛ\mu\Lambda. We say it is a proper bolt if Es(μ)Λs(μ)E\subset s(\mu)\Lambda\setminus s(\mu).

The term “bolt” comes from a tortured metaphor: a representation is tight if all of its bolts are fastened down (i.e. equal to 0).

Lemma 4.11.

Let (G,P)(G,P) be a WQLO group, let Λ\Lambda be a finitely aligned PP-graph, and let tt a representation of Λ\Lambda.

Let {An}n\{A_{n}\}_{n\in\mathbb{N}} be an increasing sequence of subalgebras of (𝒯)\mathcal{B}(\mathcal{T}), and let A=nAn¯A=\overline{\bigcup_{n\in\mathbb{N}}A_{n}}. Suppose that for each nn\in\mathbb{N}, kerψt𝒯An\ker\psi^{\mathcal{T}}_{t}\cap A_{n} is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains. Then kerψt𝒯A\ker\psi^{\mathcal{T}}_{t}\cap A is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains.

Proof.

Let J=kerψt𝒯AJ=\ker\psi^{\mathcal{T}}_{t}\cap A which is an ideal in AA, and let II denote the ideal generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} contained in JJ. Certainly IJI\subseteq J, and it suffices to show I=JI=J.

By [5, Lemma III.4.1],

J=n(JAn)¯J=\overline{\bigcup_{n\in\mathbb{N}}(J\cap A_{n})}

But note that JAn=kerψt𝒯AAn=kerψt𝒯AnJ\cap A_{n}=\ker\psi^{\mathcal{T}}_{t}\cap A\cap A_{n}=\ker\psi^{\mathcal{T}}_{t}\cap A_{n}, so JAnJ\cap A_{n} is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains. Such a bolt or 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} is certainly in JJAnJ\supseteq J\cap A_{n}, and thus JAnIJ\cap A_{n}\subseteq I, and taking the union and closure, we have

In(JAn)¯=JI\supseteq\overline{\bigcup_{n\in\mathbb{N}}(J\cap A_{n})}=J

so I=JI=J as desired.

We will now build towards our dense family of subalgebras.

Definition 4.12.

Let (G,P)(G,P) be a WQLO group, let Λ\Lambda be a finitely aligned PP-graph, and let SΛS\subseteq\Lambda. We will say that SS is MCE closed if for all μ,νS\mu,\nu\in S, MCE(μ,ν)SMCE(\mu,\nu)\subseteq S.

We will say that SS is substitution closed if for all μ,νS\mu,\nu\in S with d(μ)=d(ν)d(\mu)=d(\nu) and s(μ)=s(ν)s(\mu)=s(\nu), and for all αΛ\alpha\in\Lambda, we have that μαS\mu\alpha\in S implies ναS\nu\alpha\in S.

For SΛS\subseteq\Lambda, we will write D(S)={d(μ):μS}D(S)=\{d(\mu):\mu\in S\}.

As we will see in the next lemma, the condition that SS is MCE closed and substitution closed is the correct condition in order to make AS=span¯{𝒯μ𝒯ν:μ,νS,d(μ)=d(ν)}A_{S}=\overline{\operatorname{span}}\{\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}:\mu,\nu\in S,d(\mu)=d(\nu)\} a subalgebra of the balanced algebra.

Lemma 4.13.

Let (G,P)(G,P) be a WQLO group and let Λ\Lambda be a finitely-aligned PP-graph. For each SΛS\subseteq\Lambda, let AS=span¯{𝒯μ𝒯ν:μ,νS,d(μ)=d(ν)}A_{S}=\overline{\operatorname{span}}\{\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}:\mu,\nu\in S,d(\mu)=d(\nu)\}. Then:

  1. 1.

    If SS is MCE closed and substitution closed, then ASA_{S} is a closed -subalgebra of (𝒯)\mathcal{B}(\mathcal{T}).

  2. 2.

    If SS is MCE closed and substitution closed, and D(S)={d(μ):μS}D(S)=\{d(\mu):\mu\in S\} contains a minimal element mm, then S={μS:d(μ)m}S^{\prime}=\{\mu\in S:d(\mu)\neq m\} is MCE closed and substitution closed.

  3. 3.

    Let S,D(S),m,S,D(S),m, and SS^{\prime} be as above. For any representation tt of Λ\Lambda, if μ,νΛmS={λΛS:d(λ)=m}\mu,\nu\in\Lambda^{m}\cap S=\{\lambda\in\Lambda\cap S:d(\lambda)=m\} with s(μ)=s(ν)s(\mu)=s(\nu), then tμtνψt𝒯(AS)t_{\mu}t_{\nu}^{*}\in\psi^{\mathcal{T}}_{t}(A_{S^{\prime}}) implies that either tμ=0t_{\mu}=0 or there is a finite set EμΛSE\subset\mu\Lambda\cap S^{\prime} which is exhaustive for μΛ\mu\Lambda such that the bolt B=αE(tμtμtαtα)B=\displaystyle\prod_{\alpha\in E}(t_{\mu}t_{\mu}^{*}-t_{\alpha}t_{\alpha}^{*}) is equal to 0.

Proof.

(1): It is immediate that ASA_{S} is a closed *-invariant subspace of (𝒯)\mathcal{B}(\mathcal{T}). It then suffices to check that it is closed under multiplication. If 𝒯μ𝒯ν,𝒯α𝒯βAS\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*},\mathcal{T}_{\alpha}\mathcal{T}_{\beta}^{*}\in A_{S}, then

(𝒯μ𝒯ν)(𝒯α𝒯β)\displaystyle(\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*})(\mathcal{T}_{\alpha}\mathcal{T}_{\beta}^{*}) =\displaystyle= (𝒯μ𝒯ν)λMCE(ν,α)𝒯λ𝒯λ(𝒯α𝒯β)\displaystyle(\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*})\displaystyle\sum_{\lambda\in MCE(\nu,\alpha)}\mathcal{T}_{\lambda}\mathcal{T}_{\lambda}^{*}(\mathcal{T}_{\alpha}\mathcal{T}_{\beta}^{*})
=\displaystyle= λMCE(ν,α)𝒯μ(ν1λ)𝒯α(β1λ).\displaystyle\displaystyle\sum_{\lambda\in MCE(\nu,\alpha)}\mathcal{T}_{\mu(\nu^{-1}\lambda)}\mathcal{T}_{\alpha(\beta^{-1}\lambda)}^{*}.

Note that since SS is MCE closed, then each λMCE(ν,α)\lambda\in MCE(\nu,\alpha) is in SS, and since SS is substitution closed, then μ(ν1λ),α(β1λ)S\mu(\nu^{-1}\lambda),\alpha(\beta^{-1}\lambda)\in S. Thus (𝒯μ𝒯ν)(𝒯α𝒯β)AS(\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*})(\mathcal{T}_{\alpha}\mathcal{T}_{\beta}^{*})\in A_{S}, as desired.

(2): Given μ,νS\mu,\nu\in S^{\prime}, MCE(μ,ν)SMCE(\mu,\nu)\subseteq S since SS is MCE closed, and for λMCE(μ,ν)\lambda\in MCE(\mu,\nu), we have d(λ)d(μ)md(\lambda)\geq d(\mu)\neq m, so d(λ)md(\lambda)\neq m, so λS\lambda\in S^{\prime}. Similarly, given μ,μα,νS\mu,\mu\alpha,\nu\in S^{\prime} with d(μ)=d(ν)d(\mu)=d(\nu), we have that ναS\nu\alpha\in S since SS is substitution closed, but d(να)=d(μα)md(\nu\alpha)=d(\mu\alpha)\neq m, so ναS\nu\alpha\in S^{\prime}. Thus SS^{\prime} is MCE closed, and substitution closed.

(3): Suppose that tμtνψt𝒯(AS)=span¯{tαtβ:α,βS,d(α)=d(β)}t_{\mu}t_{\nu}^{*}\in\psi^{\mathcal{T}}_{t}(A_{S^{\prime}})=\overline{\operatorname{span}}\{t_{\alpha}t_{\beta}^{*}:\alpha,\beta\in S^{\prime},d(\alpha)=d(\beta)\}. Then there is an element of span{tαtβ:α,βS,d(α)=d(β)}\operatorname{span}\{t_{\alpha}t_{\beta}^{*}:\alpha,\beta\in S^{\prime},d(\alpha)=d(\beta)\} within a distance of 11 of tμtνt_{\mu}t_{\nu}^{*}, which is to say there are cic_{i}\in\mathbb{C} and αi,βiS\alpha_{i},\beta_{i}\in S^{\prime} with d(αi)=d(βi)d(\alpha_{i})=d(\beta_{i}) such that writing x=tμtνi=1ncitαitβix=t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{i=1}^{n}c_{i}t_{\alpha_{i}}t_{\beta_{i}}^{*}, then x<1\left\|{x}\right\|<1.

Now, let y=tμtμxy=t_{\mu}t_{\mu}^{*}x, we have ytμ2x<1\left\|{y}\right\|\leq\left\|{t_{\mu}}\right\|^{2}\cdot\left\|{x}\right\|<1, and

y\displaystyle y =\displaystyle= tμtμ(tμtνi=1ncitαitβi)\displaystyle t_{\mu}t_{\mu}^{*}\left(t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{i=1}^{n}c_{i}t_{\alpha_{i}}t_{\beta_{i}}^{*}\right)
=\displaystyle= tμtνi=1ncitμtμtαitβi\displaystyle t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{i=1}^{n}c_{i}t_{\mu}t_{\mu}^{*}t_{\alpha_{i}}t_{\beta_{i}}^{*}
=\displaystyle= tμtνi=1nλMCE(αi,μ)citλtβ(α1λ)\displaystyle t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{i=1}^{n}\displaystyle\sum_{\lambda\in MCE(\alpha_{i},\mu)}c_{i}t_{\lambda}t_{\beta(\alpha^{-1}\lambda)}
=\displaystyle= tμtνj=1Ncjtλjtηj\displaystyle t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{j=1}^{N}c_{j}t_{\lambda_{j}}t_{\eta_{j}}

where the last line is simply relabelling the previous line. Note that the sum is still indeed finite since Λ\Lambda is finitely aligned. Also, since SS is substitution closed and MCE closed, then λj,ηjS\lambda_{j},\eta_{j}\in S for all jj. Since d(λj)d(αi)md(\lambda_{j})\geq d(\alpha_{i})\neq m and mm was minimal, then d(λj)md(\lambda_{j})\neq m and thus λjS\lambda_{j}\in S^{\prime}. Also, λj\lambda_{j} is a common extension of μ\mu and some αi\alpha_{i}, so λjμΛ\lambda_{j}\in\mu\Lambda.

Now let E={λ1,,λN}E=\{\lambda_{1},...,\lambda_{N}\}, which is finite and satisfies EμΛSE\subseteq\mu\Lambda\cap S^{\prime} by our previous remarks. Suppose for the sake of contradiction that EE were not exhaustive for μΛ\mu\Lambda. Then there would be some γμΛ\gamma\in\mu\Lambda such that MCE(γ,λi)=MCE(\gamma,\lambda_{i})=\emptyset for all αiE\alpha_{i}\in E. Then, (tγtγ)ytγ2y<1\left\|{(t_{\gamma}t_{\gamma}^{*})y}\right\|\leq\left\|{t_{\gamma}}\right\|^{2}\left\|{y}\right\|<1, and

(tγtγ)y\displaystyle(t_{\gamma}t_{\gamma}^{*})y =\displaystyle= (tγtγ)(tμtνj=1Ncjtλjtηj)\displaystyle(t_{\gamma}t_{\gamma}^{*})(t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{j=1}^{N}c_{j}t_{\lambda_{j}}t_{\eta_{j}})
=\displaystyle= tγtγtμtν\displaystyle t_{\gamma}t_{\gamma}^{*}t_{\mu}t_{\nu}^{*}

since each MCE(λi,γ)=MCE(\lambda_{i},\gamma)=\emptyset for 1iN1\leq i\leq N. Next, since γμΛ\gamma\in\mu\Lambda, then γ=μγ\gamma=\mu\gamma^{\prime} for some γΛ\gamma^{\prime}\in\Lambda, so we may simplify this expression to

(tγtγ)(tμtν)=tμγtγtμtμtν=tμγtνγ(t_{\gamma}t_{\gamma}^{*})(t_{\mu}t_{\nu}^{*})=t_{\mu\gamma^{\prime}}t_{\gamma^{\prime}}^{*}t_{\mu}^{*}t_{\mu}t_{\nu}^{*}=t_{\mu\gamma^{\prime}}t_{\nu\gamma^{\prime}}^{*}

which is a partial isometry. Since partial isometries have norm either 0 or 1, and we’ve seen it has norm less than 1, we have that 0=tμγtνγ0=t_{\mu\gamma^{\prime}}t_{\nu\gamma^{\prime}}^{*}, and since s(μγ)=s(μ)=s(ν)=s(νγ)s(\mu\gamma^{\prime})=s(\mu)=s(\nu)=s(\nu\gamma^{\prime}), we have 0=ts(μμ1)=ts(μ)=tμtμ=tμ0=t_{s(\mu\mu_{1})}=t_{s(\mu)}=t_{\mu}^{*}t_{\mu}=t_{\mu}, showing that either EE is exhaustive or tμ=0t_{\mu}=0, as desired.

Finally, we wish to show that B=0B=0, where BB is the bolt B=λE(tμtμtλtλ)B=\displaystyle\prod_{\lambda\in E}(t_{\mu}t_{\mu}^{*}-t_{\lambda}t_{\lambda}^{*}). To this end, we will make a similar argument that ByBy is a partial isometry of norm less than 1, so it must be 0. The latter is immediate: ByBy<1\left\|{By}\right\|\leq\left\|{B}\right\|\cdot\left\|{y}\right\|<1. Now let us simplify the expression ByBy:

By\displaystyle By =\displaystyle= B(tμtνj=1Ncjtλjtηj)\displaystyle B(t_{\mu}t_{\nu}^{*}-\displaystyle\sum_{j=1}^{N}c_{j}t_{\lambda_{j}}t_{\eta_{j}})
=\displaystyle= Btμtνj=1Ncj(k=1N(tμtμtλktλk))tλjtηj)\displaystyle Bt_{\mu}t_{\nu}^{*}-\displaystyle\sum_{j=1}^{N}c_{j}\left(\displaystyle\prod_{k=1}^{N}(t_{\mu}t_{\mu}^{*}-t_{\lambda_{k}}t_{\lambda_{k}}^{*})\right)t_{\lambda_{j}}t_{\eta_{j}})
=\displaystyle= Btμtνj=1Ncj(kjN(tμtμtλktλk))((tμtμtλjtλj)tλj)tηj)\displaystyle Bt_{\mu}t_{\nu}^{*}-\displaystyle\sum_{j=1}^{N}c_{j}\left(\displaystyle\prod_{k\neq j}^{N}(t_{\mu}t_{\mu}^{*}-t_{\lambda_{k}}t_{\lambda_{k}}^{*})\right)\left((t_{\mu}t_{\mu}^{*}-t_{\lambda_{j}}t_{\lambda_{j}}^{*})t_{\lambda_{j}}\right)t_{\eta_{j}})
=\displaystyle= Btμtνj=1N0\displaystyle Bt_{\mu}t_{\nu}^{*}-\displaystyle\sum_{j=1}^{N}0
=\displaystyle= Btμtν\displaystyle Bt_{\mu}t_{\nu}^{*}

where the main simplification occurs since (tμtμtλjtλj)tλj=tλjtλj=0(t_{\mu}t_{\mu}^{*}-t_{\lambda_{j}}t_{\lambda_{j}}^{*})t_{\lambda_{j}}=t_{\lambda_{j}}-t_{\lambda_{j}}=0.

Thus By=BtμtνBy=Bt_{\mu}t_{\nu}^{*}, so it is again a partial isometry of norm strictly less than 1, and thus By=0By=0. Finally, since Btμtν=0Bt_{\mu}t_{\nu}^{*}=0, then

0\displaystyle 0 =\displaystyle= (Btμtν)(Btμtν)\displaystyle(Bt_{\mu}t_{\nu}^{*})(Bt_{\mu}t_{\nu}^{*})^{*}
=\displaystyle= BtμtνtνtμB\displaystyle Bt_{\mu}t_{\nu}^{*}t_{\nu}t_{\mu}^{*}B
=\displaystyle= BtμtμB\displaystyle Bt_{\mu}t_{\mu}^{*}B

and recalling that BB is a subprojection of tμtμt_{\mu}t_{\mu}^{*}, we have that 0=B0=B, completing the proof of (3).

