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

An extension property for noncommutative convex sets and duality for operator systems

Adam Humeniuk Department of Mathematics & Computing
Mount Royal University
Calgary, Alberta, Canada
[email protected]
Matthew Kennedy Department of Pure Mathematics
University of Waterloo
Waterloo, Ontario, Canada
[email protected]
 and  Nicholas Manor Department of Pure Mathematics
University of Waterloo
Waterloo, Ontario, Canada
[email protected]
Abstract.

We characterize inclusions of compact noncommutative convex sets with the property that every continuous affine function on the smaller set can be extended to a continuous affine function on the larger set with a uniform bound. As an application of this result, we obtain a simple geometric characterization of (possibly nonunital) operator systems that are dualizable, meaning that their dual can be equipped with an operator system structure. We further establish some permanence properties of dualizability, and provide a large new class of dualizable operator systems. These results are new even when specialized to ordinary compact convex sets.

2020 Mathematics Subject Classification. Primary: 46A55, 46L07, 46L52, 47L25, 47L50
Second author supported by NSERC Grant Number 418585.

1. Introduction

A unital operator system SS is a \ast-closed unital subspace of the bounded operators B(H)B(H) on a Hilbert space HH. In this paper, we assume that all operator spaces and operator systems are norm-complete. Choi and Effros [choi1977injectivity] gave an abstract characterization of unital operator systems as matrix ordered \ast-vector spaces which contain an archimedean matrix order unit. In light of this, it is natural to ask if the dual space SS^{\ast} is an abstract unital operator system.

The dual SS^{\ast} is at least a complete operator space, and inherits a \ast-operation and matrix ordering from SS. One says that SS^{\ast} is a matrix ordered operator space. However, SS^{\ast} typically fails to have an order unit in infinite dimensions. So, one requires a theory of nonunital operator systems if SS^{\ast} is to be an operator system.

Werner [werner2002subspaces] defined nonunital operator systems, which we hereafter refer to simply as “operator systems”, as matrix ordered operator spaces which embed completely isometrically and completely order isomorphically into B(H)B(H). Werner gave an abstract characterization of operator systems that extends the Choi-Effros axioms in the unital setting. One would hope that SS^{\ast} is such an operator system, but it turns out that this is too much to ask for.

It is natural to say that an operator system SS is dualizable if the dual matrix ordered operator space SS^{\ast} embeds into B(H)B(H) via a map which is both a complete order isomorphism and is completely bounded below. That is, SS^{\ast} can be re-normed with completely equivalent matrix norms in a way that makes it an operator system.

It is easy to construct examples of non-dualizable operator systems. For instance, the operator system SS consisting of 2×22\times 2 matrices with diagonal entries equal to zero, i.e.

S=span{E12,E21}M2,S=\text{span}\{E_{12},E_{21}\}\subseteq M_{2},

contains no nonzero positive elements, implying in particular that the set of positive elements in SS^{\ast} contains a line, precluding the existence of an order isomorphism into B(H)B(H).

Recently, Ng [ng2022dual] obtained an intrinsic characterization of dualizability for operator systems. Specifically, he proved that an operator system SS is dualizable if and only if it has the completely bounded positive decomposition property, meaning that there is a constant C>0C>0 such that for every nn\in\mathbb{N} and every self-adjoint element xMn(S)sax\in M_{n}(S)^{\text{sa}}, there are positives y,zMn(S)+y,z\in M_{n}(S)^{+} with x=yzx=y-z and y+zCx\|y\|+\|z\|\leq C\|x\|. Ng observed that every unital operator system SS is dualizable with C=2C=2. Further, using the continuous functional calculus, he showed that every C*-algebra is dualizable with C=1C=1. So the dualizable operator systems do form a large subclass of operator systems.


It turns out that the property of dualizability for operator systems is closely related to an important extension property for noncommutative convex sets. These are the main objects of interest in the theory of noncommutative convexity, which was recently introduced by Davidson and the second author [davidson2019noncommutative].

A noncommutative (nc) convex set KK over an operator space EE is a graded set

K=nKnnMn(E)K=\coprod_{n}K_{n}\subseteq\coprod_{n}M_{n}(E)

that is closed under direct sums and compression by scalar isometries. Equivalently, KK is closed under nc convex combinations, meaning that αixiαiKn\sum\alpha_{i}^{*}x_{i}\alpha_{i}\in K_{n} for every bounded family of points xiKnix_{i}\in K_{n_{i}} and every family of scalar matrices αiMni,n\alpha_{i}\in M_{n_{i},n}. If each KnK_{n} is compact, then KK is said to be compact.

An important point is that the above union is taken over all nκn\leq\kappa for some sufficiently large cardinal number κ\kappa, with the convention Mn=B(Hn)M_{n}=B(H_{n}) for a Hilbert space HnH_{n} of dimension nn. In the separable setting, it typically suffices to take κ=0\kappa=\aleph_{0}. This is in contrast to Wittstock’s definition of a matrix convex set, where the union is taken over all n<0n<\aleph_{0}. While every matrix convex set uniquely determines a nc convex set, the ability to work with points at infinity is essential for a robust theory (see [davidson2019noncommutative] for more details).

The duality theorem from [davidson2019noncommutative], which is essentially equivalent to a result of Webster-Winkler [webster1999krein] for compact matrix convex sets, asserts that the category of compact nc convex sets is dual to the category of unital operator systems. In particular, every unital operator system is isomorphic to the operator system A(K)A(K) of continuous affine nc functions on KK, where K=𝒮(S)K=\mathcal{S}(S) is the nc state space of SS, i.e.

Kn={φ:SMnφ is unital and completely positive}.K_{n}=\{\varphi:S\to M_{n}\mid\varphi\text{ is unital and completely positive}\}.

We say that an inclusion KLK\subseteq L of compact nc convex sets has the bounded extension property if there is a constant C>0C>0 such that every continuous affine nc function aA(K)a\in A(K) extends to a continuous affine nc function bA(L)b\in A(L) such that bCa\|b\|\leq C\|a\|. In Section 4, we establish the following characterization of the bounded extension property in terms of the operator space structure of the restriction map A(L)A(K)A(L)\to A(K), and in terms of the geometry of the inclusion KLK\subseteq L.

Theorem A.

Let KLK\subseteq L be an inclusion of compact nc convex sets. The following are equivalent:

  1. (1)

    The inclusion KLK\subseteq L has the bounded extension property.

  2. (2)

    The restriction map A(L)A(K)A(L)\to A(K) is an operator space quotient map.

  3. (3)

    There is a constant C>0C>0 such that (KK)spanLC(LL)(K-K)\cap\operatorname{span}_{\mathbb{R}}L\subseteq C(L-L).