The next lemma shows that a sufficient collection of ASA_{S} have the property that ASkerψt𝒯A_{S}\cap\ker\psi^{\mathcal{T}}_{t} is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}s it contains.

Lemma 4.14.

Let (G,P)(G,P) be a WQLO group and let Λ\Lambda be a finitely-aligned PP-graph. For each SΛS\subseteq\Lambda which is MCE closed and substitution closed, let AS=span¯{𝒯μ𝒯ν:μ,νS,d(μ)=d(ν)}A_{S}=\overline{\operatorname{span}}\{\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}:\mu,\nu\in S,d(\mu)=d(\nu)\}. Let tt be a representation of Λ\Lambda, and let ψt𝒯:(𝒯)(t)\psi^{\mathcal{T}}_{t}:\mathcal{B}(\mathcal{T})\rightarrow\mathcal{B}(t) be the balanced covering. If D(S)={d(μ):μS}D(S)=\{d(\mu):\mu\in S\} is finite, then kerψt𝒯AS\ker\psi^{\mathcal{T}}_{t}\cap A_{S} is generated (as an ideal) by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}s it contains.

Proof.

Given a MCEMCE-closed and substitution closed subset SΛS\subseteq\Lambda, let JSJ_{S} denote kerψt𝒯AS\ker\psi^{\mathcal{T}}_{t}\cap A_{S}, which is an ideal in ASA_{S}. Let ISI_{S} denote the ideal in ASA_{S} generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}s contained in JSJ_{S}. Then certainly ISJSI_{S}\subseteq J_{S}, and our claim is equivalent to proving that JS/IS=0J_{S}/I_{S}=0.

We proceed by induction on |D(S)||D(S)|. In the base case of |D(S)|=0|D(S)|=0, the claim is trivial since AS=JS=IS=0A_{S}=J_{S}=I_{S}=0.

In the inductive case, suppose that |D(S)|>0|D(S)|>0, and that the claim is true for all DD^{\prime} with |D|<|D(S)||D^{\prime}|<|D(S)|. Since D(S)D(S) is finite, it has some minimal element mm. First consider the case where Sm={μS:d(μ)=m}S^{m}=\{\mu\in S:d(\mu)=m\} is finite. Note that SmS^{m} is itself substitution closed and MCEMCE-closed and that AS=AS+ASmA_{S}=A_{S^{\prime}}+A_{S^{m}}.

Fix some x+ISJS/ISx+I_{S}\in J_{S}/I_{S}, in which case xJSx\in J_{S}. Since xJSAS=AS+ASmx\in J_{S}\subseteq A_{S}=A_{S^{\prime}}+A_{S^{m}}, we may write x=xm+xx=x_{m}+x^{\prime}, where xmASmx_{m}\in A_{S^{m}} and xASx^{\prime}\in A_{S^{\prime}}. Additionally, since ASmA_{S^{m}} is finite dimensional, then we may write xm=i=1nci𝒯μi𝒯νix_{m}=\displaystyle\sum_{i=1}^{n}c_{i}\mathcal{T}_{\mu_{i}}\mathcal{T}_{\nu_{i}}^{*} where d(μi)=d(νi)=md(\mu_{i})=d(\nu_{i})=m for all ii. There are potentially many such representations of x+ISx+I_{S} as xm+x+ISx_{m}+x^{\prime}+I_{S}. We will assume without loss of generality we have chosen the decomposition with the fewest terms in the sum xm=i=1nci𝒯μi𝒯νix_{m}=\displaystyle\sum_{i=1}^{n}c_{i}\mathcal{T}_{\mu_{i}}\mathcal{T}_{\nu_{i}}^{*}, and our goal is to show that in fact there are zero terms in the sum and thus xm=0x_{m}=0.

To this end, suppose for the sake of contradiction that xm=i=1nci𝒯μi𝒯νix_{m}=\displaystyle\sum_{i=1}^{n}c_{i}\mathcal{T}_{\mu_{i}}\mathcal{T}_{\nu_{i}}^{*} has a nonzero summand, and without loss of generality that it is the first summand. That is, suppose c1𝒯μ1𝒯ν10c_{1}\mathcal{T}_{\mu_{1}}\mathcal{T}_{\nu_{1}}^{*}\neq 0.

Now,

(𝒯μ1𝒯μ1)x(𝒯ν1𝒯ν1)\displaystyle(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*})x(\mathcal{T}_{\nu_{1}}\mathcal{T}_{\nu_{1}}^{*}) =\displaystyle= (𝒯μ1𝒯μ1)(i=1nci𝒯μi𝒯νi+x)(𝒯ν1𝒯ν1)\displaystyle(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*})\left(\displaystyle\sum_{i=1}^{n}c_{i}\mathcal{T}_{\mu_{i}}\mathcal{T}_{\nu_{i}}^{*}+x^{\prime}\right)(\mathcal{T}_{\nu_{1}}\mathcal{T}_{\nu_{1}}^{*})
=\displaystyle= c1𝒯μ1𝒯ν1+(𝒯μ1𝒯μ1)x(𝒯ν1𝒯ν1).\displaystyle c_{1}\mathcal{T}_{\mu_{1}}\mathcal{T}_{\nu_{1}}^{*}+(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*})x^{\prime}(\mathcal{T}_{\nu_{1}}\mathcal{T}_{\nu_{1}}^{*}).

Since xJSkerψt𝒯x\in J_{S}\subseteq\ker\psi^{\mathcal{T}}_{t}, by applying kerψt𝒯\ker\psi^{\mathcal{T}}_{t} to both sides, we have that

0=citμ1tν1+ψt𝒯((𝒯μ1𝒯μ1)x(𝒯ν1𝒯ν1))0=c_{i}t_{\mu_{1}}t_{\nu_{1}}^{*}+\psi^{\mathcal{T}}_{t}((\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*})x^{\prime}(\mathcal{T}_{\nu_{1}}\mathcal{T}_{\nu_{1}}^{*}))

so solving for tμ1tν1t_{\mu_{1}}t_{\nu_{1}}^{*}, we have

tμ1tν1=ψt𝒯(1c1(𝒯μ1𝒯μ1)x(𝒯ν1𝒯ν1))t_{\mu_{1}}t_{\nu_{1}}^{*}=\psi^{\mathcal{T}}_{t}(\frac{-1}{c_{1}}(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*})x^{\prime}(\mathcal{T}_{\nu_{1}}\mathcal{T}_{\nu_{1}}^{*}))

but xASx^{\prime}\in A_{S^{\prime}}, and ASmASASA_{S^{m}}A_{S^{\prime}}\subseteq A_{S^{\prime}}, so the righthand side is an element of ψ𝒯t(AS)\psi^{\mathcal{T}_{t}}(A_{S^{\prime}}). Now by part (3) of Lemma 4.13, we have either tμ1=0t_{\mu_{1}}=0 or there is a finite exhaustive set Eμ1ΛE\subset{\mu_{1}}\Lambda with d(α)S{m}d(\alpha)\in S\setminus\{m\} for all αE\alpha\in E, and such that the bolt B=αE(tμ1tμ1tαtα)B=\displaystyle\prod_{\alpha\in E}(t_{\mu_{1}}t_{\mu_{1}}^{*}-t_{\alpha}t_{\alpha}^{*}) is equal to 0.

In the former case, 𝒯μ1𝒯μ1IS\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*}\in I_{S}, so taking y=c1(𝒯μ1𝒯μ1)𝒯μ1𝒯ν1=c1𝒯μ1𝒯ν1y=c_{1}(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*})\mathcal{T}_{\mu_{1}}\mathcal{T}_{\nu_{1}}^{*}=c_{1}\mathcal{T}_{\mu_{1}}\mathcal{T}_{\nu_{1}}^{*}, we have yISy\in I_{S}, so x+IS=(xmy)+x+ISx+I_{S}=(x_{m}-y)+x^{\prime}+I_{S} is a decomposition with one fewer terms in the sum xmyx_{m}-y, a contradiction. In the latter case, αE(𝒯μ1𝒯μ1𝒯α𝒯α)\displaystyle\prod_{\alpha\in E}(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*}-\mathcal{T}_{\alpha}\mathcal{T}_{\alpha}^{*}) is a bolt in kerπt𝒯\ker\pi^{\mathcal{T}}_{t}, so B1=αE(𝒯μ1𝒯μ1𝒯α𝒯α)ISB_{1}=\displaystyle\prod_{\alpha\in E}(\mathcal{T}_{\mu_{1}}\mathcal{T}_{\mu_{1}}^{*}-\mathcal{T}_{\alpha}\mathcal{T}_{\alpha}^{*})\in I_{S}. Let y=c1B1𝒯μ1𝒯ν1y=c_{1}B_{1}\mathcal{T}_{\mu_{1}}\mathcal{T}_{\nu_{1}}^{*}, and then we may write x+IS=xmy+x+ISx+I_{S}=x_{m}-y+x^{\prime}+I_{S} which will have zeroed out the c1𝒯μ1𝒯ν1c_{1}\mathcal{T}_{\mu_{1}}\mathcal{T}_{\nu_{1}}^{*} summand in the xmx_{m} term (while possibly adding more terms to the xx^{\prime} part, which is acceptable), again a contradiction.

In either case, we found a contradiction with xmx_{m} being the representative with the fewest summands. Thus xmx_{m} had no terms in its summation, and thus xm=0x_{m}=0. That is to say, x=xm+x=xx=x_{m}+x^{\prime}=x^{\prime}, where xASx^{\prime}\in A_{S^{\prime}}, and therefore xASkert𝒯=JSx\in A_{S^{\prime}}\cap\ker^{\mathcal{T}}_{t}=J_{S^{\prime}}. By the inductive hypothesis, since D={d(μ):μS}=D(S){m}D^{\prime}=\{d(\mu):\mu\in S^{\prime}\}=D(S)\setminus\{m\} has fewer terms than D(S)D(S), we have that JS=ISJ_{S^{\prime}}=I_{S^{\prime}} so x+ISIS+IS=ISx+I_{S}\in I_{S^{\prime}}+I_{S}=I_{S}, so JSISJ_{S}\subseteq I_{S}, as desired.

Now in the case that SmS^{m} is not finite, we let FnF_{n} be an increasing sequence of finite subsets of SmS^{m} with Sm=FnS^{m}=\bigcup F_{n}, in which case SFnS^{\prime}\cup F_{n} is MCE closed and substitution-closed, so the previous case applies to ASFnA_{S^{\prime}\cup F_{n}}. But AS=n=1ASFn¯A_{S}=\overline{\bigcup_{n=1}^{\infty}A_{S^{\prime}\cup F_{n}}}, and so the general case follows from Lemma 4.11. ∎

We can now add a second dash of Lemma 4.11 to get our desired result:

Theorem 4.15.

Let (G,P)(G,P) be a WQLO group and let Λ\Lambda be a finitely-aligned PP-graph. Let tt be a representation of Λ\Lambda, and ψt𝒯:(𝒯)(t)\psi^{\mathcal{T}}_{t}:\mathcal{B}(\mathcal{T})\rightarrow\mathcal{B}(t) be the balanced covering. Then kerψt𝒯\ker\psi^{\mathcal{T}}_{t} is generated (as an ideal) by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}s it contains.

Proof.

As above, for any SΛS\subseteq\Lambda, let D(S)={d(μ):μΛ}D(S)=\{d(\mu):\mu\in\Lambda\}. Note that taking a substitution preserves the degree of a path, and taking minimal common extensions of paths can take joins of degrees (that is, replace paths of length pp and qq with paths of length pqp\vee q). Therefore, if FΛF\subseteq\Lambda and SS is the set of all paths in Λ\Lambda obtained by taking substitutions and MCEs of FF, then D(S){p1p2pn:piD(F)}D(S)\subseteq\{p_{1}\vee p_{2}\vee...\vee p_{n}:p_{i}\in D(F)\}. In particular, if FF is finite, then |D(S)|2|F|<|D(S)|\leq 2^{|F|}<\infty, so any finite FΛF\subseteq\Lambda is contained in a set SΛS\subseteq\Lambda which is MCE closed and substitution closed and for which D(S)D(S) is finite.

Since Λ\Lambda is countable, we can enumerate the elements as λ1,λ2,\lambda_{1},\lambda_{2},.... For each nn\in\mathbb{N}, let Fn={λ1,,λn}F_{n}=\{\lambda_{1},...,\lambda_{n}\}, and let SnS_{n} denote the set of all paths in Λ\Lambda obtained by taking substitutions and MCEs of FnF_{n}.

Then SnS_{n} is substitution closed and MCE closed and by the above argument, D(Sn)<D(S_{n})<\infty. Therefore, by Lemma 4.14, kerψt𝒯ASn\ker\psi^{\mathcal{T}}_{t}\cap A_{S_{n}} is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains.

But recall that (𝒯)=span¯{𝒯μ𝒯ν:d(μ)=d(ν)}\mathcal{B}(\mathcal{T})=\overline{\operatorname{span}}\{\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}:d(\mu)=d(\nu)\}, and for any μ,νΛ\mu,\nu\in\Lambda with d(μ)=d(ν)d(\mu)=d(\nu), since we have an enumeration, there exists some m,nm,n\in\mathbb{N} with μ=λm,ν=λn\mu=\lambda_{m},\nu=\lambda_{n}, so μ,νFmax(m,n)Smax(m,n)\mu,\nu\in F_{\max(m,n)}\subset S_{\max(m,n)}, so 𝒯μ𝒯νASmax(m,n)\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}\in A_{S_{\max(m,n)}}. Thus (𝒯)=nASn¯\mathcal{B}(\mathcal{T})=\overline{\bigcup_{n\in\mathbb{N}}A_{S_{n}}}.

Then by Lemma 4.11, kerψt𝒯(𝒯)=kerψt𝒯\ker\psi^{\mathcal{T}}_{t}\cap\mathcal{B}(\mathcal{T})=\ker\psi^{\mathcal{T}}_{t} is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} it contains. ∎

The above result classifies ideals in the balanced algebra (𝒯)\mathcal{B}(\mathcal{T}), and when combined with the Factors Through Theorem (Lemma 2.8) shows that two balanced algebras are isomorphic if and only if they have the same set of bolts and range projections which are set equal to 0. We make use of this fact in the following lemma:

Lemma 4.16.

Let (G,P)(G,P) be a WQLO group, Λ\Lambda a PP-graph, and let ss and tt be two representations of (G,P)(G,P). Then

  1. 1.

    If tt is tight and ss is Λ\Lambda-faithful, then tbalst\leq_{bal}s.

  2. 2.

    If tt and ss are both tight and Λ\Lambda-faithful, then tbalst\cong_{bal}s.

Proof.

For the first claim, let 𝒯\mathcal{T} denote the universal representation of Λ\Lambda. Then, there are balanced coverings ψs𝒯:(𝒯)(s)\psi^{\mathcal{T}}_{s}:\mathcal{B}(\mathcal{T})\rightarrow\mathcal{B}(s) and ψt𝒯:(𝒯)(t)\psi^{\mathcal{T}}_{t}:\mathcal{B}(\mathcal{T})\rightarrow\mathcal{B}(t). We wish to use the Factors Through Theorem (Lemma 2.8) to conclude there is a balanced covering ψts:(s)(t)\psi^{s}_{t}:\mathcal{B}(s)\rightarrow\mathcal{B}(t) given by sμsνtμtνs_{\mu}s_{\nu}^{*}\rightarrow t_{\mu}t_{\nu}^{*}, which we can do if and only if kerψs𝒯kerψt𝒯\ker\psi^{\mathcal{T}}_{s}\subseteq\ker\psi^{\mathcal{T}}_{t}.

By Theorem 4.15, kerψs𝒯\ker\psi^{\mathcal{T}}_{s} and kerψt𝒯\ker\psi^{\mathcal{T}}_{t} are generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} they contain. Since ss is Λ\Lambda-faithful, its kernel contains no 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*} (and possibly some bolts), and since tt is tight, its kernel contains every bolt (and possibly some 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}). Thus kerψs𝒯kerψt𝒯\ker\psi^{\mathcal{T}}_{s}\subseteq\ker\psi^{\mathcal{T}}_{t}, so tbalst\leq_{bal}s as desired.