The duality theorem from [kennedy2023nonunital] extends the duality theorem for unital operator systems to (possibly nonunital) operator systems. It asserts that the category of pointed compact nc convex sets is dual to the category of operator systems, where a pointed compact convex set is a pair (K,z)(K,z) consisting of a compact nc convex set KK and a distinguished point zK1z\in K_{1} that behaves like zero in a certain precise sense. In particular, every operator system is isomorphic to the operator system A(K,z)A(K,z) of continuous affine nc functions on KK that vanish at z=0z=0, where K=𝒬𝒮(S)K=\mathcal{QS}(S) is the nc quasistate space of SS, i.e.

Kn={φ:SMnφ is completely contractive and completely positive}.K_{n}=\{\varphi:S\to M_{n}\mid\varphi\text{ is completely contractive and completely positive}\}.

This result was utilized to provide additional insight on some recent results of Connes and van Suijlekom [connes2021spectral].

In Section 5, as an application of Theorem A and the duality between operator systems and pointed compact nc convex sets, we obtain a simple new geometric characterization of dualizability for operator systems.

Theorem B.

Let SS be an operator system with nc quasistate space KK. The following are equivalent:

  1. (1)

    The operator system SS is dualizable.

  2. (2)

    There is a constant C>0C>0 such that (K+K)+KCK(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K\subseteq CK.

  3. (3)

    The set (K+K)+K(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K is bounded.

In Section 6, we prove the dualizability of a large new class of operator systems. Specifically, we consider operator systems with nc quasistate spaces that are nc simplices in the sense of [kennedy2022noncommutative]. Although this class of operator systems contains all C*-algebras, it is much larger.

Theorem C.

Let SS be an operator system with nc quasistate space KK. If KK is an nc simplex and 0K10\in K_{1} is an extreme point in KK, then SS is dualizable.

In Section 7, we discuss the existing corresponding classical duality theory for function systems, which are the commutative operator systems. Here, the dual of a function system is again equivalent to a function system if we have (ordinary) bounded positive decomposition. By completeness, this is equivalent to ordinary positive generation. However, we don’t know if positive generation ensures completely bounded positive generation for function systems.

In Section 8, we relate the completely bounded positive decomposition property of an operator system to the property of being positively generated. In contrast to the classical situation for function systems in Section 7, positive generation at all matrix levels is not enough to imply the completely bounded positive decomposition property. We provide an example of a matrix ordered operator space that is positively generated but does not have the bounded positive decomposition property. However, we do show that positive generation of an operator system at the first level is enough to automatically imply positive generation at all matrix levels.

Finally, in Section 9, we establish some permanence properties of dualizability, showing that quotients and pushouts of dualizable operator systems are again dualizable.

Acknowledgements

The authors are grateful to David Blecher, Ken Davidson, Nico Spronk, Ivan Todorov and Vern Paulsen for their helpful comments and suggestions. The authors are also grateful to C.K. Ng for providing a preprint of [ng2022dual].

2. Background

2.1. Nonunital operator systems

All vector spaces in this paper are over \mathbb{C}, unless stated otherwise. If VV is a vector space and nn\in\mathbb{N}, we let Mn(V)M_{n}(V) be the vector space of n×nn\times n-matrices with entries in VV. We will typically identify Mn(V)M_{n}(V) with MnVM_{n}\otimes V, and write for instance

12x=(x00x),1_{2}\otimes x=\begin{pmatrix}x&0\\ 0&x\end{pmatrix},

where xVx\in V and 12M21_{2}\in M_{2} is the identity matrix. We will also use the notation

(V):=n1Mn(V)\mathcal{M}(V):=\coprod_{n\geq 1}M_{n}(V)

to denote the matrix universe over VV.

If VV is any normed vector space and r0r\geq 0, we will frequently use Br(V)B_{r}(V) to denote the closed ball in VV with radius rr and center 0V0\in V.

Following Ng [ng2022dual], we fix the following definitions. An operator space EE is a vector space equipped with a complete family of LL^{\infty}-matrix norms, which we will denote either by \|\cdot\|, E\|\cdot\|_{E}, or Mn(E)\|\cdot\|_{M_{n}(E)} as appropriate.

Definition 2.1.

A semi-matrix ordered operator space (X,P)(X,P) consists of an operator space XX equipped with a conjugate-linear completely isometric involution xxx\mapsto x^{\ast}, and a distinguished selfadjoint matrix convex cone P=n1PnnMn(X)saP=\coprod_{n\geq 1}P_{n}\subseteq\coprod_{n}M_{n}(X)^{\text{sa}} such that each PnP_{n} is norm-closed in Mn(X)M_{n}(X). Usually we omit the symbol PP and write Mn(X)+:=PnM_{n}(X)^{+}:=P_{n}. If in addition each cone Mn(X)+M_{n}(X)^{+} satisfies Mn(X)+(Mn(X)+)={0}M_{n}(X)^{+}\cap(-M_{n}(X)^{+})=\{0\}, then we say XX is a matrix ordered operator space. If XX is in addition a dual space X=(X)X=(X_{\ast})^{\ast}, we say XX is a dual matrix ordered operator space if the positive cones Mn(X)+M_{n}(X)^{+} are weak-\ast closed.

Definition 2.2.

A semi-matrix ordered operator space XX is positively generated if

Mn(X)sa=Mn(X)+Mn(X)+M_{n}(X)^{\text{sa}}=M_{n}(X)^{+}-M_{n}(X)^{+}

for all n1n\geq 1.

Example 2.3.

If XX is a positively generated matrix ordered operator space, then XX^{\ast} is naturally a dual matrix ordered operator space with the standard norm and order structure that identifies

Mn(X)\displaystyle M_{n}(X^{\ast}) CB(X,Mn) isometrically, and\displaystyle\cong\text{CB}(X,M_{n})\quad\text{ isometrically, and}
Mn(X)+\displaystyle M_{n}(X^{\ast})^{+} CP(X,Mn).\displaystyle\cong\text{CP}(X,M_{n}).
Definition 2.4.

Let XX and YY be matrix ordered operator spaces, and let φ:XY\varphi:X\to Y be a linear map. For any n1n\geq 1, φ\varphi induces a linear map φn:Mn(X)Mn(Y)\varphi_{n}:M_{n}(X)\to M_{n}(Y). We say that φ\varphi is completely bounded, contractive, bounded below, isometric, positive, or a complete order isomorphism when each induced map φn\varphi_{n} satisfies the same property uniformly in nn. If φ\varphi is completely bounded below and positive, we say φ\varphi is a complete embedding. If φ\varphi is completely isometric and positive, we say φ\varphi is a completely isometric embedding. If φ\varphi is also a linear isomorphism, we call φ\varphi a complete isomorphism or completely isometric isomorphism as appropriate.

The class of all matrix ordered operator spaces forms a category, where one usually chooses the morphisms to be completely contractive and completely positive (ccp) maps, or completely bounded and completely positive (cbp) maps. In the interest of readability, we hereafter adopt the convention that “completely contractive and positive" always means “completely bounded and completely positive”, and similarly for “completely bounded and positive”. That is, “completely" modifies both the words “contractive” and “positive”. Since we have no need to consider maps which are positive but not completely positive, there should be no risk of confusion.

Example 2.5.