For the second claim, by applying (1) twice, we have tbalst\leq_{bal}s and sbalts\leq_{bal}t, so tbalst\cong_{bal}s. ∎

4.1.4 Normalizations of Coactions

We will now touch briefly on the normalization of a coaction. For a more thorough introduction, the reader is directed to [6, Appendix A.7]

Remark 4.17.

Given a discrete coaction (A,G,δ)(A,G,\delta), define jA:AACr(G)j_{A}:A\rightarrow A\otimes C^{*}_{r}(G) by jA=(idAπLU)δj_{A}=(\operatorname{id}_{A}\otimes\pi^{U}_{L})\circ\delta, which is a *-homomorphism.

Let An=jA(A)A/kerjAA^{n}=j_{A}(A)\cong A/\ker j_{A}. Then by Lemma A.55 and Definition A.56 of [6] there is a coaction (An,G,δn)(A^{n},G,\delta^{n}) which is normal, and such that jAj_{A} is δ\delta-δn\delta^{n} equivariant, meaning that δnjA=(jAidG)δ\delta^{n}\circ j_{A}=(j_{A}\otimes\operatorname{id}_{G})\circ\delta. We call (An,G,δn)(A^{n},G,\delta^{n}) the normalization of (A,G,δ)(A,G,\delta).

Note that [6] uses a definition of normality which is equivalent to ours, but not identical. Readers may wish to read [17, Lemma 1.4] for a proof of the equivalence of the two definitions.

The following lemma is analogous to Lemma 2.50 but refers to a “normalization” process for representations instead of a “coactionization” process. Note that while coactionization made the representation larger with respect to rep\leq_{rep}, normalization will make the representation smaller.

Lemma 4.18.

Let (G,P)(G,P) be a WQLO group, and Λ\Lambda a PP-graph. Let tt be a representation of Λ\Lambda with a gauge coaction δ\delta. Then,

  1. 1.

    Let jA:C(t)C(t)nj_{A}:C^{*}(t)\rightarrow C^{*}(t)^{n} denote the quotient map taking the cosystem (C(t),G,δ)(C^{*}(t),G,\delta) to its normalization (C(t)n,G,δn)(C^{*}(t)^{n},G,\delta^{n}), and for λΛ\lambda\in\Lambda let t~λ:=jA(tλ)\tilde{t}_{\lambda}:=j_{A}(t_{\lambda}). Then t~\tilde{t} is a representation of Λ\Lambda.

  2. 2.

    There is a canonical covering tt~t\mapsto\tilde{t}.

  3. 3.

    The coaction δn\delta^{n} on C(t~)C^{*}(\tilde{t}) is normal and is a gauge coaction.

  4. 4.

    tt is canonically isomorphic to t~\tilde{t} if and only if δ\delta is a normal coaction.

  5. 5.

    There is a balanced covering ψt~t:(t)(t~)\psi^{t}_{\tilde{t}}:\mathcal{B}(t)\rightarrow\mathcal{B}(\tilde{t}), which is a balanced isomorphism.

  6. 6.

    t~\tilde{t} is Λ\Lambda-faithful (respectively, tight) if and only if tt is.

Definition 4.19.

We call the representation t~\tilde{t} the given in the previous lemma the normalization of tt.

Proof of 4.18.

(1) The fact that t~λ\tilde{t}_{\lambda} is a representation is immediate since it is the image of a representation under a homomorphism.

(2) jAj_{A} is a canonical covering.

(3) δn\delta^{n} is normal by construction. To show it is a gauge coaction, recall from Definition A.56 and Lemma A.55 of [6] that the canonical covering jAj_{A} is δ\delta-δn\delta^{n} equivariant, meaning that δnjA=(jAidG)δ\delta^{n}\circ j_{A}=(j_{A}\otimes\operatorname{id}_{G})\circ\delta. Then, for any μΛ\mu\in\Lambda,

δn(t~μ)\displaystyle\delta^{n}(\tilde{t}_{\mu}) =\displaystyle= δn(jA(tμ))\displaystyle\delta^{n}(j_{A}(t_{\mu}))
=\displaystyle= (jAidG)δ(tμ)\displaystyle(j_{A}\otimes\operatorname{id}_{G})\circ\delta(t_{\mu})
=\displaystyle= (jAidG)(tμUd(μ))\displaystyle(j_{A}\otimes\operatorname{id}_{G})(t_{\mu}\otimes U_{d(\mu)})
=\displaystyle= t~μUd(μ)\displaystyle\tilde{t}_{\mu}\otimes U_{d(\mu)}

as desired.

(4) By [17, Lemma 1.4], a coaction is normal (in the sense of its conditional expectation Φ\Phi being faithful) if and only if jAj_{A} is injective, which is equivalent to jAj_{A} being a canonical isomorphism.

(5) The balanced covering ψt~t\psi^{t}_{\tilde{t}} exists because it is a restriction of the canonical covering jAj_{A} from part (2). Since it is a balanced covering, it is surjective, so it suffices to show ψt~t\psi^{t}_{\tilde{t}} is injective. To this end, recall that ψt~t=jA(t)\psi^{t}_{\tilde{t}}=j_{A}\mid_{\mathcal{B}(t)} and that jA:=(idAπLU)δj_{A}:=(\operatorname{id}_{A}\otimes\pi^{U}_{L})\circ\delta. Now for x(t)x\in\mathcal{B}(t) we have δ(x)=xUe\delta(x)=x\otimes U_{e}, so jA(x)=((idAπLU)δ)(x)=xLe=x1j_{A}(x)=((\operatorname{id}_{A}\otimes\pi^{U}_{L})\circ\delta)(x)=x\otimes L_{e}=x\otimes 1. Then since \left\|{\cdot}\right\| is a CC^{*}-cross norm, jA(x)=x1=x\left\|{j_{A}(x)}\right\|=\left\|{x}\right\|\cdot\left\|{1}\right\|=\left\|{x}\right\|, so x=0x=0 if and only if ψt~t(x)=jA(x)=0\psi^{t}_{\tilde{t}}(x)=j_{A}(x)=0, as desired.

(6) Having proven (5), this is immediate: since tightness is a relation among elements of the balanced algebra, tt is tight if and only if t~\tilde{t} is tight. Similarly, tt is Λ\Lambda-faithful if and only if every tλ0t_{\lambda}\neq 0 if and only if every tλtλ0t_{\lambda}^{*}t_{\lambda}\neq 0, which is a relation among elements of the balanced algebra, so tt is Λ\Lambda-faithful if and only if t~\tilde{t} is.

Corollary 4.20.

Let (G,P)(G,P) be a WQLO group, and let Λ\Lambda be a finitely aligned PP-graph. For any Λ\Lambda-faithful, tight representation tt of Λ\Lambda, the normalization of the coactionization of tt is a Λ\Lambda-faithful, tight representation with a normal gauge coaction.

Proof.

Recall that tt^{\prime} denotes the coactionization of tt as in Lemma 2.50 and let t~\widetilde{t^{\prime}} denoting the normalization of tt^{\prime} as in Lemma 4.18. By those two lemmas, t~\widetilde{t^{\prime}} has a normal gauge coaction, and since tt is Λ\Lambda-faithful and tight, so is t~\widetilde{t^{\prime}}. ∎

4.1.5 The Co-Universal Algebra

The following result shows that a balanced covering extends to a covering of the entire algebras in the context of a normal coaction:

Lemma 4.21.

Let (G,P)(G,P) be a WQLO group, let Λ\Lambda be a finitely-aligned PP-graph, and let ss and tt be two representations of Λ\Lambda. If sbalts\leq_{bal}t, ss and tt have gauge coactions, and the gauge coaction on ss is normal, then srepts\leq_{rep}t.

Proof.

Our argument will make use of several functions, organized according to the following diagram (which will commute):

C(𝒯){C^{*}(\mathcal{T})}C(t){C^{*}(t)}(t){\mathcal{B}(t)}C(s){C^{*}(s)}(s){\mathcal{B}(s)}πt𝒯\scriptstyle{\pi^{\mathcal{T}}_{t}}πs𝒯\scriptstyle{\pi^{\mathcal{T}}_{s}}Φt\scriptstyle{\Phi_{t}}πst\scriptstyle{\exists\pi^{t}_{s}}ψst\scriptstyle{\psi^{t}_{s}}Φs\scriptstyle{\Phi_{s}}

Here 𝒯\mathcal{T} denotes the universal representation of Λ\Lambda, and πt𝒯\pi^{\mathcal{T}}_{t} and πs𝒯\pi^{\mathcal{T}}_{s} are the canonical coverings of C(t)C^{*}(t) and C(s)C^{*}(s) respectively. We wish to show the existence of the canonical covering πst\pi^{t}_{s}. The algebras (t)\mathcal{B}(t) and (s)\mathcal{B}(s) are the balanced subalgebras as in Lemma 4.7 and the maps Φt:C(t)(t)\Phi_{t}:C^{*}(t)\rightarrow\mathcal{B}(t) and Φs:C(s)(s)\Phi_{s}:C^{*}(s)\rightarrow\mathcal{B}(s) are the conditional expectations arising from Lemma 4.8. From that lemma, these conditional expectations satisfy

Φt(tμtν)={tμtν if d(μ)=d(ν)0 otherwise and similarly Φs(sμsν)={sμsν if d(μ)=d(ν)0 otherwise.\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ otherwise}\end{cases}\text{ and similarly }\Phi_{s}(s_{\mu}s_{\nu}^{*})=\begin{cases}s_{\mu}s_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ otherwise}\end{cases}.

Also note that by Definition 2.27, Φs\Phi_{s} is faithful because the gauge coaction on ss is normal. Finally, ψst\psi^{t}_{s} is the balanced covering that exists because tbalst\geq_{bal}s.

We will now verify that this diagram commutes. That is, we will show that

ψstΦtπt𝒯=Φsπs𝒯.\psi^{t}_{s}\circ\Phi_{t}\circ\pi^{\mathcal{T}}_{t}=\Phi_{s}\circ\pi^{\mathcal{T}}_{s}.

Since C(𝒯)=span¯{𝒯μ𝒯ν:μ,νΛ}C^{*}(\mathcal{T})=\overline{\operatorname{span}}\{\mathcal{T}_{\mu}\mathcal{T}_{\nu}:\mu,\nu\in\Lambda\} and all the maps are linear and continuous, it suffices to check that the two functions agree on each 𝒯μ𝒯ν\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}. To this end, we fix some μ,νΛ\mu,\nu\in\Lambda, and we have

ψstΦtπt𝒯(𝒯μ𝒯ν)\displaystyle\psi^{t}_{s}\circ\Phi_{t}\circ\pi^{\mathcal{T}}_{t}(\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}) =\displaystyle= ψstΦt(tμtν)\displaystyle\psi^{t}_{s}\circ\Phi_{t}(t_{\mu}t_{\nu}^{*})
=\displaystyle= ψst({tμtν if d(μ)=d(ν)0 otherwise)\displaystyle\psi^{t}_{s}\left(\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ otherwise}\end{cases}\right)
=\displaystyle= {sμsν if d(μ)=d(ν)0 otherwise\displaystyle\begin{cases}s_{\mu}s_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ otherwise}\end{cases}
=\displaystyle= Φs(sμsν)\displaystyle\Phi_{s}(s_{\mu}s_{\nu}^{*})
=\displaystyle= Φsπs𝒯(𝒯μ𝒯ν)\displaystyle\Phi_{s}\circ\pi^{\mathcal{T}}_{s}(\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*})

as desired.

We will now show that kerπt𝒯kerπs𝒯\ker\pi^{\mathcal{T}}_{t}\subseteq\ker\pi^{\mathcal{T}}_{s}. To this end, suppose that xkerπt𝒯x\in\ker\pi^{\mathcal{T}}_{t}. Then xxkerπt𝒯x^{*}x\in\ker\pi^{\mathcal{T}}_{t}, so

0\displaystyle 0 =\displaystyle= ψstΦtπt𝒯(xx)\displaystyle\psi^{t}_{s}\circ\Phi_{t}\circ\pi^{\mathcal{T}}_{t}(x^{*}x)
=\displaystyle= Φsπs𝒯(xx)\displaystyle\Phi_{s}\circ\pi^{\mathcal{T}}_{s}(x^{*}x)

and since πs𝒯(xx)0\pi^{\mathcal{T}}_{s}(x^{*}x)\geq 0 and Φs\Phi_{s} is faithful, then we must have that πs𝒯(xx)=0\pi^{\mathcal{T}}_{s}(x^{*}x)=0, so πs𝒯(x)=0\pi^{\mathcal{T}}_{s}(x)=0, and thus xkerπs𝒯x\in\ker\pi^{\mathcal{T}}_{s} as desired.

Thus by Lemma 2.8 (the Factors Through Theorem), there exists a map πst\pi^{t}_{s} satisfying πs𝒯=πstπt𝒯\pi^{\mathcal{T}}_{s}=\pi_{s}^{t}\circ\pi^{\mathcal{T}}_{t}, from which it is immediate that πst\pi^{t}_{s} is a canonical covering. Thus srepts\leq_{rep}t.

Now we may prove the main theorem of this section, which is a slight generalization of [3, Theorem 5.3] from the context of quasi-lattice ordered groups to weakly quasi-lattice ordered groups. We have also rephrased the result:

Theorem 4.22.

Let (G,P)(G,P) be a WQLO group, and let Λ\Lambda be a finitely aligned PP-graph. If ss is a Λ\Lambda-faithful, tight representation with a normal gauge coaction, and tt is a Λ\Lambda-faithful gauge-coacting representation, then sts\leq t.

In particular, there is a unique representation SS of Λ\Lambda which is Λ\Lambda-faithful, tight, and has a normal gauge coaction, and StS\leq t for any Λ\Lambda-faithful gauge-coacting representation tt.

We write Cmin(Λ)C^{*}_{min}(\Lambda) for C(S)C^{*}(S), and call Cmin(Λ)C^{*}_{min}(\Lambda) the co-universal algebra of the graph.

Proof.

Suppose that ss was a Λ\Lambda-faithful, tight representation with a normal gauge coaction, and that tt was any Λ\Lambda-faithful gauge-coacting representation. Since ss is tight and tt is Λ\Lambda-faithful, by Lemma 4.16, sbalts\leq_{bal}t. Then since ss has a normal gauge coaction and tt has a gauge coaction, by Lemma 4.21, srepts\leq_{rep}t.

To show the existence of such an SS, recall that ff denotes the ultrafilter representation from Lemma 4.5, which is Λ\Lambda-faithful and tight by that lemma. By Corollary 4.20, S:=f~S:=\widetilde{f^{\prime}} is a Λ\Lambda-faithful, tight representation with a normal gauge coaction.

If tt were another Λ\Lambda-faithful, tight representation that had a normal gauge coaction, then by the first part of the claim trepSt\leq_{rep}S and SreptS\leq_{rep}t, so trepSt\cong_{rep}S. That is, SS is unique up to canonical isomorphism.

4.2 The Tight Algebra

The following result is a relatively straightforward combination of the properties of Cmin(Λ)C^{*}_{min}(\Lambda) and the construction of a universal algebra (Lemma 2.9).

Proposition 4.23.

Let (G,P)(G,P) be a WQLO group, and Λ\Lambda a finitely-aligned PP-graph.

  1. 1.

    There is a tight representation TT of Λ\Lambda which is universal for tight representations. We will denote the algebra generated by this representation as Ctight(Λ)C^{*}_{tight}(\Lambda).

  2. 2.

    There is a canonical covering πST:Ctight(Λ)Cmin(Λ)\pi_{S}^{T}:C^{*}_{tight}(\Lambda)\rightarrow C^{*}_{min}(\Lambda).

  3. 3.

    TT is Λ\Lambda-faithful.

  4. 4.

    TT has a gauge coaction δ\delta.

  5. 5.

    The gauge coaction δ\delta on Ctight(Λ)C^{*}_{tight}(\Lambda) is normal if and only if πST\pi_{S}^{T} is an isomorphism, in which case Ctight(Λ)Cmin(Λ)C^{*}_{tight}(\Lambda)\cong C^{*}_{min}(\Lambda).

  6. 6.

    If δ\delta is normal, there is a gauge-invariant uniqueness theorem of this form: if tt is another Λ\Lambda-faithful, gauge coacting, tight representation, then Cmin(Λ)C^{*}_{min}(\Lambda) is canonically isomorphic to C(t)C^{*}(t).