Let SS be a unital operator system, i.e. an \ast-matrix ordered space with archimedian matrix order unit 1S1_{S}. Then SS is a matrix ordered operator space with norm

xMn(S)=inf{t0|(t(1n1s)xxt(1n1s))0 in M2n(S)sa}.\|x\|_{M_{n}(S)}=\inf\left\{t\geq 0\;\middle|\;\begin{pmatrix}t(1_{n}\otimes 1_{s})&x\\ x^{\ast}&t(1_{n}\otimes 1_{s})\end{pmatrix}\geq 0\text{ in }M_{2n}(S)^{\text{sa}}\right\}.

This norm agrees with the induced norm from any unital complete order embedding SB(H)S\subseteq B(H). In particular, for any Hilbert space HH, the space B(H)B(H) is a unital operator system.

Definition 2.6.

Let SS be a matrix ordered operator space. We say that SS is a quasi-operator system if there is a complete embedding SB(H)S\to B(H) for some Hilbert space HH, and that SS is a operator space if there is a completely isometric embedding SB(H)S\to B(H). If SS is in addition a dual matrix ordered operator space, then we say SS is a dual (quasi-)operator system if there is a weak-\ast homeomorphic (complete embedding) completely isometric embedding into some B(H)B(H).

That is, a quasi-operator system SS is a matrix ordered operator space which is completely isomorphic to an operator system. Put another way, one can choose a completely equivalent system of norms on SS, for which SS embeds completely isometrically and order isomorphically into B(H)B(H).

2.2. Pointed noncommutative convex sets

Suppose that E=(E)E=(E_{\ast})^{\ast} is a dual operator space. Let

(E):=n1Mn(E),\mathcal{M}(E):=\coprod_{n\geq 1}M_{n}(E),

where the union is taken over all cardinals n1n\geq 1 up to some fixed cardinal α\alpha at least as large as the density character of EE. (In practice we suppress α\alpha.) When nn is infinite, we take the convention Mn:=B(Hn)M_{n}:=B(H_{n}), where HnH_{n} is a Hilbert space of dimension nn. By naturally identifying

Mn(E)=CB(E,E),M_{n}(E)=\text{CB}(E_{\ast},E),

we may equip each Mn(E)M_{n}(E) with its corresponding point-weak-\ast topology. Note that if E=MkE=M_{k}, this is the just the usual weak-\ast topology on Mn(Mk)MnkM_{n}(M_{k})\cong M_{nk}.

Definition 2.7.

We say that a graded subset

K=n1Kn(E)K=\coprod_{n\geq 1}K_{n}\subseteq\mathcal{M}(E)

is an nc convex set if for every norm-bounded family (xi)Kni(x_{i})\in K_{n_{i}} and every family of matrices αiMni,n\alpha_{i}\in M_{n_{i},n} which satisfies

(1) iαiαi=1n,\sum_{i}\alpha_{i}^{\ast}\alpha_{i}=1_{n},

we have

(2) iαixiαiKn.\sum_{i}\alpha_{i}^{\ast}x_{i}\alpha_{i}\in K_{n}.

Here the sums (1) and (2) are required to converge in the point-weak-\ast topologies on MnM_{n} and Mn(E)M_{n}(E), respectively. We say in addition that KK is a compact nc convex set if each matrix level KnK_{n} is point-weak-\ast compact in Mn(E)M_{n}(E).

Usually we refer to the sum in (2) as an nc convex combination of the points xix_{i}. Succinctly, an nc convex set is one that is closed under nc convex combinations. It is equivalent to require only that KK is closed under direct sums (1) in which the αi\alpha_{i}’s are co-isometries with orthogonal domain projections, and compressions (2) when there is only one αi\alpha_{i}, which must be an isometry.

Definition 2.8.

Let KK and LL be nc convex sets. A function a:KLa:K\to L is an affine nc function if it is graded

a(Kn)Ln, for all n1,a(K_{n})\subseteq L_{n},\quad\text{ for all }n\geq 1,

and respects nc convex combinations, i.e. whenever xiKx_{i}\in K are bounded and αi\alpha_{i} are scalar matrices of appropriate sizes such that iαixiαi\sum_{i}\alpha_{i}^{\ast}x_{i}\alpha_{i}, then

a(iαixiαi)=iαia(xi)αi.a\Big{(}\sum_{i}\alpha_{i}^{\ast}x_{i}\alpha_{i}\Big{)}=\sum_{i}\alpha_{i}^{\ast}a(x_{i})\alpha_{i}.

The function aa is continuous if each restriction a|Kna|_{K_{n}} is point-weak-\ast continuous.

Kadison’s classical representation theorem [kadison1951representation] asserts that the category of function systems, or archimedean order unit spaces, is equivalent to the category of compact convex sets with continuous affine functions as morphisms. The following noncommutative generalization from [davidson2019noncommutative, Theorem 3.2.5] for compact nc convex sets is essentially equivalent to [webster1999krein, Proposition 3.5] for compact matrix convex sets.

Theorem 2.9.

The category of unital operator systems with ucp maps as morphisms is contravariantly equivalent to the category of compact nc convex sets with continuous affine nc functions as morphisms. On objects, the essential inverse functors send an operator system SS to its nc state space

𝒮(S)=n1{φ:SMnφ is unital and completely positive},\mathcal{S}(S)=\coprod_{n\geq 1}\{\varphi:S\to M_{n}\mid\varphi\text{ is unital and completely positive}\},

and send a compact nc convex set KK to the operator system

A(K)={a:K=()a is a continuous affine nc function}.A(K)=\{a:K\to\mathcal{M}=\mathcal{M}(\mathbb{C})\mid a\text{ is a continuous affine nc function}\}.

The operator system structure and norm on A(K)A(K) is pointwise, i.e. one identifies Mn(A(K))A(K,(Mn))M_{n}(A(K))\cong A(K,\mathcal{M}(M_{n})), and declares a matrix valued affine nc function if it takes positive values at every point. The order unit is the “constant function" xKn1nMnx\in K_{n}\mapsto 1_{n}\in M_{n}. Both essential inverse functors act on morphisms by precomposition. That is, if π:ST\pi:S\to T is a ucp map between operator systems, then the corresponding map on nc state spaces sends ρ:TMn\rho:T\to M_{n} to ρπ:SMn\rho\pi:S\to M_{n}. Likewise, if a:KLa:K\to L is affine nc, then ffa:A(L)A(K)f\mapsto f\circ a:A(L)\to A(K) is affine nc.

Recently, the second and third authors together with Kim [kennedy2023nonunital] settled the question of Kadison duality for nonunital operator systems. The key challenge is that in the absence of order units, if SS is a nonunital operator system then one must remember the whole nc quasistate space

𝒬𝒮(S)=n1{φ:SMnφ is contractive and completely positive}\mathcal{QS}(S)=\coprod_{n\geq 1}\{\varphi:S\to M_{n}\mid\varphi\text{ is contractive and completely positive}\}

and consider pointed affine nc functions which fix the zero quasistates.

Definition 2.10.

Let KK be a compact nc convex set and fix a distinguished point zz. We let

A(K,z)={aA(K)a(z)=0}A(K,z)=\{a\in A(K)\mid a(z)=0\}

denote the operator system of affine nc functions which vanish at zz. We say that the pair (K,z)(K,z) is a pointed nc convex set if the natural evaluation map

K\displaystyle K 𝒬𝒮(A(K,z))\displaystyle\to\mathcal{QS}(A(K,z))
x\displaystyle x (aa(x))\displaystyle\mapsto(a\mapsto a(x))

is surjective (and hence bijective).

The following result from [kennedy2023nonunital, Theorem 4.9] provides a nonunital generalization of the duality between the category of unital operator systems and the category of compact nc convex sets. The main subtlety is that while the correspondence S(𝒬𝒮(S),0)S\mapsto(\mathcal{QS}(S),0) is a full and faithful functor, it is only essentially surjective onto the pointed compact nc convex sets.

Theorem 2.11.

The category of operator systems with ccp maps as morphisms is contravariantly equivalent to the category of pointed compact nc convex sets with pointed continuous affine nc functions as morphism. On objects, the essential inverse functors send an operator system SS to is pointed nc quasistate space (𝒬𝒮(S),0)(\mathcal{QS}(S),0), and send a pointed compact nc convex set KK to the operator system A(K,z)A(K,z) of pointed continuous affine nc functions on (K,z)(K,z).

Again, on morphisms the essential inverse functors in Theorem 2.11 act in the natural way by precomposition on either affine nc functions or on nc quasistates.

3. Quotients of matrix ordered spaces

3.1. Operator space quotients

In this section, if VV is a normed vector space and r>0r>0, then we let Br(V)B_{r}(V) denote the closed ball in VV with radius rr and center 0.

Here, we recall the basic theory of quotients for operator spaces. If EE is an operator space, and FEF\subseteq E is a closed subspace, then the quotient vector space E/FE/F is an operator space where the matrix norms isometrically identify Mn(E/F)M_{n}(E/F) with the standard Banach space quotient Mn(E)/Mn(F)M_{n}(E)/M_{n}(F).

Definition 3.1.

Let φ:EF\varphi:E\to F be a completely bounded map between operator spaces EE and FF. We will say that φ:EF\varphi:E\to F is a operator space quotient map with constant C>0C>0 if any of the following equivalent conditions hold

  • (1)

    B1(Mn(F))φn(BC(Mn(E)))¯=Cφn(B1(Mn(E)))¯B_{1}(M_{n}(F))\subseteq\overline{\varphi_{n}(B_{C}(M_{n}(E)))}=C\overline{\varphi_{n}(B_{1}(M_{n}(E)))} for all nn\in\mathbb{N}.

  • (2)

    B1(Mn(F))(C+ϵ)φn(B1(Mn(E)))B_{1}(M_{n}(F))\subseteq(C+\epsilon)\cdot\varphi_{n}(B_{1}(M_{n}(E))) for all nn\in\mathbb{N} and every ϵ>0\epsilon>0.

  • (3)

    The induced map φ~:E/kerφF\tilde{\varphi}:E/\ker\varphi\to F is an isomorphism and satisfies φ~1cbC\|\tilde{\varphi}^{-1}\|_{\text{cb}}\leq C.

The equivalence of (1) and (2) follows from a standard series argument using completeness of EE. We will simply say operator space quotient map if we have no need to refer to CC explicitly.

The following fact is standard in operator space theory, but we provide a proof for completeness.

Proposition 3.2.

Let φ:EF\varphi:E\to F be a completely bounded map between operator spaces EE and FF. The map φ\varphi is a quotient map with constant C>0C>0 if and only if the dual map φ:FE\varphi^{\ast}:F^{\ast}\to E^{\ast} is completely bounded below by 1/C1/C. Moreover, in this case, φ\varphi^{\ast} is weak-\ast homeomorphism onto its range.

Proof.

Suppose that Cφn(B1(Mn(E)))C\varphi_{n}(B_{1}(M_{n}(E))) is dense in B1(Mn(F))B_{1}(M_{n}(F)) for every nn. Given fMm(F)CB(F,Mm)f\in M_{m}(F^{\ast})\cong\text{CB}(F,M_{m}), approximating unit vector yB1(Mn(F))y\in B_{1}(M_{n}(F)) with vectors of the form φ(x)\varphi(x) for xBC(E)x\in B_{C}(E) shows that fcbCφm(f)cb\|f\|_{\text{cb}}\leq C\|\varphi^{\ast}_{m}(f)\|_{\text{cb}}.

Conversely, suppose that

n1B1(Mn(F))Cn1φn(B1(Mn(E))).\coprod_{n\geq 1}B_{1}(M_{n}(F))\not\subseteq C\coprod_{n\geq 1}\varphi_{n}(B_{1}(M_{n}(E))).

By the Effros-Winkler nc Bipolar theorem [effros1997matrix], there are m,n1m,n\geq 1, an xCB1(Mn(E))x\in CB_{1}(M_{n}(E)), and an fMm(F)CB(F,Mm)f\in M_{m}(F^{\ast})\cong\text{CB}(F,M_{m}), such that

Refk(y)1mkfor all k1,yB1(Mk(F)),\operatorname{Re}f_{k}(y)\leq 1_{mk}\quad\text{for all }k\geq 1,y\in B_{1}(M_{k}(F)),

and yet Refn(x)1mn\operatorname{Re}f_{n}(x)\not\leq 1_{mn}. It follows that f1\|f\|\leq 1, but xC\|x\|\leq C and fn(φn(x))>1\|f_{n}(\varphi_{n}(x))\|>1, so φm(f)>fcb/C\|\varphi^{\ast}_{m}(f)\|>\|f\|_{\text{cb}}/C. This shows φ\varphi^{\ast} is not completely bounded below by 1/C1/C.

Finally, if φ\varphi is an operator space quotient map, it is bounded and surjective, and so its dual map φ\varphi^{\ast} is weak-\ast homeomorphic onto its range. ∎

3.2. Matrix ordered operator space quotients

Definition 3.3.

Let XX be a matrix ordered operator space. We call a closed subspace JXJ\subseteq X a kernel if it is the kernel of a ccp map φ:XY\varphi:X\to Y for some matrix ordered operator space YY. In this case, we define an matrix ordered operator space structure on the operator space X/JX/J with involution

(x+J):=x+J(x+J)^{\ast}:=x^{\ast}+J

and matrix order

Mn(X/J)+:={x+Mn(J)xMn(X)+}¯,M_{n}(X/J)^{+}:=\overline{\{x+M_{n}(J)\mid x\in M_{n}(X)^{+}\}},

where the closure is taken in the quotient norm topology on

Mn(X/J)Mn(X)/Mn(J).M_{n}(X/J)\cong M_{n}(X)/M_{n}(J).
Proposition 3.4.

If XX is a matrix ordered operator space, and J=kerφJ=\ker\varphi is a kernel, then X/JX/J is a matrix ordered operator space.