Proof.

For (1), the relators (T1)-(T4) and the tightness condition are all polynomial relations in the generators, and (T1)(T1) and (T3)(T3) together imply that all the generators are partial isometries, so by Lemma 2.9, there is a universal CC^{*}-algebra for such tight representations. Let us denote this representation by TT and the algebra it generates by Ctight(Λ)C^{*}_{tight}(\Lambda).

For (2), by Theorem 4.22, the representation SS which generates Cmin(Λ)C^{*}_{min}(\Lambda) is a tight representation, so by the universality of TT, there is a canonical covering πST:Ctight(Λ)=C(T)C(S)=Cmin(Λ)\pi^{T}_{S}:C^{*}_{tight}(\Lambda)=C^{*}(T)\rightarrow C^{*}(S)=C^{*}_{min}(\Lambda).

For (3), by Theorem 4.22(1), SS is Λ\Lambda-faithful, so each SλS_{\lambda} is nonzero, and since πST(Tλ)=Sλ\pi^{T}_{S}(T_{\lambda})=S_{\lambda}, then each TλT_{\lambda} is nonzero as well.

For (4), let TT^{\prime} denote the coactionization of TT given by Proposition 2.50. By part (7) of that proposition, since TT is tight, then TT^{\prime} is also tight, and therefore by the universality of TT, there is a canonical covering TλTλ=TλUd(λ)T_{\lambda}\mapsto T_{\lambda}^{\prime}=T_{\lambda}\otimes U_{d(\lambda)}, which is the desired gauge coaction.

For (5), if STS\cong T, then the gauge coaction on TT is normal since the gauge coaction on SS is normal. Conversely, if the gauge coaction on TT is normal, then by the uniqueness part of Theorem 4.22, the canonical covering πST\pi_{S}^{T} is an isomorphism.

For (6), suppose that δ\delta is normal and tt is a Λ\Lambda-faithful gauge coacting, tight representation. Then by the universality of Ctight(Λ)=C(T)C^{*}_{tight}(\Lambda)=C^{*}(T), TreptT\geq_{rep}t (here we are using the partial order notation of Lemma 2.48). Similarly, by the co-universality of Cmin(Λ)=C(S)C^{*}_{min}(\Lambda)=C^{*}(S), trepSt\geq_{rep}S. Finally, by (5) since δ\delta is normal, then TST\cong S. Putting these together, we have that STreptrepSS\cong T\geq_{rep}t\geq_{rep}S, so tSt\cong S.

This is a pleasant result if δ\delta is normal, but there are examples arising from non-amenable groups where δ\delta is not normal:

Lemma 4.24.

Let (G,P)(G,P) be a group with a total ordering (meaning G=PP1G=P\cup P^{-1} in addition to PP1={1}P\cap P^{-1}=\{1\}), and let Λ=P\Lambda=P, thought of as a small category with one object and with degree functor idP\operatorname{id}_{P}. Then any tight representation tt of Λ=P\Lambda=P extends uniquely to a unitary representation t¯\bar{t} of GG by t¯g={tg if gPtg1 if gP1\bar{t}_{g}=\begin{cases}t_{g}&\text{ if }g\in P\\ t_{g^{-1}}^{*}&\text{ if }g\in P^{-1}\end{cases}.

Therefore, Ctight(Λ)C(G)C^{*}_{tight}(\Lambda)\cong C^{*}(G), and in particular the gauge coaction on Ctight(Λ)C^{*}_{tight}(\Lambda) is normal if and only if GG is amenable.

Proof.

Let (G,P)(G,P) be a group with a total ordering, and let Λ=P\Lambda=P, thought of as a small category with the natural degree functor. Let tt be a tight representation of PP.

We will first show that tpt_{p} is a unitary for each pPp\in P. For all p,qPp,q\in P, max(p,q)p,q\max(p,q)\geq p,q, so pp and qq have a common extension. That is, every pPp\in P is exhaustive for PP, so by tightness, we have that 1=tete=tptp1=t_{e}t_{e}^{*}=t_{p}t_{p}^{*} for all pPp\in P. Since also tptp=ts(p)=te=1t_{p}t_{p}^{*}=t_{s(p)}=t_{e}=1 by the (T3) relator, this says that the tpt_{p}s are unitaries.

Now, extend tt to t¯\bar{t} on all of G=PP1G=P\cup P^{-1} by t¯g={tg if gPtg1 if gP1\bar{t}_{g}=\begin{cases}t_{g}&\text{ if }g\in P\\ t_{g^{-1}}^{*}&\text{ if }g\in P^{-1}\end{cases}, or equivalently t¯p=tp\bar{t}_{p}=t_{p} and t¯p1=tp\bar{t}_{p^{-1}}=t_{p}^{*} for pPp\in P. We wish to show that t¯\bar{t} is a representation of all of GG. To this end, it suffices to check that it is a group homomorphism, and since it extends a representation of PP, it suffices to check that multiplication is correct for one element in PP and one element in P1P^{-1}.

To this end, suppose p,qPp,q\in P. We have several cases depending on which products are positive. If 1p1q1\leq p^{-1}q, then

t¯p1t¯q=tptq=tptptp1q=tp1q=t¯p1q\bar{t}_{p^{-1}}\bar{t}_{q}=t_{p}^{*}t_{q}=t_{p}^{*}t_{p}t_{p^{-1}q}=t_{p^{-1}q}=\bar{t}_{p^{-1}q}

and if instead 1p1q1\geq p^{-1}q, then 1q1p1\leq q^{-1}p, so

t¯p1t¯q=tptq=(tqtq1p)tq=tq1ptqtq=tq1p=t¯p1q\bar{t}_{p^{-1}}\bar{t}_{q}=t_{p}^{*}t_{q}=(t_{q}t_{q^{-1}p})^{*}t_{q}=t_{q^{-1}p}^{*}t_{q}^{*}t_{q}=t_{q^{-1}p}^{*}=\bar{t}_{p^{-1}q}

so in either case t¯p1t¯q=t¯p1q\bar{t}_{p^{-1}}\bar{t}_{q}=\bar{t}_{p^{-1}q}.

For products the other way, if pq11pq^{-1}\geq 1, then

t¯pt¯q1=tptq=tpq1tqtq=tpq1=t¯pq1\bar{t}_{p}\bar{t}_{q^{-1}}=t_{p}t_{q}^{*}=t_{pq^{-1}}t_{q}t_{q}^{*}=t_{pq^{-1}}=\bar{t}_{pq^{-1}}

while if pq11pq^{-1}\leq 1, then qp11qp^{-1}\geq 1, so

t¯pt¯q1=tptq=tp(tqp1tp)=tptptqp1=tqp1=t¯pq1\bar{t}_{p}\bar{t}_{q^{-1}}=t_{p}t_{q}^{*}=t_{p}(t_{qp^{-1}}t_{p})^{*}=t_{p}t_{p}^{*}t_{qp^{-1}}^{*}=t_{qp^{-1}}^{*}=\bar{t}_{pq^{-1}}

so in either case t¯pt¯q1=t¯pq1\bar{t}_{p}\bar{t}_{q^{-1}}=\bar{t}_{pq^{-1}}. Thus t¯\bar{t} is indeed a representation of GG.

To show that t¯\bar{t} is the unique extension of tt to a unitary representation of GG, observe that if t~\tilde{t} were another such representation, then for all pPp\in P, 1=t~pt~p1=tpt~p11=\tilde{t}_{p}\tilde{t}_{p^{-1}}=t_{p}\tilde{t}_{p^{-1}}, so since tpt_{p} is a unitary we get that tp=t~p1t_{p}^{*}=\tilde{t}_{p^{-1}}, so t~=t¯\tilde{t}=\bar{t}.

Thus the algebra generated by the universal tight representation TT is the algebra generated by the universal group representation. That is, Ctight(Λ)=C(G)C^{*}_{tight}(\Lambda)=C^{*}(G), and the gauge coaction on Ctight(Λ)C^{*}_{tight}(\Lambda) is the standard coaction δG\delta_{G} on C(G)C^{*}(G) given in Example 2.24.

Now for normality, by Lemma 2.28, if GG is amenable, then all of its coactions are amenable, and in particular δG\delta_{G} is amenable. Conversely, if the gauge coaction on Ctight(Λ)C(G)C^{*}_{tight}(\Lambda)\cong C^{*}(G) is normal, then considering {Ug}g\{\mathbb{C}U_{g}\}_{g\in\mathbb{C}} as a topological grading over GG with conditional faithful expectation, by [9, Proposition 3.7], C(G)C^{*}(G) is naturally isomorphic to Cr({Ug}g)=Cr(G)C^{*}_{r}(\{\mathbb{C}U_{g}\}_{g\in\mathbb{C}})=C^{*}_{r}(G), so the full and reduced CC^{*}-algebras are isomorphic and thus GG is amenable.

Corollary 4.25.

If (G,P)=(F2,R)(G,P)=(F_{2},R) where RR a total ordering on GG such as the Magnus expansion given in Section 3.2 of [4], then Ctight(Λ)≇Cmin(Λ)C^{*}_{tight}(\Lambda)\not\cong C^{*}_{min}(\Lambda). Since both Ctight(Λ)C^{*}_{tight}(\Lambda) and Cmin(Λ)C^{*}_{min}(\Lambda) are Λ\Lambda-faithful, tight, gauge coacting representations, there is no gauge-invariant uniqueness theorem for this RR-graph.

Proof.

By the previous lemma, Ctight(Λ)C(G)C^{*}_{tight}(\Lambda)\cong C^{*}(G), and by Proposition 4.23 Ctight(Λ)C^{*}_{tight}(\Lambda) is a tight, Λ\Lambda-faithful representation of Λ\Lambda with a gauge coaction. Since G=F2G=F_{2} is not amenable, the gauge coaction is not normal. But by part (5) of 4.23, since the gauge coaction is not normal, then πST\pi^{T}_{S} is not an isomorphism, so Cmin(Λ)C^{*}_{min}(\Lambda) is another tight, Λ\Lambda-faithful representation of Λ\Lambda with a gauge coaction, but which is not canonically isomorphic to Ctight(Λ)C^{*}_{tight}(\Lambda).

The reader may recognize the ordered group (F2,R)(F_{2},R) from Example 3.6 earlier in this manuscript, where we used it as an example of an ordering that had no reduction to an amenable group. This recognition may ignite a spark of hope: perhaps the existence of a reduction to an amenable group prevents the obstruction seen in the previous lemma and guarantees that δ\delta is normal. We shall see in the next section that this is indeed the case.

4.3 Reductions and Representations

In this section, we develop the notion that if (G,P)(G,P) reduces to (H,Q)(H,Q), then PP-graphs are QQ-graphs and they have the same representations.

Remark 4.26.

In this section, we will be considering a category Λ\Lambda as a graph with respect to more than one semigroup (ex: as a PP-graph and as a QQ-graph). In cases where there may be confusion, we will write Cmin(Λ,P)C^{*}_{min}(\Lambda,P) and Ctight(Λ,P)C^{*}_{tight}(\Lambda,P) to indicate the semigroup PP, and so on for other semigroups.

First we will show that if Λ\Lambda is a PP-graph, and (G,P)(G,P) has reduction (H,Q)(H,Q), then Λ\Lambda is also a QQ-graph in a natural way.

Lemma 4.27.

Let (G,P)(G,P) be an ordered group, and let (H,Q)(H,Q) be a reduction of (G,P)(G,P) with degree functor dQP:PQd^{P}_{Q}:P\rightarrow Q. Then a PP-graph Λ\Lambda with degree functor dPΛd_{P}^{\Lambda} is also a QQ-graph with degree functor d=dQPdPΛd=d^{P}_{Q}\circ d_{P}^{\Lambda}.

Proof.

Certainly, dd maps Λ\Lambda into QQ, so, it suffices to check that Λ\Lambda has unique factorization with respect to dd.

Let λΛ\lambda\in\Lambda, and qQq\in Q satisfy qd(λ)q\leq d(\lambda). We must show that there is a unique α,βΛ\alpha,\beta\in\Lambda such that λ=αβ\lambda=\alpha\beta and d(α)=qd(\alpha)=q.

For existence, since dPΛ(λ)Pd_{P}^{\Lambda}(\lambda)\in P is a path of length dQP(dPΛ(λ))=d(λ)qd^{P}_{Q}(d_{P}^{\Lambda}(\lambda))=d(\lambda)\geq q, then by unique factorization of dQPd^{P}_{Q}, there exist unique p,p1Pp,p_{1}\in P such that pp1=dPΛ(λ)pp_{1}=d_{P}^{\Lambda}(\lambda) and dQP(p)=qd^{P}_{Q}(p)=q. Then pdPΛ(λ)p\leq d_{P}^{\Lambda}(\lambda), so by unique factorization of dPΛd_{P}^{\Lambda}, there exist unique α,βΛ\alpha,\beta\in\Lambda such that αβ=pp1\alpha\beta=pp_{1} and dPΛ(α)=pd_{P}^{\Lambda}(\alpha)=p. Then d(α)=dQP(dPΛ(α))=dQP(p)=qd(\alpha)=d^{P}_{Q}(d_{P}^{\Lambda}(\alpha))=d^{P}_{Q}(p)=q, so the desired α,β\alpha,\beta exist.

For uniqueness, if α,βΛ\alpha^{\prime},\beta^{\prime}\in\Lambda were two other paths with λ=αβ\lambda=\alpha^{\prime}\beta^{\prime} and d(α)=qd(\alpha^{\prime})=q, then dPΛ(λ)=dPΛ(α)dPΛ(β)d_{P}^{\Lambda}(\lambda)=d_{P}^{\Lambda}(\alpha^{\prime})d_{P}^{\Lambda}(\beta^{\prime}) would be a factorization of dPΛ(λ)d_{P}^{\Lambda}(\lambda) where the first term has degree dQP(dPΛ(α))=d(α)=qd^{P}_{Q}(d_{P}^{\Lambda}(\alpha^{\prime}))=d(\alpha^{\prime})=q, so by uniqueness of the factorization of dQPd^{P}_{Q}, we have that dPΛ(α)=pd_{P}^{\Lambda}(\alpha^{\prime})=p. Then by the uniqueness of the factorization dPΛd_{P}^{\Lambda}, we have α=α\alpha^{\prime}=\alpha, as desired.

Next, we will show that whether Λ\Lambda is regarded as a PP-graph or a QQ-graph, it has the same representations.

Lemma 4.28.

Let (G,P)(G,P) and (H,Q)(H,Q) be two WQLO groups, and Λ\Lambda a small category with two functors dPΛ:ΛPd^{\Lambda}_{P}:\Lambda\rightarrow P and dQΛ:ΛQd^{\Lambda}_{Q}:\Lambda\rightarrow Q such that Λ\Lambda is a PP-graph with respect to dPΛd^{\Lambda}_{P} and a QQ-graph with respect to dQΛd^{\Lambda}_{Q}. Then,

  1. 1.

    Λ\Lambda is finitely-aligned as a PP-graph if and only if it is finitely-aligned as a QQ-graph.

  2. 2.

    The (T1)-(T4) relators are the same for Λ\Lambda independently of being a PP-graph or a QQ-graph.

  3. 3.

    A representation is tight independently of Λ\Lambda being a PP-graph or a QQ-graph.

Proof.

For (1), recall that MCE(μ,ν)MCE(\mu,\nu) depends only on the category structure of Λ\Lambda by definition, so it is finite regardless of the degree functor given to Λ\Lambda.

Similarly for (2) and (3), the T1-T4 and tightness relations as written only depend on the category structure of Λ\Lambda.

Corollary 4.29.

Let (G,P)(G,P) and (H,Q)(H,Q) be two WQLO groups, and Λ\Lambda a small category with two functors dPΛ:ΛPd^{\Lambda}_{P}:\Lambda\rightarrow P and dQΛ:ΛQd^{\Lambda}_{Q}:\Lambda\rightarrow Q such that Λ\Lambda is a finitely-aligned PP-graph with respect to dPΛd^{\Lambda}_{P} and a finitely-aligned QQ-graph with respect to dQΛd^{\Lambda}_{Q}. Then Ctight(Λ,P)C^{*}_{tight}(\Lambda,P) is canonically isomorphic to Ctight(Λ,Q)C^{*}_{tight}(\Lambda,Q).

Proof.