Proof.

Since the involution on XX is completely isometric and JJ is selfadjoint, it follows that the involution on Mn(X/J)M_{n}(X/J) is completely isometric. It is straightforward to check that X/JX/J is a matrix ordered operator space. To prove that it is a matrix ordered operator space, suppose x+JMn(X/J)+(Mn(X/J)+)x+J\in M_{n}(X/J)^{+}\cap(-M_{n}(X/J)^{+}). Then for any ϵ\epsilon, there are y,zMn(X)+y,z\in M_{n}(X)^{+} with xy+Mn(J),x+z+Mn(J)<ϵ\|x-y+M_{n}(J)\|,\|x+z+M_{n}(J)\|<\epsilon. Hence

φn(x)φn(y)xy+Mn(J)<ϵ\|\varphi_{n}(x)-\varphi_{n}(y)\|\leq\|x-y+M_{n}(J)\|<\epsilon

and similarly φn(x)+φn(z)<ϵ\|\varphi_{n}(x)+\varphi_{n}(z)\|<\epsilon. Since φ\varphi is cp, φn(y),φn(z)0\varphi_{n}(y),\varphi_{n}(z)\geq 0. As ϵ\epsilon is arbitrary and YY is a matrix ordered operator space, this shows

φn(x)Mn(Y)+¯(Mn(Y)+¯)=Mn(Y)+(Mn(Y)+)={0}.\varphi_{n}(x)\in\overline{M_{n}(Y)^{+}}\cap(-\overline{M_{n}(Y)^{+}})=M_{n}(Y)^{+}\cap(-M_{n}(Y)^{+})=\{0\}.

Therefore xMn(kerφ)=Mn(J)x\in M_{n}(\ker\varphi)=M_{n}(J), and so x+Mn(J)=0x+M_{n}(J)=0. This shows

Mn(X/J)+(Mn(X/J)+)={0},M_{n}(X/J)^{+}\cap(-M_{n}(X/J)^{+})=\{0\},

so X/JX/J is a matrix ordered operator space. ∎

One can form a category of matrix ordered operator spaces with morphisms as either completely contractive and positive (ccp) or completely bounded and positive (cbp) maps.

Definition 3.5.

Let XX and YY be matrix ordered operator spaces, and let φ:XY\varphi:X\to Y be a cbp map. We say that φ\varphi is a matrix ordered operator space quotient map with constant C>0C>0 if for all nn\in\mathbb{N} we have both

  • (1)

    B1(Mn(Y))Cφn(B1(Mn(X)))¯B_{1}(M_{n}(Y))\subseteq C\overline{\varphi_{n}(B_{1}(M_{n}(X)))}, and

  • (2)

    Mn(Y)+=φn(Mn(X))+¯.M_{n}(Y)^{+}=\overline{\varphi_{n}(M_{n}(X))^{+}}.

For brevity, we will usually simply refer to φ\varphi as a quotient map, whenever it is clear that we are speaking only in the context of matrix ordered operator spaces.

That is, a matrix ordered operator space quotient map is just an operator space quotient map that maps the positives (densely) onto the positives at each matrix level. Comparing to Definition 3.1.(2), a quotient map is surjective. Each map φn:Mn(X)Mn(Y)\varphi_{n}:M_{n}(X)\to M_{n}(Y) is therefore open and closed, and since the positive cones Mn(X)+M_{n}(X)^{+} and Mn(Y)+M_{n}(Y)^{+} are norm-closed, it follows that φn(Mn(X)+)\varphi_{n}(M_{n}(X)^{+}) is closed and φn(Mn(X)+)=Mn(Y)+\varphi_{n}(M_{n}(X)^{+})=M_{n}(Y)^{+} for all nn. That is, the closure in condition (2) is redundant. The first thing to show is that such maps are in fact categorical quotients in the category of matrix ordered operator spaces.

Proposition 3.6.

Let φ:XY\varphi:X\to Y be a cbp map between matrix ordered operator spaces. The following are equivalent.

  • (1)

    The map φ\varphi is a quotient map with constant C>0C>0.

  • (2)

    The dual map φ:YX\varphi^{\ast}:Y^{\ast}\to X^{\ast} is completely bounded below by 1/C1/C and a complete order injection.

  • (3)

    With J=kerφJ=\ker\varphi, the induced map φ~:X/JY\tilde{\varphi}:X/J\to Y such that

    X{X}Y{Y}X/J{X/J}φ\scriptstyle{\varphi}q\scriptstyle{q}φ~\scriptstyle{\tilde{\varphi}}

    commutes is an isomorphism with cbp inverse satisfying φ~1cbC\|\tilde{\varphi}^{-1}\|_{\text{cb}}\leq C.

  • (4)

    For every matrix ordered operator space ZZ and cbp map ψ:XZ\psi:X\to Z with kerφkerψ\ker\varphi\subseteq\ker\psi, there is a unique cbp map ψ~:YZ\tilde{\psi}:Y\to Z making the diagram

    X{X}Z{Z}Y{Y}ψ\scriptstyle{\psi}φ\scriptstyle{\varphi}ψ~\scriptstyle{\tilde{\psi}}

    commute, with ψ~cbCψcb\|\tilde{\psi}\|_{\text{cb}}\leq C\|\psi\|_{\text{cb}}.

In this case, φ\varphi^{\ast} is weak-\ast homeomorphic onto its range.

Proof.

To prove (1) and (2) are equivalent, after invoking Proposition 3.2, it suffices to show that φ\varphi^{\ast} is a complete order injection if and only if Condition (2) in Definition 3.5 holds. Note that because φ\varphi is completely positive, so is φ\varphi^{\ast}. Suppose φn(Mn(X)+)\varphi_{n}(M_{n}(X)^{+}) is dense in Mn(Y)+M_{n}(Y)^{+} for every n0n\geq 0. Let fMm(Y)f\in M_{m}(Y^{\ast}) with φm(f)0\varphi_{m}^{\ast}(f)\geq 0. Given n1n\geq 1 and yMn(Y)+y\in M_{n}(Y)^{+}, approximating yy with a net of points of the form φn(xi)\varphi_{n}(x_{i}) for xiMn(X)+x_{i}\in M_{n}(X)^{+} shows that

fn(y)=limifn(φn(xi))=limi(φm(f))n(xi)0.f_{n}(y)=\lim_{i}f_{n}(\varphi_{n}(x_{i}))=\lim_{i}(\varphi^{\ast}_{m}(f))_{n}(x_{i})\geq 0.

This shows f0f\geq 0.

Conversely, suppose that φn(Mn(X)+)\varphi_{n}(M_{n}(X)^{+}) is not dense in Mn(Y)+M_{n}(Y)^{+} for some n1n\geq 1. By the Effros-Winkler nc Bipolar Theorem [effros1997matrix] applied to the closed nc convex sets

k1φk(Mn(X)+)¯k1Mk(Y)+,\coprod_{k\geq 1}\overline{\varphi_{k}(M_{n}(X)^{+})}\quad\not\supseteq\quad\coprod_{k\geq 1}M_{k}(Y)^{+},

there is a selfadjoint matrix functional fMm(Y)saf\in M_{m}(Y^{\ast})^{\text{sa}} such that fk(y)1mkf_{k}(y)\geq-1_{mk} for every kk and every yMk(Y)+y\in M_{k}(Y)^{+}, but

fn(z)1mnf_{n}(z)\not\geq-1_{mn}

for some zφn(Mn(X)+)¯Mk(Y)+z\in\overline{\varphi_{n}(M_{n}(X)^{+})}\setminus M_{k}(Y)^{+}. A rescaling argument shows that f0f\geq 0 in Mk(Y)M_{k}(Y). However, approximating xx by points of the form φn(x)\varphi_{n}(x), xMn(X)+x\in M_{n}(X)^{+} shows that φm(f)\varphi^{\ast}_{m}(f) cannot be positive. Hence, φ\varphi^{\ast} is not a complete order isomorphism.

If φ\varphi is a quotient map with constant C>0C>0, then it follows immediately from the definition of the matrix order and matrix norms on X/JX/J that φ~:X/JY\tilde{\varphi}:X/J\to Y is a complete order and norm isomorphism with φ~1cbC\|\tilde{\varphi}^{-1}\|_{\text{cb}}\leq C. Conversely, note that by definition the quotient map q:XX/Jq:X\to X/J is a quotient map with constant 11. Hence, if φ~\tilde{\varphi} is a complete order isomorphism with φ~1cbC\|\tilde{\varphi}^{-1}\|_{\text{cb}}\leq C, it follows that φ=φ~q\varphi=\tilde{\varphi}\circ q is a quotient map with constant CC. This proves (1) and (3) are equivalent.

To show (3) and (4) are equivalent, it is enough to note that the quotient map q:XX/Jq:X\to X/J satisfies the universal property (4) with constant C=1C=1. In detail, if (3) holds, composing the universal map from (4) applied to q:XX/Jq:X\to X/J with φ~1\tilde{\varphi}^{-1} shows that (4) holds for φ\varphi with constant CC. Conversely, if (4) holds, then it holds for both φ\varphi and qq, and there are induced maps φ~:X/JY\tilde{\varphi}:X/J\to Y and q~:YX/J\tilde{q}:Y\to X/J with q~φ~cb\|\tilde{q}\|\leq\|\tilde{\varphi}\|_{\text{cb}} and q~Cqcb=C\|\tilde{q}\|\leq C\|q\|_{\text{cb}}=C. Comparing diagrams shows q~=φ~1\tilde{q}=\tilde{\varphi}^{-1}, and φ~\tilde{\varphi} is an isomorphism. ∎

Condition (4) in Proposition 3.6 shows that a matrix ordered operator space quotient map is a categorical quotient in the category of matrix ordered operator spaces with cbp maps as morphisms. Moreover, the norm bound shows that a quotient map with constant C=1C=1 is a categorical quotient in the subcategory of matrix ordered operator spaces with ccp maps as morphisms.

Remark 3.7.

Every unital operator system is a matrix ordered operator space, and so if φ:ST\varphi:S\to T is a ucp map between operator systems with J=kerSJ=\ker S, we may form the quotient matrix ordered operator space S/kerφS/\ker\varphi, but there is no a priori guarantee that this quotient is again an operator system. The matrix ordered operator space quotient is generally not isomorphic to the unital operator system quotient defined by Kavruk, Paulsen, Todorov, and Tomforde [kavruk2013quotients]. For example, they show in [kavruk2013quotients, Example 4.4] that the order norm on the unital operator system quotient need not be completely equivalent to the quotient operator space norm.

4. Extension property for compact nc convex sets

If K=nKnK=\coprod_{n}K_{n} is a compact nc convex set, we will define

spanK:=n1spanKn(E).\operatorname{span}_{\mathbb{R}}K:=\coprod_{n\geq 1}\operatorname{span}_{\mathbb{R}}K_{n}\subseteq\mathcal{M}(E).

The set spanK\operatorname{span}_{\mathbb{R}}K is also nc convex, but need not be closed in EE.

Lemma 4.1.

Let 0K(E)0\in K\subseteq\mathcal{M}(E) be a compact nc convex set containing 0. Let KKK-K denote the levelwise Minkowski difference of KK with itself. Then we have inclusions

KK2ncconv¯(K(K))KK.\frac{K-K}{2}\subseteq\overline{\mathrm{ncconv}}(K\cup(-K))\subseteq K-K.

Consequently, ncconv¯(K(K))spanK\overline{\mathrm{ncconv}}(K\cup(-K))\subseteq\operatorname{span}_{\mathbb{R}}K.

Proof.

It is immediate that (KK)/2ncconv(K(K))ncconv¯(K(K))(K-K)/2\subseteq\mathrm{ncconv}(K\cup(-K))\subseteq\overline{\mathrm{ncconv}}(K\cup(-K)). Given zncconv(K(K))nz\in\mathrm{ncconv}(K\cup(-K))_{n}, we can write

z=iαixiαijβjyjβjz=\sum_{i}\alpha_{i}^{\ast}x_{i}\alpha_{i}-\sum_{j}\beta_{j}^{\ast}y_{j}\beta_{j}

for uniformly bounded families {xi},{yi}\{x_{i}\},\{y_{i}\} in KK and matrix coefficients satisfying iαiαi+jβjβj=1n\sum_{i}\alpha_{i}^{\ast}\alpha_{i}+\sum_{j}\beta_{j}^{\ast}\beta_{j}=1_{n}. Since 0K0\in K and iαiαi1\sum_{i}\alpha_{i}^{\ast}\alpha_{i}\leq 1, we have x:=iαixiαiKnx:=\sum_{i}\alpha_{i}^{\ast}x_{i}\alpha_{i}\in K_{n}. Similarly y:=jβjyjβjKy:=\sum_{j}\beta_{j}^{\ast}y_{j}\beta_{j}\in K, and so z=xyz=x-y is in (KK)n=KnKn(K-K)_{n}=K_{n}-K_{n}. Therefore

ncconv(K(K))KK,\mathrm{ncconv}(K\cup(-K))\subseteq K-K,

and since the latter is compact, ncconv¯(K(K))KK\overline{\mathrm{ncconv}}(K\cup(-K))\subseteq K-K. ∎

When 0K0\in K, by extending the inclusion map KnMn(A(K,0))K\subseteq\coprod_{n}M_{n}(A(K,0)^{\ast}) linearly at each level, we will think of elements in (spanK)n(\operatorname{span}_{\mathbb{R}}K)_{n} as nc functionals in

Mn(A(K,0))=CB(A(K,0),Mn).M_{n}(A(K,0)^{\ast})=\text{CB}(A(K,0),M_{n}).
Proposition 4.2.