Let aa and bb be the universal tight representations of Λ\Lambda as a PP-graph and QQ-graph, respectively, so that Ctight(Λ,P)=C(a)C^{*}_{tight}(\Lambda,P)=C^{*}(a) and Ctight(Λ,Q)=C(b)C^{*}_{tight}(\Lambda,Q)=C^{*}(b).

By the previous Lemma, both the aλa_{\lambda}s and the bλb_{\lambda}s satisfy the equivalent T1-T4 and tightness relators, so by their respective universal properties, there are canonical covers taking aλbλa_{\lambda}\mapsto b_{\lambda} and bλaλb_{\lambda}\mapsto a_{\lambda}, which are inverses, and hence canonical isomorphisms. ∎

Lemma 4.30.

Let φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) be a reduction of WQLO groups. Let Λ\Lambda be a finitely-aligned PP-graph, and tt a representation of Λ\Lambda. If tt has a gauge coaction by GG, then it has a gauge coaction by HH (when Λ\Lambda inherits the structure of a QQ-graph as described in 4.27).

In particular, if δ\delta is the gauge coaction by GG, then ϵ:=(idC(t)φ¯)δ\epsilon:=(\operatorname{id}_{C^{*}(t)}\otimes\bar{\varphi})\circ\delta is the gauge coaction by HH, where φ¯:C(G)C(H)\bar{\varphi}:C^{*}(G)\rightarrow C^{*}(H) is given by UgVφ(g)U_{g}\mapsto V_{\varphi(g)} .

Proof.

First let us fix some notation. Let {Ug}gG\{U_{g}\}_{g\in G} and {Vh}hH\{V_{h}\}_{h\in H} denote the universal unitary representations of GG and HH respectively, so that C(G)C^{*}(G) and C(H)C^{*}(H) are the closed span of the UgU_{g} and VhV_{h}, respectively. Let δ\delta denote the gauge coaction on C(t)C^{*}(t), dPΛd^{\Lambda}_{P} the degree function on Λ\Lambda as a PP-graph, and recall that the degree functor on Λ\Lambda as a QQ-graph was dQΛ=φdPΛd^{\Lambda}_{Q}=\varphi\circ d^{\Lambda}_{P}.

Then we have that gVφ(g)g\mapsto V_{\varphi(g)} is a unitary representation of GG, so by the universal property of C(G)C^{*}(G), there is a *-homomorphism φ¯:C(G)C(H)\bar{\varphi}:C^{*}(G)\rightarrow C^{*}(H) given by UgVφ(g)U_{g}\mapsto V_{\varphi(g)}. We will prove that there is a gauge coaction of HH on C(t)C^{*}(t) given by ϵ:=(idC(t)φ¯)δ\epsilon:=(\operatorname{id}_{C^{*}(t)}\otimes\bar{\varphi})\circ\delta. First note that by Lemma 2.14, there is such a function ϵ\epsilon, it is a *-homomorphism, and it maps C(t)C(t)C(H)C^{*}(t)\rightarrow C^{*}(t)\otimes C^{*}(H).

But now observe that

ϵ(tλ)\displaystyle\epsilon(t_{\lambda}) =\displaystyle= (idAφ¯)δ(tλ)\displaystyle(\operatorname{id}_{A}\otimes\bar{\varphi})\circ\delta(t_{\lambda})
=\displaystyle= (idAφ¯)(tλUdPΛ(λ))\displaystyle(\operatorname{id}_{A}\otimes\bar{\varphi})(t_{\lambda}\otimes U_{d_{P}^{\Lambda}(\lambda)})
=\displaystyle= tλVφ(dPΛ(λ))\displaystyle t_{\lambda}\otimes V_{\varphi(d_{P}^{\Lambda}(\lambda))}
=\displaystyle= tλVdQΛ(λ)\displaystyle t_{\lambda}\otimes V_{d^{\Lambda}_{Q}(\lambda)}

so by Proposition 2.50(3), ϵ\epsilon is a gauge coaction by HH, as desired. ∎

4.4 Gauge-Invariant Uniqueness for Ordered Groups which Reduce to Amenable Groups

The following is the mathematical core of our gauge-invariant uniqueness theorem:

Proposition 4.31.

Let (G,P)(G,P) be a WQLO group and suppose there is a reduction to an amenable group φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q). Then for any finitely-aligned PP-graph Λ\Lambda, the following are canonically isomorphic:

  1. a.

    Ctight(Λ,P)C^{*}_{tight}(\Lambda,P).

  2. b.

    Ctight(Λ,Q)C^{*}_{tight}(\Lambda,Q).

  3. c.

    Cmin(Λ,P)C^{*}_{min}(\Lambda,P).

  4. d.

    Cmin(Λ,Q)C^{*}_{min}(\Lambda,Q).

  5. e.

    C(t)C^{*}(t), where tt is any representation of Λ\Lambda as a PP-graph which is tight, Λ\Lambda-faithful, and which has a gauge coaction by GG.

Proof.

For the sake of abbreviation, let us denote the representations generating Ctight(Λ,P)C^{*}_{tight}(\Lambda,P), Ctight(Λ,Q)C^{*}_{tight}(\Lambda,Q), Cmin(Λ,P),C^{*}_{min}(\Lambda,P), and Cmin(Λ,Q)C^{*}_{min}(\Lambda,Q) as A,A, B,B, C,C, and DD, respectively. Our argument is organized by the following diagram:

C(A)=Ctight(Λ,P){C^{*}(A)=C^{*}_{tight}(\Lambda,P)}C(B)=Ctight(Λ,Q){C^{*}(B)=C^{*}_{tight}(\Lambda,Q)}C(C)=Cmin(Λ,P){C^{*}(C)=C^{*}_{min}(\Lambda,P)}C(D)=Cmin(Λ,Q){C^{*}(D)=C^{*}_{min}(\Lambda,Q)}πBA\scriptstyle{\pi_{B}^{A}}πCA\scriptstyle{\pi_{C}^{A}}πDB\scriptstyle{\pi_{D}^{B}}πDC\scriptstyle{\pi^{C}_{D}}

All arrows here denote canonical coverings.

The map πBA\pi_{B}^{A} exists and is an isomorphism by Corollary 4.29.

The maps πCA\pi_{C}^{A} and πDB\pi_{D}^{B} are the canonical coverings of Proposition 4.23(2). Since HH is amenable, then by Lemma 2.28 the gauge coaction by HH on Ctight(Λ,Q)C^{*}_{tight}(\Lambda,Q) is normal, and so πDB\pi_{D}^{B} is a canonical isomorphism by Proposition 4.23 (5).

Combining these two, we have that ABDA\cong B\cong D.

The map πDC\pi_{D}^{C} arises from the work established in the previous lemmas: by Lemma 4.30 C(C)C^{*}(C) carries a gauge coaction by HH. Since also CC is Λ\Lambda-faithful by 4.22, then by the minimality of Cmin(Λ,Q)C^{*}_{min}(\Lambda,Q), there is a canonical covering πDC:Cmin(Λ,P)Cmin(Λ,Q)\pi_{D}^{C}:C^{*}_{min}(\Lambda,P)\rightarrow C^{*}_{min}(\Lambda,Q).

Now, with the order notation of Lemma 2.48, we have that ArepCrepDBAA\geq_{rep}C\geq_{rep}D\cong B\cong A, so ABCDA\cong B\cong C\cong D. That is, (a.)-(d.) are canonically isomorphic.

For (e.), we have that AreptrepCAA\geq_{rep}t\geq_{rep}C\cong A by the respective universality and co-universality of C(A)C^{*}(A) and C(C)C^{*}(C), so AtCA\cong t\cong C. ∎

One nice outcome is that this result is independent of the amenable group (H,Q)(H,Q) and the reduction φ\varphi, and simply requires that one exist. In particular, we have a gauge-invariant uniqueness theorem for WQLO groups which have a reduction to an amenable group:

\uniquenesstheorem
Proof.

Let TT be the universal tight representation as in Proposition 4.23 (1), so C(T)Ctight(Λ)C^{*}(T)\cong C^{*}_{tight}(\Lambda). By that theorem, TT is Λ\Lambda-faithful, tight, and has a gauge coaction. It is also universal for tight representations.

By Proposition 4.31, C(T)Cmin(Λ)C^{*}(T)\cong C^{*}_{min}(\Lambda), so TT is co-universal for representations which are Λ\Lambda-faithful and have gauge coaction.

To see that TT is the unique representation (up to canonical isomorphism) which is Λ\Lambda-faithful, tight, and has a gauge coaction, if there were such another representation tt, it would be covered by TT (since tt is tight, and TT is universal for tight representations) and it would cover TT (since tt is Λ\Lambda-faithful and has a gauge coaction and TT is co-universal for such representations). Thus tTt\cong T.

Remark 4.32.

In the case that (G,P)(G,P) reduces to an amenable group, we believe that the CC^{*}-algebra generated by this unique Λ\Lambda-faithful, tight, gauge coacting representation deserves the title of the CC^{*}-algebra of the graph. We will write C(Λ)C^{*}(\Lambda) for this algebra.

4.5 Additional Results in the case of a Strong Reduction

In this section we will consider what further results can be provided about a PP-graph algebra under the assumption that (G,P)(G,P) strongly reduces to an amenable group.

Lemma 4.33.

Let (G,P)(G,P) be a WQLO group, and suppose (G,P)(G,P) has a strong reduction to an amenable group. Then every gauge coaction on a finitely-aligned PP-graph is normal.

Proof.

Let φ:(G,P)(H,Q)\varphi:(G,P)\rightarrow(H,Q) be the strong reduction where HH is amenable. Fix some finitely-aligned PP-graph Λ\Lambda and a representation ss of Λ\Lambda which has a gauge coaction by GG.

Then by Lemma 4.30, ss has a gauge coaction by HH. For sake of clarity, let Φs\Phi_{s} and Ωs\Omega_{s} denote the conditional expectations on C(s)C^{*}(s) as in Lemma 2.26 which respectively arise from its GG-coaction and HH-coaction.

By Lemma 2.28, since HH is amenable, then Ωs\Omega_{s} is faithful. Our desired result is that Φs\Phi_{s} is faithful.

To this end, it suffices to show that Ωs=Φs\Omega_{s}=\Phi_{s} as maps on C(s)C^{*}(s), and since C(s)=span¯{sαsβ:α,βΛ}C^{*}(s)=\overline{\operatorname{span}}\{s_{\alpha}s_{\beta}^{*}:\alpha,\beta\in\Lambda\} and both functions are continuous and linear, it suffices to check that Ωs=Φs\Omega_{s}=\Phi_{s} on a single term sαsβs_{\alpha}s_{\beta}^{*}.

Recall that by Lemma 2.26, we have

Φs(sαsβ)={sαsβ if d(α)=d(β)0 if d(α)d(β)\Phi_{s}(s_{\alpha}s_{\beta}^{*})=\begin{cases}s_{\alpha}s_{\beta}^{*}&\text{ if }d(\alpha)=d(\beta)\\ 0&\text{ if }d(\alpha)\neq d(\beta)\end{cases}

and

Ωs(sαsβ)={sαsβ if φ(d(α))=φ(d(β))0 if φ(d(α))φ(d(β)).\Omega_{s}(s_{\alpha}s_{\beta}^{*})=\begin{cases}s_{\alpha}s_{\beta}^{*}&\text{ if }\varphi(d(\alpha))=\varphi(d(\beta))\\ 0&\text{ if }\varphi(d(\alpha))\neq\varphi(d(\beta))\end{cases}.

But φ\varphi is a strong reduction, so for the positive elements d(α),d(β)Pd(\alpha),d(\beta)\in P, we have d(α)=d(β)d(\alpha)=d(\beta) if and only if φ(d(α))=φ(d(β))\varphi(d(\alpha))=\varphi(d(\beta)).

Thus Ωs=Φs\Omega_{s}=\Phi_{s} on each spanning element, so they are equal as functions, and thus Φs\Phi_{s} is faithful, so ss is normal.

Lemma 4.34.

Let (G,P)(G,P) be a WQLO group with a strong reduction onto an amenable group, let Λ\Lambda be a PP-graph, and let ss be a representations of (G,P)(G,P). If ss is Λ\Lambda-faithful, has a gauge coaction, and every proper bolt in ss is nonzero, then s𝒯s\cong\mathcal{T}.

Proof.

Since 𝒯\mathcal{T} covers ss, it suffices to show that kerπs𝒯=0\ker\pi^{\mathcal{T}}_{s}=0.

To this end, consider the diagram

C(𝒯){C^{*}(\mathcal{T})}(𝒯){\mathcal{B}(\mathcal{T})}C(s){C^{*}(s)}(s){\mathcal{B}(s)}Φ𝒯\scriptstyle{\Phi_{\mathcal{T}}}πs𝒯\scriptstyle{\pi^{\mathcal{T}}_{s}}ψs𝒯\scriptstyle{\psi^{\mathcal{T}}_{s}}Φs\scriptstyle{\Phi_{s}}

which commutes by a routine computation on the spanning elements 𝒯μ𝒯ν\mathcal{T}_{\mu}\mathcal{T}_{\nu}^{*}.

Fix some xkerπs𝒯=0x\in\ker\pi^{\mathcal{T}}_{s}=0. Then πs𝒯(xx)=0\pi^{\mathcal{T}}_{s}(x^{*}x)=0, so

0=Φs(πs𝒯(xx))=ψs𝒯(Φ𝒯(xx)).0=\Phi_{s}(\pi^{\mathcal{T}}_{s}(x^{*}x))=\psi^{\mathcal{T}}_{s}(\Phi_{\mathcal{T}}(x^{*}x)).

Now by Theorem 4.15, since kerψs𝒯\ker\psi^{\mathcal{T}}_{s} is generated by the bolts and 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}s it contains, and by the hypotheses on ss this kernel contains no proper bolts or 𝒯μ𝒯μ\mathcal{T}_{\mu}\mathcal{T}_{\mu}^{*}, then we have kerψs𝒯=0\ker\psi^{\mathcal{T}}_{s}=0, and thus Φ𝒯(xx)=0\Phi_{\mathcal{T}}(x^{*}x)=0.

By Lemma 4.33, Φ𝒯\Phi_{\mathcal{T}} is faithful, so xx=0x^{*}x=0 and thus x=0x=0. Thus kerπs𝒯={0}\ker\pi_{s}^{\mathcal{T}}=\{0\}, as desired. ∎

Corollary 4.35.

Let Λ\Lambda be a finitely-aligned 2\mathbb{N}^{2}*\mathbb{N}-graph. Then any gauge coaction by G=2G=\mathbb{Z}^{2}*\mathbb{Z} is automatically normal.

Proof.

Recall that (2,2)(\mathbb{Z}^{2}*\mathbb{Z},\mathbb{N}^{2}*\mathbb{N}) strongly reduces to (2,2)(\mathbb{Z}^{2}\wr\mathbb{Z},\mathbb{N}^{2}\wr\mathbb{N}), which is amenable. Then by Lemma 4.33, the gauge coaction is normal.

Lemma 4.36.

Let (G,P)(G,P) be a WQLO group and suppose that (G,P)(G,P) has a strong reduction to an amenable group. Let Λ\Lambda be a finitely-aligned PP-graph, and let tt be a representation of Λ\Lambda. Then the following are equivalent:

  1. 1.

    tt has a gauge coaction by GG

  2. 2.

    There is a bounded linear map Φt:C(t)(t)\Phi_{t}:C^{*}(t)\rightarrow\mathcal{B}(t) satisfying

    Φt(tμtν)={tμtν if d(μ)=d(ν)0 if d(μ)d(ν)\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ if }d(\mu)\neq d(\nu)\end{cases}
Proof.

(12)(1\Rightarrow 2) If tt has a gauge coaction, then such a map Φt\Phi_{t} exists by Lemma 4.8.

(21)(2\Rightarrow 1) If such a Φt\Phi_{t} exists, then by Theorem 19.1 and Definition 19.2 of [10], writing Bg=span¯{tμtν:d(μ)d(ν)1=g}B_{g}=\overline{\operatorname{span}}\{t_{\mu}t_{\nu}^{*}:d(\mu)d(\nu)^{-1}=g\}, we have that ={Bg}gG\mathcal{B}=\{B_{g}\}_{g\in G} are linearly independent and form a topological grading of C(t)C^{*}(t).

In particular, \mathcal{B} is a Fell bundle, and we can form the cross-sectional algebra C()C^{*}(\mathcal{B}) as in Remark 2.21. Each graded component BgB_{g} is naturally embedded in C()C^{*}(\mathcal{B}), so for each λΛ\lambda\in\Lambda, let sλs_{\lambda} denote the operator arising from embedding tλBd(λ)t_{\lambda}\in B_{d(\lambda)} into C()C^{*}(\mathcal{B}).

But one may immediately check the T1-T4 relators to check that {sλ}λΛ\{s_{\lambda}\}_{\lambda\in\Lambda} is a representation of Λ\Lambda in C()C^{*}(\mathcal{B}). Note that since the {sλ}λΛ\{s_{\lambda}\}_{\lambda\in\Lambda} generate each fiber BgB_{g}, then they generate all of C()C^{*}(\mathcal{B}), so C()=C(s)C^{*}(\mathcal{B})=C^{*}(s).

Our goal is now to show that ss has a gauge coaction and that sts\cong t. For the former, by [17, Proposition 3.3], there is a coaction δ\delta on C()C^{*}(\mathcal{B}) satisfying

δ(b)=bUg for all bBg,gG.\delta(b)=b\otimes U_{g}\text{ for all }b\in B_{g},g\in G.

In particular, since sλBd(λ)s_{\lambda}\in B_{d(\lambda)}, then δ(sλ)=sλUd(λ)\delta(s_{\lambda})=s_{\lambda}\otimes U_{d(\lambda)}, so δ\delta is a gauge coaction on ss.

Now we must show that sts\cong t. By [10, Theorem 19.5], since C(t)C^{*}(t) is a topologically graded CC^{*}-algebra, with grading \mathcal{B} there is a commutative diagram of surjective *-homomorphisms

C()=C(s){C^{*}(\mathcal{B})=C^{*}(s)}Cr(){C^{*}_{r}(\mathcal{B})}C(t){C^{*}(t)}L\scriptstyle{L}πts\scriptstyle{\pi^{s}_{t}}ψ\scriptstyle{\psi}

where LL denotes the regular representation from [10, Definition 17.6] (in that definition, this map is called Λ\Lambda, but we have changed its name to avoid confusion with the PP-graph Λ\Lambda).

Recall that C()C^{*}(\mathcal{B}) is topologically graded with conditional expectation Φs\Phi_{s}. Since GG strongly reduces to an amenable group, then by Lemma 4.33, the gauge action ss is normal and thus the conditional expectation Φs\Phi_{s} is faithful. But by [10, Proposition 19.6], the kernel of the regular representation LL is given by

ker(L)={xC():Φs(xx)=0}\ker(L)=\{x\in C^{*}(\mathcal{B}):\Phi_{s}(x^{*}x)=0\}

and since Φs\Phi_{s} is faithful, then kerL={0}\ker L=\{0\}. Since LL is injective and L=ψπstL=\psi\circ\pi_{s}^{t}, then πst\pi_{s}^{t} is injective, so sts\cong t, so tt has a gauge coaction as desired.

The reader should note that if such a bounded linear map exists, then there is a gauge coaction, so that bounded linear map is the conditional expectation of Lemma 2.26.

The next result shows that if you represent a PP-graph on a Hilbert space in such a way that the Hilbert space also has a “PP-grading” which interacts with the grading on Λ\Lambda in a natural way, then this representation has a gauge coaction.

Lemma 4.37.

Let (G,P)(G,P) be a WQLO group and suppose that (G,P)(G,P) has a strong reduction to an amenable group. Let Λ\Lambda be a finitely-aligned PP graph. Let tt be a representation of Λ\Lambda in ()\mathcal{B}(\mathcal{H}) for some Hilbert space \mathcal{H}, and suppose that =pPp\mathcal{H}=\bigoplus_{p\in P}\mathcal{H}_{p} such that tμpd(μ)pt_{\mu}\mathcal{H}_{p}\subseteq\mathcal{H}_{d(\mu)p} for all μΛ\mu\in\Lambda, pPp\in P. Then tt has a gauge coaction.

Proof.

By Lemma 4.36, it suffices to show that there is a bounded linear map Φt:C(t)(t)\Phi_{t}:C^{*}(t)\rightarrow\mathcal{B}(t) satisfying

Φt(tμtν)={tμtν if d(μ)=d(ν)0 if d(μ)d(ν).\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ if }d(\mu)\neq d(\nu)\end{cases}.

For each pPp\in P, let Qp()Q_{p}\in\mathcal{B}(\mathcal{H}) denote the projection onto p\mathcal{H}_{p}. Then define

Φt(x)=pPWOT limitQpxQp\Phi_{t}(x)=\displaystyle\sum_{\begin{subarray}{c}p\in P\\ WOT\text{ limit}\end{subarray}}Q_{p}xQ_{p}

which is bounded and converges in the weak operator topology because the {Qp}pP\{Q_{p}\}_{p\in P} are pairwise orthogonal projections. We then wish to show that

Φt(tμtν)={tμtν if d(μ)=d(ν)0 if d(μ)d(ν)\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ if }d(\mu)\neq d(\nu)\end{cases}

for all μ,νΛ\mu,\nu\in\Lambda.

To this end, fix some qPq\in P, hqqh_{q}\in\mathcal{H}_{q}, and xx\in\mathcal{H}. Then we have that

Φt(tμtν)hq,x\displaystyle\langle\Phi_{t}(t_{\mu}t_{\nu}^{*})h_{q},x\rangle =\displaystyle= pPWOT limitQptμtνQphq,x\displaystyle\displaystyle\sum_{\begin{subarray}{c}p\in P\\ WOT\text{ limit}\end{subarray}}\langle Q_{p}t_{\mu}t_{\nu}^{*}Q_{p}h_{q},x\rangle
=\displaystyle= QqtμtνQqhq,x\displaystyle\langle Q_{q}t_{\mu}t_{\nu}^{*}Q_{q}h_{q},x\rangle
=\displaystyle= Qqtμtνhq,x\displaystyle\langle Q_{q}t_{\mu}t_{\nu}^{*}h_{q},x\rangle

since Qphq=0Q_{p}h_{q}=0 if pqp\neq q. Since our choice of xx\in\mathcal{H} was arbitrary, we may unbind the xx term, so we have shown that Φt(tμtν)hq=Qqtμtνhq\Phi_{t}(t_{\mu}t_{\nu}^{*})h_{q}=Q_{q}t_{\mu}t_{\nu}^{*}h_{q}.

But note that tμtνhqd(μ)d(ν)1qt_{\mu}t_{\nu}^{*}h_{q}\in\mathcal{H}_{d(\mu)d(\nu)^{-1}q} and q=d(μ)d(ν)1qq=d(\mu)d(\nu)^{-1}q if and only if d(μ)=d(ν)d(\mu)=d(\nu), so

Φt(tμtν)hq=Qqtμtνhq={tμtνhq if d(μ)=d(ν)0 if d(μ)d(ν).\Phi_{t}(t_{\mu}t_{\nu}^{*})h_{q}=Q_{q}t_{\mu}t_{\nu}^{*}h_{q}=\begin{cases}t_{\mu}t_{\nu}^{*}h_{q}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ if }d(\mu)\neq d(\nu)\end{cases}.

Since this is true for each qPq\in P and hqqh_{q}\in\mathcal{H}_{q}, and the q\mathcal{H}_{q} span all of \mathcal{H}, we may unbind the hqh_{q} to get that

Φt(tμtν)={tμtν if d(μ)=d(ν)0 if d(μ)d(ν)\Phi_{t}(t_{\mu}t_{\nu}^{*})=\begin{cases}t_{\mu}t_{\nu}^{*}&\text{ if }d(\mu)=d(\nu)\\ 0&\text{ if }d(\mu)\neq d(\nu)\end{cases}

as desired.

Corollary 4.38.

Let (G,P)(G,P) be a WQLO group and suppose that (G,P)(G,P) has a strong reduction to an amenable group. Let Λ\Lambda be a finitely-aligned PP-graph. Then L𝒯L\cong\mathcal{T}, where LL denotes the regular representation of Λ\Lambda and 𝒯\mathcal{T} denotes the Toeplitz representation of Λ\Lambda.

Proof.

To show that LL is universal, by Lemma 4.34 it suffices to show that LL it is Λ\Lambda-faithful, has a gauge coaction, and that every proper bolt is nonzero. By Remark 2.45, LL is Λ\Lambda-faithful and its proper bolts are nonzero.

To show that LL has a gauge coaction, we will appeal to Lemma 4.37. Recall that LL is a representation in (2(Λ))\mathcal{B}(\ell^{2}(\Lambda)) by

Lμeα={eμα if s(μ)=r(α)0 if s(μ)r(α).L_{\mu}e_{\alpha}=\begin{cases}e_{\mu\alpha}&\text{ if }s(\mu)=r(\alpha)\\ 0&\text{ if }s(\mu)\neq r(\alpha)\end{cases}.

Therefore, writing Λp={μΛ:d(μ)=p}\Lambda^{p}=\{\mu\in\Lambda:d(\mu)=p\}, we can see that 2(Λ)=pP2(Λp)\ell^{2}(\Lambda)=\bigoplus_{p\in P}\ell^{2}(\Lambda^{p}), and a simple calculation shows that Lμ2(Λp)2(Λd(μ)p)L_{\mu}\ell^{2}(\Lambda^{p})\subseteq\ell^{2}(\Lambda^{d(\mu)p}). Therefore, by Lemma 4.37, LL has a gauge coaction.

Thus by by Lemma 4.34, L𝒯L\cong\mathcal{T}.

The following result shows that if Λ\Lambda has no “infinite paths”, then the ultrafilter representation is the universal tight representation.

Lemma 4.39.

Let (G,P)(G,P) be a WQLO group and suppose that (G,P)(G,P) has a strong reduction to an amenable group. Let Λ\Lambda be a finitely-aligned PP-graph. Recall that Λ\Lambda has a partial order given by αβ\alpha\leq\beta if βαΛ\beta\in\alpha\Lambda. Suppose that with respect to this partial order, every chain in Λ\Lambda has an upper bound in Λ\Lambda. Then the ultrafilter representation ff of Λ\Lambda is the universal tight representation of Λ\Lambda.

Proof.

By our gauge-invariant uniqueness theorem for PP-graphs, it suffices to show that ff is Λ\Lambda-faithful, tight, and has a gauge coaction. The former two properties are shown in Lemma 4.5. It then suffices to use Lemma 4.37 to show that ff has a gauge coaction.

To this end, we wish to show that the ultrafilters in ff correspond to maximal paths in Λ\Lambda. More formally, for each path αΛ\alpha\in\Lambda, let Uα={μΛ:μα}U_{\alpha}=\{\mu\in\Lambda:\mu\leq\alpha\}. It is straightforward to check that UαU_{\alpha} is a filter, and that it is an ultrafilter if and only if α\alpha is maximal. We wish to show that all ultrafilters in Λ\Lambda are of the form UαU_{\alpha}.

Fix some ultrafilter UU. If UU is finite, it must contain a maximal element α\alpha, in which case U=UαU=U_{\alpha}. If instead UU is infinite, since UΛU\subseteq\Lambda is countable, we may enumerate U={αn}nU=\{\alpha_{n}\}_{n\in\mathbb{N}}. Inductively define a sequence {βn}nU\{\beta_{n}\}_{n\in\mathbb{N}}\subseteq U by β1=α1\beta_{1}=\alpha_{1}, and for n>1n>1, let βn\beta_{n} be a common upper bound of βn1\beta_{n-1} and αn\alpha_{n} which is guaranteed to exist by the F2F2 property of filters. Note that U=n=1UβnU=\bigcup_{n=1}^{\infty}U_{\beta_{n}}.

Then {βn}n\{\beta_{n}\}_{n\in\mathbb{N}} is a chain in Λ\Lambda, so by our hypothesis, it has an upper bound βΛ\beta\in\Lambda. Then UβUβnU_{\beta}\supseteq U_{\beta_{n}} for all nn, and since U=n=1UβnU=\bigcup_{n=1}^{\infty}U_{\beta_{n}}, then UβUU_{\beta}\supseteq U, and since UU is an ultrafilter, U=UβU=U_{\beta}, as desired.

Therefore, we can partition the set of ultrafilters by the degree of their unique maximal element. That is, writing Λ^p={UαΛ^:d(α)=p}\widehat{\Lambda}_{\infty}^{p}=\{U_{\alpha}\in\widehat{\Lambda}_{\infty}:d(\alpha)=p\}, we have Λ^=pPΛ^p\widehat{\Lambda}_{\infty}=\bigsqcup_{p\in P}\widehat{\Lambda}_{\infty}^{p}, and therefore

2(Λ^)=pP2(Λ^p).\ell^{2}(\widehat{\Lambda}_{\infty})=\bigoplus_{p\in P}\ell^{2}(\widehat{\Lambda}_{\infty}^{p}).

Finally, a simple calculation shows that fμ2(Λ^p)2(Λ^d(μ)p)f_{\mu}\ell^{2}(\widehat{\Lambda}_{\infty}^{p})\subseteq\ell^{2}(\widehat{\Lambda}_{\infty}^{d(\mu)p}), so by Lemma 4.37, ff has a gauge coaction, as desired.

Example 4.40.

Let G=BS(1,2)=c,t|tc=c2tG=BS(1,2)=\langle c,t|tc=c^{2}t\rangle, and let P={c,t}P=\{c,t\}^{*}. Note that BS(1,2)BS(1,2) is amenable, and therefore has a strong reduction to an amenable group (itself). Define a PP-graph Λ\Lambda as in this diagram:

v1{v_{1}}v2{v_{2}}v3{v_{3}}v4{v_{4}}v5{v_{5}}t1\scriptstyle{t_{1}}c1\scriptstyle{c_{1}}t2\scriptstyle{t_{2}}c2\scriptstyle{c_{2}}c3\scriptstyle{c_{3}}

so Λ={v1,v2,v3,v4,v5,c1,c2,c3,t1,t2,c2t1,c3c2,c3c2t1=c1t2}\Lambda=\{v_{1},v_{2},v_{3},v_{4},v_{5},c_{1},c_{2},c_{3},t_{1},t_{2},c_{2}t_{1},c_{3}c_{2},c_{3}c_{2}t_{1}=c_{1}t_{2}\}, where the range and source maps are as indicated in the diagram, the operation is concatenation of paths, and the degrees are given by d(vi)=1d(v_{i})=1, d(ci)=cd(c_{i})=c, d(ti)=td(t_{i})=t, and so on.

Since Λ\Lambda is finite, all chains have an upper bound, so by Lemma 4.39, the ultrafilter representation is the universal tight representation and the set of ultrafilters is in bijection with the maximal paths in Λ\Lambda. Thus Λ^={v1,c1,t1,c2t1,c3c2t1=t2c1}\widehat{\Lambda}_{\infty}=\{v_{1},c_{1},t_{1},c_{2}t_{1},c_{3}c_{2}t_{1}=t_{2}c_{1}\}, so C(f)C^{*}(f) is represented on a 5-dimension vector space and C(f)M5C^{*}(f)\subseteq M_{5}, the space of 5-by-5 matrices. A direct calculation with respect to the basis {ev1,ec1,et1,ec2t1,ec3c2t1=et2c1}\{e_{v_{1}},e_{c_{1}},e_{t_{1}},e_{c_{2}t_{1}},e_{c_{3}c_{2}t_{1}}=e_{t_{2}c_{1}}\} shows that

fv1\displaystyle f_{v_{1}} =\displaystyle= E11\displaystyle E_{11}
fc1\displaystyle f_{c_{1}} =\displaystyle= E21\displaystyle E_{21}
ft1\displaystyle f_{t_{1}} =\displaystyle= E31\displaystyle E_{31}
fc2t1\displaystyle f_{c_{2}t_{1}} =\displaystyle= E41\displaystyle E_{41}
ft2c1\displaystyle f_{t_{2}c_{1}} =\displaystyle= E51\displaystyle E_{51}

where EijE_{ij} denotes the matrix consisting of all 0s except for a single 1 in the iith row and jjth column (a matrix unit). Since these matrix units generate M5M_{5}, then C(f)=C(Λ)=M5C^{*}(f)=C^{*}(\Lambda)=M_{5}.

It is notable that the results in this section apply only to groups with strong reductions, instead of any reduction. The only “bottleneck” for this is Lemma 4.33, which says that gauge coactions are normal when (G,P)(G,P) strongly reduces to an amenable group. We conjecture that a (non-strong) reduction to an amenable group would suffice:

Conjecture 4.41.