Let 0K(E)0\in K\subseteq\mathcal{M}(E) be a compact nc convex set in a dual operator space E=(E)E=(E_{\ast})^{\ast}. For each nn\in\mathbb{N}, the inclusion KQS(A(K,0))K\to\text{QS}(A(K,0)) extends uniquely to a well-defined affine nc isomorphism

η:n1spanKnn1Mn(A(K,0))sa\eta:\coprod_{n\geq 1}\operatorname{span}_{\mathbb{R}}K_{n}\to\coprod_{n\geq 1}M_{n}(A(K,0)^{\ast})^{\text{sa}}

which is levelwise linear. The norm unit ball in Mn(A(K,0))saM_{n}(A(K,0)^{\ast})^{\text{sa}} is

B1(Mn(A(K,0)))=CC(A(K,0),Mn)=ncconv¯(η(K)(η(K)))n,B_{1}(M_{n}(A(K,0)^{\ast}))=\text{CC}(A(K,0),M_{n})=\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K)))_{n},

and for each nn, η\eta is homeomorphic on KnKnK_{n}-K_{n}.

Proof.

Since KnK_{n} is convex, we have spanKn={sxtyx,yKn,s,t0}\operatorname{span}_{\mathbb{R}}K_{n}=\{sx-ty\mid x,y\in K_{n},s,t\geq 0\}. Given sxtyspanKnsx-ty\in\operatorname{span}_{\mathbb{R}}K_{n}, we define

η(sxty)(a)=sa(x)ta(y)\eta(sx-ty)(a)=sa(x)-ta(y)

for aA(K,0)a\in A(K,0). Since such functions aa are affine and satisfy a(0)=0a(0)=0, it follows that η|Kn\eta|_{K_{n}} is well-defined and linear, and that η\eta is affine nc. Since EE_{\ast} contains a separating family of functionals, which restrict to affine nc functions in A(K,0)A(K,0), the map η\eta is injective.

Next we will show the closed unit ball is

B1(Mn(A(K,0))sa)=ncconv¯(η(K)(η(K)))nB_{1}(M_{n}(A(K,0)^{\ast})^{\text{sa}})=\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K)))_{n}

for every nn. That is, if LL is the compact nc convex set

L=n1Ln=n1B1(Mn(A(K,0))sa),L=\coprod_{n\geq 1}L_{n}=\coprod_{n\geq 1}B_{1}(M_{n}(A(K,0)^{\ast})^{\text{sa}}),

we want to show L=ncconv¯(η(K)(η(K)))L=\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K))). Since η(K)\eta(K) consists of nc quasistates on A(K,0)A(K,0), it is clear that Lncconv¯(η(K)(η(K)))L\supseteq\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K))). To prove the reverse inclusion, by the nc Bipolar theorem of Effros and Winkler [effros1997matrix], it suffices to suppose that for some nn\in\mathbb{N} and aMn(A(K,0))saa\in M_{n}(A(K,0))^{\text{sa}} that we have

φn(a)1k1n=1kn\varphi_{n}(a)\leq 1_{k}\otimes 1_{n}=1_{kn}

for all kk\in\mathbb{N} and all φncconv¯(η(K)(η(K)))\varphi\in\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K))), and then show that ψn(a)1k1n\psi_{n}(a)\leq 1_{k}\otimes 1_{n} for all kk and all ψLk\psi\in L_{k}. Because ncconv¯(η(K)(η(K)))\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K))) contains both η(K)\eta(K) and η(K)-\eta(K), we have

1kna(x)1kn-1_{kn}\leq a(x)\leq 1_{kn}

for all kk and all xKkx\in K_{k}. Hence aMn(A(K,0))1\|a\|_{M_{n}(A(K,0))}\leq 1, and so ψn(a)a1kn1kn\psi_{n}(a)\leq\|a\|1_{kn}\leq 1_{kn} for every ψL\psi\in L. This proves L=ncconv¯(η(K)(η(K)))L=\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K))), and consequently η\eta is also surjective. Since η\eta is homeomorphic on KK and KKK-K is (levelwise) compact, it is easy to check that η\eta is continuous on each KnKnK_{n}-K_{n}. Being a continuous injection on a compact Hausdorff space, the map η|KnKn\eta|_{K_{n}-K_{n}} is automatically a homeomorphism onto its range. ∎

Recall that the pair (K,0)(K,0) in Proposition 4.2 is a pointed nc convex set exactly when we have

n1B1(Mn(A(K,0))+)=QS(A(K,0))=η(K).\coprod_{n\geq 1}B_{1}(M_{n}(A(K,0)^{\ast})^{+})=\text{QS}(A(K,0))=\eta(K).

In practice, we will often identify Mn(A(K,0))saM_{n}(A(K,0)^{\ast})^{\text{sa}} with spanKn\operatorname{span}_{\mathbb{R}}K_{n} and so omit the symbol η\eta. Note that since η\eta is homeomorphic on KKncconv¯(K(K))K-K\supseteq\overline{\mathrm{ncconv}}(K\cup(-K)) (Lemma 4.1), we are free to identify

ncconv¯(η(K)(η(K)))=η(ncconv¯(K(K))).\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K)))=\eta(\overline{\mathrm{ncconv}}(K\cup(-K))).

That is, when we identify Mn(A(K,0))sa=spanKnM_{n}(A(K,0)^{\ast})^{\text{sa}}=\operatorname{span}_{\mathbb{R}}K_{n}, the unit ball of Mn(A(K,0))saM_{n}(A(K,0)^{\ast})^{\text{sa}} is ncconv¯(K(K))n\overline{\mathrm{ncconv}}(K\cup(-K))_{n}.

For a closed convex set XX in a vector space VV containing 0, we use the usual Minkowski functional

γX(v):=inf{t0vtX},vV.\gamma_{X}(v):=\inf\{t\geq 0\mid v\in tX\},\qquad v\in V.

If 0K=nKn0\in K=\bigcup_{n}K_{n} is a compact nc convex set over a dual operator space EE, we will use the shorthand

γK(x)=γKn(x)\gamma_{K}(x)=\gamma_{K_{n}}(x)

when xMn(E)x\in M_{n}(E).

Definition 4.3.

[taylor1972extension] If XX is a closed convex set in some vector space VV, then for dVd\in V, we define the width of VV (with respect to dd) or the dd-width of VV as

|X|d\displaystyle|X|_{d} :=sup{t0tdXX}\displaystyle:=\sup\{t\geq 0\mid td\in X-X\}
=1γXX(d).\displaystyle=\frac{1}{\gamma_{X-X}(d)}.
Definition 4.4.

If K=nKn(E)K=\coprod_{n}K_{n}\subseteq\mathcal{M}(E) is a closed nc convex set over a dual operator space EE, then for any nn and any dMn(E)d\in M_{n}(E) we define the width

|K|d:=|Kn|d=1γKK(d),|K|_{d}:=|K_{n}|_{d}=\frac{1}{\gamma_{K-K}(d)},
Lemma 4.5.