Let (G,P)(G,P) be a WQLO group, let Λ\Lambda be a finitely aligned PP-graph, and suppose (G,P)(G,P) reduces to an amenable group. Then every gauge coaction on a representation of Λ\Lambda is normal.

Therefore, the other results in this section only require the hypothesis that (G,P)(G,P) reduces to an amenable group.

4.6 Applications to Kirchberg Algebras in the UCT Class

Definition 4.42.

A CC^{*}-algebra AA is called a Kirchberg algebra if it is purely infinite, simple, nuclear, and separable.

A CC^{*}-algebra is said to be in the UCT class if it satisfies the hypotheses of the universal coefficient theorem. (For a more thorough treatment, the reader is directed to the introduction of [1]).

Proposition 4.43.

Let AA be a Kirchberg algebra in the UCT class. Then there is a finitely aligned 2\mathbb{N}^{2}*\mathbb{N}-graph such that AA is stably isomorphic to C(Λ)C^{*}(\Lambda).

Proof.

By [3, Corollary 6.3], there is a finitely aligned 2\mathbb{N}^{2}*\mathbb{N}-graph Λ\Lambda such that AA is stably isomorphic to Cmin(Λ)C^{*}_{min}(\Lambda). But since 2\mathbb{N}^{2}*\mathbb{N} strongly reduces to an amenable group by Corollary 3.30, then Cmin(Λ)=C(Λ)C^{*}_{min}(\Lambda)=C^{*}(\Lambda) by the gauge-invariant uniqueness theorem for PP-graphs (Theorem 1). Thus AA is stably isomorphic to C(Λ)C^{*}(\Lambda). ∎

5 Appendix 1: Tight Representations

In this section, our goal is to justify our use of the term “tight” in the context of representations of PP-graphs by showing that a representation is tight in the sense of Definition 2.44(c) if and only if it corresponds to a tight representation of a semilattice in the sense of Exel’s definition in [8].

In this section, we fix (G,P)(G,P) a WQLO group, Λ\Lambda a finitely-aligned PP-graph, and t:ΛC(t)t:\Lambda\rightarrow C^{*}(t) a representation. We will often write qλ=tλtλq_{\lambda}=t_{\lambda}t_{\lambda}^{*}.

It can be easy to be confused about what symbols denote paths, sets of paths, and sets of sets of paths. To avoid confusion, we will try to stick to the convention that greek letters like α\alpha denote paths in Λ\Lambda, lower case letters (and μΛ\mu\Lambda where μΛ\mu\in\Lambda) denote subsets of Λ\Lambda, and capital letters denote collections of subsets.

5.1 Background

Recall the following definition of the tightness from 2.44(c), slightly modified to keep with our notation for this section:

Definition 5.1.

A representation tt of a graph Λ\Lambda is tight if for every μΛ\mu\in\Lambda and finite zμΛz\subseteq\mu\Lambda which is exhaustive for μΛ\mu\Lambda, we have that

αz(qμqα)=0\displaystyle\prod_{\alpha\in z}(q_{\mu}-q_{\alpha})=0

where qλ=tλtλq_{\lambda}=t_{\lambda}t_{\lambda}^{*}.

This condition was referred to a tightness before, but in this section will be referred to as H-tightness to avoid confusion.

The following terminology is from [8, Section 2]:

Definition 5.2.

In [8]’s context, a partially ordered set contains a smallest element, denoted 0. A semilattice is a partially ordered set where for all x,yXx,y\in X, the set {zX:zx,y}\{z\in X:z\leq x,y\} has a maximum element, denoted xyx\wedge y. That is, xyx\wedge y is the infimum of xx and yy, and every two elements have an infimum (although it may be 0).

Given, x,yXx,y\in X, we say xx and yy are disjoint and write xyx\perp y if there is no nonzero zXz\in X with zx,yz\leq x,y. Otherwise we say that xx and yy intersect, and write xyx\Cap y.

Definition 5.3.

Let EE be a semilattice. Given finite X,YEX,Y\subseteq E, let EX,YE^{X,Y} denote the subset of EE given by

EX,Y={zE:zxxX,zyyY}.E^{X,Y}=\{z\in E:z\leq x\ \forall x\in X,z\perp y\ \forall y\in Y\}.

Given any subset FEF\subseteq E, we shall say that a subset ZFZ\subseteq F is a cover for FF if, for every nonzero xFx\in F there exists zZz\in Z such xx and zz have a common extension.

Definition 5.4.

Let σ:E\sigma:E\rightarrow\mathcal{B} be a representation of a semilattice EE in a Boolean algebra \mathcal{B}. We shall say that σ\sigma is tight if for every finite subsets X,YEX,Y\subseteq E and for every finite cover ZZ for EX,YE^{X,Y}, one has that

zZσ(z)xXσ(x)yY¬σ(y).\bigvee_{z\in Z}\sigma(z)\geq\bigwedge_{x\in X}\sigma(x)\wedge\bigwedge_{y\in Y}\neg\sigma(y).

Since the reverse inequality is always true, tightness is equivalent to equality in this expression.

To avoid confusion with H-tightness, we will refer to this notion of tightness as E-tightness.

5.2 Building our Semilattice

We’ll now define our semilattice EE and develop its basic properties.

Remark 5.5.

Given Λ\Lambda, let E(Λ)={i=1nαiΛ:αiΛ}E(\Lambda)=\{\bigcup_{i=1}^{n}\alpha_{i}\Lambda:\alpha_{i}\in\Lambda\} denote the collection of finite unions of αΛ\alpha\Lambdas, ordered under containment. When Λ\Lambda is clear from context, we will write EE for E(Λ)E(\Lambda). The empty set serves as the zero element for EE, and two elements a,ba,b of EE have a least lower bound aba\cap b, which is also in EE by finite alignment of Λ\Lambda. Thus EE is a semilattice.

Recall that Λ\Lambda is given a partial ordering \leq by αβ\alpha\leq\beta if and only if β=αα1\beta=\alpha\alpha_{1}. Note that αβ\alpha\leq\beta if and only if βΛαΛ\beta\Lambda\subseteq\alpha\Lambda.

For aEa\in E, we let m(a)={αa:α minimal in a}m(a)=\{\alpha\in a:\alpha\text{ minimal in }a\} denote the set of maximal elements of aa (with respect to the ordering on Λ\Lambda). For any aEa\in E, we may write a=i=1nαiΛa=\bigcup_{i=1}^{n}\alpha_{i}\Lambda for some αiΛ\alpha_{i}\in\Lambda, so it follows that m(a){αi}i=1nm(a)\subseteq\{\alpha_{i}\}_{i=1}^{n}, so m(a)m(a) is finite. Note also that m(a)m(a) is empty if and only if aa is empty. Finally, note that a=αm(a)αΛa=\bigcup_{\alpha\in m(a)}\alpha\Lambda. So in this way, mm provides a unique representation of each element of EE, and provides a bijection between EE and the set of finite, pairwise incomparable subsets of Λ\Lambda.

If tt be a representation of Λ\Lambda, let qα=tαtαq_{\alpha}=t_{\alpha}t_{\alpha}^{*} for all αΛ\alpha\in\Lambda, and let tC(t)\mathcal{B}_{t}\subseteq C^{*}(t) be the Boolean algebra generated by 11 and {qα:αΛ}\{q_{\alpha}:\alpha\in\Lambda\}, where qαqβ=qαqβq_{\alpha}\wedge q_{\beta}=q_{\alpha}q_{\beta}.

Finally, given such a representation tt, define σt:Et\sigma_{t}:E\rightarrow\mathcal{B}_{t} by

σt:a=αm(a)αΛαm(a)tαtα=αm(a)qα.\sigma_{t}:a=\bigcup_{\alpha\in m(a)}\alpha\Lambda\mapsto\bigvee_{\alpha\in m(a)}t_{\alpha}t_{\alpha}^{*}=\bigvee_{\alpha\in m(a)}q_{\alpha}.

When the representation tt is clear from context, we will write σ\sigma and \mathcal{B} for σt\sigma_{t} and t\mathcal{B}_{t}, respectively.

5.3 Technical Lemmas

Lemma 5.6 (Intersection Criterion).

If a,bEa,b\in E, then aba\Cap b if and only if there are αm(a),βm(b)\alpha\in m(a),\beta\in m(b) such that αβ\alpha\Cap\beta (in the sense of having a common extension in Λ\Lambda).

Proof.

Observe that ab=αm(a),βm(b)αΛβΛa\cap b=\bigcup_{\alpha\in m(a),\beta\in m(b)}\alpha\Lambda\cap\beta\Lambda, so the lefthand side is nonempty if and only if one of the terms on the righthand side are nonempty. But αΛβΛ\alpha\Lambda\cap\beta\Lambda is nonempty if and only if α\alpha has a common extension with β\beta. ∎

Lemma 5.7.

Let ZZ be a finite cover of EX,YE^{X,Y}. Then there is another finite cover ZZ^{\prime} of EX,YE^{X,Y} such that every element of ZZ^{\prime} is of the form αΛ\alpha\Lambda, and zZz=zZz\bigvee_{z\in Z}z=\bigvee_{z^{\prime}\in Z^{\prime}}z^{\prime}.

Proof.

Let B=m(zZz)B=m(\bigcup_{z\in Z}z) and let Z={βΛ:βB}Z^{\prime}=\{\beta\Lambda:\beta\in B\}. Then by construction BB and hence ZZ^{\prime} are finite sets, and by our initial remark on mm, we know that zZz=zZz\bigcup_{z\in Z}z=\bigvee_{z\in Z}z can be reconstituted from its minimal elements in the sense that zZz=zZz\bigvee_{z\in Z}z=\bigvee_{z^{\prime}\in Z^{\prime}}z^{\prime}.

It then suffices to show that ZZ^{\prime} is a cover of EX,YE^{X,Y}. To first show that it is contained in EX,YE^{X,Y}, we know that if βΛZ\beta\Lambda\in Z^{\prime}, then βB\beta\in B, so β\beta must be minimal in some z0Zz_{0}\in Z. Then for all xXx\in X and yYy\in Y, βΛz0x\beta\Lambda\leq z_{0}\leq x, and βΛyz0y=0\beta\Lambda\wedge y\leq z_{0}\wedge y=0, so βΛx\beta\Lambda\leq x and βΛy\beta\Lambda\perp y for all xX,yYx\in X,y\in Y. Thus ZEX,YZ^{\prime}\subseteq E^{X,Y}.

To show it covers, fix some nonzero wEX,Yw\in E^{X,Y}. Then since ZZ covers EX,YE^{X,Y}, there is a zZz\in Z such that zwz\Cap w. Then by the intersection criterion, there is a μm(z),νm(w)\mu\in m(z),\nu\in m(w) such that μν\mu\Cap\nu. Since μm(z)\mu\in m(z), then μΛβBβΛ\mu\Lambda\leq\bigcup_{\beta\in B}\beta\Lambda, so there is some βB\beta\in B with βμ\beta\leq\mu. Then βν\beta\Cap\nu, so βΛw\beta\Lambda\Cap w. Thus ZZ^{\prime} is a finite cover.

Lemma 5.8.

Let X,YEX,Y\subseteq E be finite sets. Then there is a finite set YEY^{\prime}\subseteq E such that EX,Y=EX,YE^{X,Y}=E^{X,Y^{\prime}} and YEX,Y^{\prime}\subseteq E^{X,\emptyset}.

Proof.

By the remark in [8] following Definition 2.5, we may assume without loss of generality, that X={x}X=\{x\}, a singleton. Let Y={yx:xX}Y^{\prime}=\{y\wedge x:x\in X\} which is finite. Then by construction, y=yxxy^{\prime}=y\wedge x\leq x for all yYy^{\prime}\in Y^{\prime}, so YEX,Y^{\prime}\subseteq E^{X,\emptyset}. It then suffices to show that EX,Y=EX,YE^{X,Y}=E^{X,Y^{\prime}}.

If wEX,Yw\in E^{X,Y}, then wxw\leq x and for all yYy\in Y, wyw\perp y, so wy=0w\wedge y=0, and thus wy=wyx=0w\wedge y^{\prime}=w\wedge y\wedge x=0 for all yYy^{\prime}\in Y^{\prime}. Thus wEX,Yw\in E^{X,Y^{\prime}}.

Conversely, if wEX,Yw\in E^{X,Y^{\prime}}, then wxw\leq x and for all yYy^{\prime}\in Y, wyw\perp y^{\prime}, so wyx=0w\wedge y\wedge x=0. But wxw\leq x, so wyx=(wx)y=wy=0w\wedge y\wedge x=(w\wedge x)\wedge y=w\wedge y=0, so wEX,Yw\in E^{X,Y}.

Thus EX,Y=EX,YE^{X,Y}=E^{X,Y^{\prime}} as desired. ∎

Lemma 5.9.

Let μΛ\mu\in\Lambda and BμΛB\subseteq\mu\Lambda be finite. Then qμ=βBqβq_{\mu}=\bigvee_{\beta\in B}q_{\beta} if and only if 0=βB(qμqβ)0=\displaystyle\prod_{\beta\in B}(q_{\mu}-q_{\beta}).

Proof.

The result follows from a chain of equivalences:

qμ\displaystyle q_{\mu} =\displaystyle= βBqβ\displaystyle\bigvee_{\beta\in B}q_{\beta}
0\displaystyle 0 =\displaystyle= qμβBqβ by subtraction\displaystyle q_{\mu}-\bigvee_{\beta\in B}q_{\beta}\text{ by subtraction}
0\displaystyle 0 =\displaystyle= βB(qμqβ) by De Morgan’s laws\displaystyle\bigwedge_{\beta\in B}(q_{\mu}-q_{\beta})\text{ by De Morgan's laws}
0\displaystyle 0 =\displaystyle= βB(qμqβ) by definition of .\displaystyle\displaystyle\prod_{\beta\in B}(q_{\mu}-q_{\beta})\text{ by definition of }\bigwedge.

5.4 The Proof of Equivalence

Here we prove the following proposition. The main interest is the equivalence of the first and last criteria, and the middle statements are there to ease the proof.

Proposition 5.10.

If tt is a representation of Λ\Lambda, and σ=σt\sigma=\sigma_{t} is the associated representation of the semilattice E=E(Λ)E=E(\Lambda), then the following are equivalent:

  1. 1.

    σ\sigma is a tight representation in the EE-sense.

  2. 2.

    σ\sigma is a tight representation in the EE-sense when Y=Y=\emptyset.

  3. 3.

    σ\sigma is a tight representation in the EE-sense when Y=Y=\emptyset and X={μΛ}X=\{\mu\Lambda\}.

  4. 4.

    tt is a tight representation in the HH-sense.

Proof.

Certainly (1)(2)(3)(1)\Rightarrow(2)\Rightarrow(3). We will prove four non-obvious directions: (3)(4),(4)(3),(3)(2)(3)\Rightarrow(4),(4)\Rightarrow(3),(3)\Rightarrow(2) and (2)(1)(2)\Rightarrow(1).

(3)(4)(3)\Rightarrow(4) Fix some μΛ\mu\in\Lambda. We wish to show that if BμΛB\subseteq\mu\Lambda is exhaustive, then ZB={βΛ:βB}Z_{B}=\{\beta\Lambda:\beta\in B\} is a finite cover of E{μΛ},E^{\{\mu\Lambda\},\emptyset}. In particular, given such an exhaustive BB, if wE{μΛ},w\in E^{\{\mu\Lambda\},\emptyset} is nonzero, then taking αm(w)\alpha\in m(w), αwμΛ\alpha\in w\leq\mu\Lambda, so αμΛ\alpha\in\mu\Lambda, so since BB is exhaustive there is a βB\beta\in B such that αβ\alpha\Cap\beta, and by the intersection criterion, βΛw\beta\Lambda\Cap w. Thus ZBZ_{B} covers.

Then by (3), we have that zZBσ(z)=σ(μΛ)\bigvee_{z\in Z_{B}}\sigma(z)=\sigma(\mu\Lambda) and substituting σ(z)=σ(βΛ)=qβ\sigma(z)=\sigma(\beta\Lambda)=q_{\beta} and σ(μΛ)=qμ\sigma(\mu\Lambda)=q_{\mu}, we have

βBqβ=qμ\bigvee_{\beta\in B}q_{\beta}=q_{\mu}

which is equivalent to H-tightness by Lemma 5.9.

Given such an exhaustive BB, if wE{μΛ},w\in E^{\{\mu\Lambda\},\emptyset} is nonzero, then taking αm(w)\alpha\in m(w), αwμΛ\alpha\in w\leq\mu\Lambda, so αμΛ\alpha\in\mu\Lambda, so since BB is exhaustive there is a βB\beta\in B such that αβ\alpha\Cap\beta, and by the intersection criterion, βΛw\beta\Lambda\Cap w. Thus ZBZ_{B} covers.

(4)(3)(4)\Rightarrow(3) Fix some μΛ\mu\in\Lambda. We wish to show that if ZE{μΛ},Z\subseteq E^{\{\mu\Lambda\},\emptyset} is a finite cover, then BZ=zZm(z)B_{Z}=\bigcup_{z\in Z}m(z) is a finite exhaustive subset of μΛ\mu\Lambda. In particular, given αμΛ\alpha\in\mu\Lambda, we have that αΛE{μΛ},\alpha\Lambda\in E^{\{\mu\Lambda\},\emptyset}, so there is zZz\in Z such that zαΛz\Cap\alpha\Lambda. Then by the intersection property, there is a βm(z)\beta\in m(z) such that βα\beta\Cap\alpha. Thus BZB_{Z} is exhaustive for μΛ\mu\Lambda, as desired.

Then, by (4), and Lemma 5.9, we have that

qμ=βBZqβ.q_{\mu}=\bigvee_{\beta\in B_{Z}}q_{\beta}.

Substituting σ(μΛ)=qμ\sigma(\mu\Lambda)=q_{\mu} and σ(βΛ)=qβ\sigma(\beta\Lambda)=q_{\beta}, then regrouping the terms on the right, we have that

σ(μΛ)\displaystyle\sigma(\mu\Lambda) =\displaystyle= βBZσ(βΛ)\displaystyle\bigvee_{\beta\in B_{Z}}\sigma(\beta\Lambda)
=\displaystyle= zZβm(z)σ(βΛ)\displaystyle\bigvee_{z\in Z}\bigvee_{\beta\in m(z)}\sigma(\beta\Lambda)
=\displaystyle= zZσ(βm(z)βΛ)\displaystyle\bigvee_{z\in Z}\sigma(\bigvee_{\beta\in m(z)}\beta\Lambda)
=\displaystyle= zZσ(z)\displaystyle\bigvee_{z\in Z}\sigma(z)

as desired.

(32)(3\Rightarrow 2) Let XEX\subseteq E and let ZZ be a finite cover of EX,E^{X,\emptyset}. By the remark in [8] following Definition 2.5, we may assume without loss of generality that XX is a singleton, since otherwise we may replace XX with {xmin}\{x_{min}\} where xmin=xXxx_{min}=\bigwedge_{x\in X}x.

Let ZZ be a finite cover of EX,E^{X,\emptyset}. By Lemma 5.7, we may assume without loss of generality that Z={β1Λ,,βnΛ}Z=\{\beta_{1}\Lambda,...,\beta_{n}\Lambda\} for some β1,,βnΛ\beta_{1},...,\beta_{n}\in\Lambda. Let B={β1,βn}B=\{\beta_{1},...\beta_{n}\} and A=m(x)A=m(x).

Now, for each αA\alpha\in A and βB\beta\in B, let Zα,β={γΛ:γMCE(α,β)}Z_{\alpha,\beta}=\{\gamma\Lambda:\gamma\in MCE(\alpha,\beta)\}. Note that Zα,βZ_{\alpha,\beta} is finite and γΛZα,β\gamma\Lambda\in Z_{\alpha,\beta} is in αΛβΛ\alpha\Lambda\cap\beta\Lambda.

We wish to show that for fixed βB\beta\in B, αAZα,β\bigcup_{\alpha\in A}Z_{\alpha,\beta} is a cover for E{βΛ},E^{\{\beta\Lambda\},\emptyset} and for fixed αA\alpha\in A, that βBZα,β\bigcup_{\beta\in B}Z_{\alpha,\beta} is a cover for E{αΛ},E^{\{\alpha\Lambda\},\emptyset}. Each is a finite union of finite sets, hence finite, and is contained in the appropriate space. It then suffices to show that they are covering.

For the former set, fix some nonzero wE{βΛ},w\in E^{\{\beta\Lambda\},\emptyset}. Since ww is nonzero, there is some μm(w)\mu\in m(w). Then μΛwβΛx\mu\Lambda\leq w\leq\beta\Lambda\leq x, so μΛx=μΛ<\mu\Lambda\wedge x=\mu\Lambda<\infty, so μΛx\mu\Lambda\Cap x, and thus there is αm(x)\alpha\in m(x) such that μα\mu\Cap\alpha. That is, α\alpha and μ\mu have a common extension ν\nu. Since β\beta is a prefix of μ\mu, then ν\nu is a common extension of α\alpha and β\beta as well. Thus some prefix of ν\nu is a γMCE(α,β)\gamma\in MCE(\alpha,\beta), so ν\nu is a common extension of both γ\gamma and μ\mu. That is, γΛw\gamma\Lambda\Cap w, and γΛZα,β\gamma\Lambda\in Z_{\alpha,\beta}. Thus αAZα,β\bigcup_{\alpha\in A}Z_{\alpha,\beta} is a cover for E{βΛ},E^{\{\beta\Lambda\},\emptyset}, as desired.

For the latter set, fix some nonzero wE{αΛ},w\in E^{\{\alpha\Lambda\},\emptyset}. Then we know that wαΛxw\leq\alpha\Lambda\leq x, so wEX,w\in E^{X,\emptyset}, and because ZZ is a cover, there is some βΛZ\beta\Lambda\in Z with wβΛw\Cap\beta\Lambda. Then by the intersection criterion, there is a μm(w)\mu\in m(w) such that μβ\mu\Cap\beta. That is, μ\mu and β\beta have a common extension ν\nu. However, αμ\alpha\leq\mu as wE{αΛ},w\in E^{\{\alpha\Lambda\},\emptyset}, so ν\nu is a common extension of β\beta and α\alpha, and thus has a prefix which is some γMCE(β,α)\gamma\in MCE(\beta,\alpha). Thus ν\nu is a common extension of μ\mu and γ\gamma, so wγΛw\Cap\gamma\Lambda, and γΛZα,β\gamma\Lambda\in Z_{\alpha,\beta}. Thus βBZα,β\bigcup_{\beta\in B}Z_{\alpha,\beta} is a cover for E{αΛ},E^{\{\alpha\Lambda\},\emptyset}, as desired.

Now, using (3), the fact that these two sets cover gives us that

σ(βΛ)\displaystyle\sigma(\beta\Lambda) =\displaystyle= γαAZα,βσ(γΛ)\displaystyle\bigvee_{\gamma\in\bigcup_{\alpha\in A}Z_{\alpha,\beta}}\sigma(\gamma\Lambda)
=\displaystyle= αAγMCE(α,β)σ(γΛ)\displaystyle\bigvee_{\alpha\in A}\bigvee_{\gamma\in MCE(\alpha,\beta)}\sigma(\gamma\Lambda)

and

σ(αΛ)\displaystyle\sigma(\alpha\Lambda) =\displaystyle= γβBZα,βσ(γΛ)\displaystyle\bigvee_{\gamma\in\bigcup_{\beta\in B}Z_{\alpha,\beta}}\sigma(\gamma\Lambda)
=\displaystyle= βBγMCE(α,β)σ(γΛ)\displaystyle\bigvee_{\beta\in B}\bigvee_{\gamma\in MCE(\alpha,\beta)}\sigma(\gamma\Lambda)

Finally, using the fact that x=αAαΛx=\bigcup_{\alpha\in A}\alpha\Lambda, we have that

σ(x)\displaystyle\sigma(x) =\displaystyle= αAσ(αΛ)\displaystyle\bigvee_{\alpha\in A}\sigma(\alpha\Lambda)
=\displaystyle= αA(βBγMCE(α,β)σ(γΛ))\displaystyle\bigvee_{\alpha\in A}\left(\bigvee_{\beta\in B}\bigvee_{\gamma\in MCE(\alpha,\beta)}\sigma(\gamma\Lambda)\right)
=\displaystyle= βB(αAγMCE(α,β)σ(γΛ))\displaystyle\bigvee_{\beta\in B}\left(\bigvee_{\alpha\in A}\bigvee_{\gamma\in MCE(\alpha,\beta)}\sigma(\gamma\Lambda)\right)
=\displaystyle= βBσ(βΛ)\displaystyle\bigvee_{\beta\in B}\sigma(\beta\Lambda)
=\displaystyle= zZσ(z)\displaystyle\bigvee_{z\in Z}\sigma(z)

which is the E-tightness condition for EX,E^{X,\emptyset}.

(21)(2\Rightarrow 1) Let X,YEX,Y\subseteq E be finite, and let ZZ be a finite cover of EX,YE^{X,Y}. By Lemma 5.8, we may assume that YY is a subset of EX,E^{X,\emptyset}. Let Z=ZYZ^{\prime}=Z\cup Y. We claim that ZZ^{\prime} is a finite cover of EX,E^{X,\emptyset}. In particular, given wEX,w\in E^{X,\emptyset}, if wyw\perp y for all yYy\in Y, then wEX,Yw\in E^{X,Y}, so wzw\Cap z for some zZZz\in Z\subseteq Z^{\prime}. If instead w⟂̸yw\not\perp y for some yy, then wyw\Cap y for some yYZy\in Y\subseteq Z^{\prime}.

Then by (2), we know that

zZσ(z)=xXσ(x)\bigvee_{z^{\prime}\in Z^{\prime}}\sigma(z^{\prime})=\bigwedge_{x\in X}\sigma(x)

and meeting with yY¬σ(y)\bigwedge_{y\in Y}\neg\sigma(y), we have that

(zZσ(z))yY¬σ(y)=xXσ(x)yY¬σ(y)\left(\bigvee_{z^{\prime}\in Z^{\prime}}\sigma(z^{\prime})\right)\wedge\bigwedge_{y\in Y}\neg\sigma(y)=\bigwedge_{x\in X}\sigma(x)\wedge\bigwedge_{y\in Y}\neg\sigma(y)

Simplifying the lefthand side, we have that

(zZσ(z))yY¬σ(y)\displaystyle\left(\bigvee_{z^{\prime}\in Z^{\prime}}\sigma(z^{\prime})\right)\wedge\bigwedge_{y\in Y}\neg\sigma(y) =\displaystyle= (zZσ(z)yYσ(y))yY¬σ(y)\displaystyle\left(\bigvee_{z\in Z}\sigma(z)\vee\bigvee_{y^{\prime}\in Y}\sigma(y^{\prime})\right)\wedge\bigwedge_{y\in Y}\neg\sigma(y)
=\displaystyle= (zZσ(z)yY¬σ(y))(yYσ(y)yY¬σ(y))\displaystyle\left(\bigvee_{z\in Z}\sigma(z)\wedge\bigwedge_{y\in Y}\neg\sigma(y)\right)\vee\left(\bigvee_{y^{\prime}\in Y}\sigma(y^{\prime})\wedge\bigwedge_{y\in Y}\neg\sigma(y)\right)
=\displaystyle= (zZσ(z)yY¬σ(y))\displaystyle\left(\bigvee_{z\in Z}\sigma(z)\wedge\bigwedge_{y\in Y}\neg\sigma(y)\right)
=\displaystyle= zZ(σ(z)yY¬σ(y))\displaystyle\bigvee_{z\in Z}\left(\sigma(z)\wedge\bigwedge_{y\in Y}\neg\sigma(y)\right)

Now, for a fixed zZ,yYz\in Z,y\in Y, and for any αm(z)\alpha\in m(z), βm(y)\beta\in m(y), we have that αβ\alpha\perp\beta, so qαqβ=0q_{\alpha}q_{\beta}=0. Then σ(αΛ)¬σ(βΛ)=qα(1qβ)=qα\sigma(\alpha\Lambda)\wedge\neg\sigma(\beta\Lambda)=q_{\alpha}(1-q_{\beta})=q_{\alpha}, so σ(z)¬σ(y)=σ(z)\sigma(z)\vee\neg\sigma(y)=\sigma(z). Thus

zZ(σ(z)yY¬σ(y))=zZσ(z)\bigvee_{z\in Z}\left(\sigma(z)\wedge\bigwedge_{y\in Y}\neg\sigma(y)\right)=\bigvee_{z\in Z}\sigma(z)

so by transitivity,

zZσ(z)=xXσ(x)yY¬σ(y)\bigvee_{z\in Z}\sigma(z)=\bigwedge_{x\in X}\sigma(x)\wedge\bigwedge_{y\in Y}\neg\sigma(y)

as desired. This completes the proof that (2) implies (1).

References

  • [1] Selçuk Barlak and Xin Li. Cartan subalgebras and the UCT problem, II. Math. Ann., 378(1-2):255–287, 2020.
  • [2] Bruce Blackadar. Shape theory for CC^{\ast}-algebras. Math. Scand., 56(2):249–275, 1985.
  • [3] Nathan Brownlowe, Aidan Sims, and Sean T. Vittadello. Co-universal CC^{\ast}-algebras associated to generalised graphs. Israel J. Math., 193(1):399–440, 2013.
  • [4] Adam Clay and Dale Rolfsen. Ordered groups and topology, volume 176 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2016.
  • [5] Kenneth R. Davidson. CC^{*}-algebras by example, volume 6 of Fields Institute Monographs. American Mathematical Society, Providence, RI, 1996.
  • [6] Siegfried Echterhoff, S. Kaliszewski, John Quigg, and Iain Raeburn. A categorical approach to imprimitivity theorems for CC^{*}-dynamical systems. Mem. Amer. Math. Soc., 180(850):viii+169, 2006.
  • [7] Siegfried Echterhoff and John Quigg. Induced coactions of discrete groups on CC^{*}-algebras. Canad. J. Math., 51(4):745–770, 1999.
  • [8] R. Exel. Tight representations of semilattices and inverse semigroups. Semigroup Forum, 79(1):159–182, 2009.
  • [9] Ruy Exel. Amenability for Fell bundles. J. Reine Angew. Math., 492:41–73, 1997.
  • [10] Ruy Exel. Partial dynamical systems, Fell bundles and applications, volume 224 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2017.
  • [11] M. Hochster. Subsemigroups of amenable groups. Proc. Amer. Math. Soc., 21:363–364, 1969.
  • [12] Eberhard Kaniuth and Keith F. Taylor. Induced representations of locally compact groups, volume 197 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2013.
  • [13] David W. Kribs and Stephen C. Power. Free semigroupoid algebras. J. Ramanujan Math. Soc., 19(2):117–159, 2004.
  • [14] David W. Kribs and Stephen C. Power. The analytic algebras of higher rank graphs. Math. Proc. R. Ir. Acad., 106A(2):199–218, 2006.
  • [15] Alex Kumjian and David Pask. Higher rank graph CC^{\ast}-algebras. New York J. Math., 6:1–20, 2000.
  • [16] Roger C. Lyndon and Paul E. Schupp. Combinatorial group theory. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1977 edition.
  • [17] John C. Quigg. Discrete CC^{*}-coactions and CC^{*}-algebraic bundles. J. Austral. Math. Soc. Ser. A, 60(2):204–221, 1996.
  • [18] Iain Raeburn. Graph algebras, volume 103 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2005.
  • [19] Iain Raeburn. On graded CC^{*}-algebras. Bull. Aust. Math. Soc., 97(1):127–132, 2018.
  • [20] Camila F. Sehnem. On CC^{*}-algebras associated to product systems. J. Funct. Anal., 277(2):558–593, 2019.
  • [21] Jack Spielberg. Graph-based models for Kirchberg algebras. J. Operator Theory, 57(2):347–374, 2007.
  • [22] Jack Spielberg. Groupoids and CC^{*}-algebras for categories of paths. Trans. Amer. Math. Soc., 366(11):5771–5819, 2014.
  • [23] M. Takesaki. Theory of operator algebras. I, volume 124 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 2002. Reprint of the first (1979) edition, Operator Algebras and Non-commutative Geometry, 5.
  • [24] Jun Tomiyama. On the projection of norm one in WW^{\ast}-algebras. Proc. Japan Acad., 33:608–612, 1957.
\printindex