If 0K(E)0\in K\subseteq\mathcal{M}(E) is a compact nc convex set containing 0, then for dMn(E)d\in M_{n}(E), we have |K|d>0|K|_{d}>0 if and only if dspanKd\in\operatorname{span}_{\mathbb{R}}K. Moreover, for dspanKd\in\operatorname{span}_{\mathbb{R}}K, we have

1|K|dη(d)Mn(A(K,0))2|K|d.\frac{1}{|K|_{d}}\leq\|\eta(d)\|_{M_{n}(A(K,0)^{\ast})}\leq\frac{2}{|K|_{d}}.

That is, d1/|K|d=1/|Kn|dd\mapsto 1/|K|_{d}=1/|K_{n}|_{d} defines a norm on spanKn\operatorname{span}_{\mathbb{R}}K_{n} that is equivalent to the norm induced by the isomorphism η:spanKnMn(A(K,0))sa\eta:\operatorname{span}_{\mathbb{R}}K_{n}\to M_{n}(A(K,0)^{\ast})^{\text{sa}}.

Proof.

By Lemma 4.1, we have inclusions

KK2ncconv¯(K(K))KK.\frac{K-K}{2}\subseteq\overline{\mathrm{ncconv}}(K\cup(-K))\subseteq K-K.

It follows that for dspanKd\in\operatorname{span}_{\mathbb{R}}K, we have

2γKK(d)γncconv¯(K(K))(d)γKK(d).2\gamma_{K-K}(d)\geq\gamma_{\overline{\mathrm{ncconv}}(K\cup(-K))}(d)\geq\gamma_{K-K}(d).

By definition, γKK=1/|K|d\gamma_{K-K}=1/|K|_{d}. By Proposition 4.2, the norm unit ball of Mn(A(K,0))saM_{n}(A(K,0)^{\ast})^{\text{sa}} is

ncconv¯(η(K)(η(K)))=η(ncconv¯(K(K))),\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K)))=\eta(\overline{\mathrm{ncconv}}(K\cup(-K))),

and hence γncconv¯(K(K))(d)=γncconv¯(η(K)(η(K)))(η(d))=η(d)\gamma_{\overline{\mathrm{ncconv}}(K\cup(-K))}(d)=\gamma_{\overline{\mathrm{ncconv}}(\eta(K)\cup(-\eta(K)))}(\eta(d))=\|\eta(d)\|. ∎

Given compact nc convex sets 0LK0\in L\subseteq K. The restriction map ρ:A(K,0)A(L,0)\rho:A(K,0)\to A(L,0) is always completely contractive and positive, and has dense range. When is this map an operator space quotient map? Equivalently, this means there is a constant C>0C>0 so that any affine nc function gMn(A(L,0))g\in M_{n}(A(L,0)) extends to an affine nc function ff on all of KK with

f|L=gandfMn(A(K,0))CgMn(A(L,0)).f|_{L}=g\quad\text{and}\quad\|f\|_{M_{n}(A(K,0))}\leq C\|g\|_{M_{n}(A(L,0))}.

Here is a noncommutative version of [taylor1972extension, Theorem 1].

Proposition 4.6.

Let 0LK(E)0\in L\subseteq K\subseteq\mathcal{M}(E) be compact nc convex sets containing 0. The following are equivalent

  • (1)

    The restriction map A(K)A(L)A(K)\to A(L) is an operator space quotient map.

  • (2)

    The restriction map ρ:A(K,0)A(L,0)\rho:A(K,0)\to A(L,0) is an operator space quotient map.

  • (3)

    The dual map ρ:A(L,0)A(K,0)\rho^{\ast}:A(L,0)^{\ast}\to A(K,0)^{\ast} is completely bounded below.

  • (4)

    There is a constant c>0c>0 such that for all n1n\geq 1 and all dMn(E)d\in M_{n}(E) with |L|d>0|L|_{d}>0, we have

    |L|dc|K|d.|L|_{d}\geq c|K|_{d}.
  • (5)

    There is a constant C>0C>0 such that

    (KK)spanLC(LL).(K-K)\cap\operatorname{span}_{\mathbb{R}}L\subseteq C(L-L).
Proof.

Clearly (1) implies (2). Suppose ρ:A(K,0)A(L,0)\rho:A(K,0)\to A(L,0) is an operator space quotient map with constant C0C\geq 0. Given aA(L)a\in A(L), we have aa(0)1A(L)A(L,0)a-a(0)\otimes 1_{A(L)}\in A(L,0). Thus there is a bA(K,0)b\in A(K,0) with b|L=aa(0)1A(L)b|_{L}=a-a(0)\otimes 1_{A(L)} and bCaa(0)1A(L)2Ca\|b\|\leq C\|a-a(0)\otimes 1_{A(L)}\|\leq 2C\|a\|. Then, b+a(0)1A(K)A(K)b+a(0)\otimes 1_{A(K)}\in A(K) restricts to aa on LL and satisfies b+a(0)1A(K)b+a(2C+1)a\|b+a(0)\otimes 1_{A(K)}\|\leq\|b\|+\|a\|\leq(2C+1)\|a\|. This proves A(K)A(L)A(K)\to A(L) is an operator space quotient map with constant 2C+12C+1, so (2) implies (1).

The equivalence of (2) and (3) is Proposition 3.2. To prove (3) is equivalent to (4), first note by taking real and imaginary parts that (3) occurs if and only if the restrictions ρn:Mn(A(L,0))saMn(A(K,0))sa\rho^{\ast}_{n}:M_{n}(A(L,0)^{\ast})^{\text{sa}}\to M_{n}(A(K,0)^{\ast})^{\text{sa}} are bounded below by a universal constant. By Proposition 4.2, we may identify

spanLn=Mn(A(L,0))saandspanKn=Mn(A(K,0))sa.\operatorname{span}_{\mathbb{R}}L_{n}=M_{n}(A(L,0)^{\ast})^{\text{sa}}\quad\text{and}\quad\operatorname{span}_{\mathbb{R}}K_{n}=M_{n}(A(K,0)^{\ast})^{\text{sa}}.

With this identification, ρ\rho^{\ast} is just the inclusion map spanLnspanKn\operatorname{span}_{\mathbb{R}}L_{n}\to\operatorname{span}_{\mathbb{R}}K_{n}. By Lemma 4.5, the induced norms on spanL\operatorname{span}_{\mathbb{R}}L and spanK\operatorname{span}_{\mathbb{R}}K are completely equivalent to d1/|L|dd\mapsto 1/|L|_{d} and d1/|K|dd\mapsto 1/|K|_{d}. Thus the dual map ρ\rho^{\ast} is completely bounded below if and only if for some constant c>0c>0, we have

1c|K|d1|L|d|L|dc|K|d\frac{1}{c|K|_{d}}\leq\frac{1}{|L|_{d}}\iff|L|_{d}\geq c|K|_{d}

whenever dspanL={d(E)|L|d>0}d\in\operatorname{span}_{\mathbb{R}}L=\{d\in\mathcal{M}(E)\mid|L|_{d}>0\}, by Lemma 4.5.

For d(E)d\in\mathcal{M}(E), recall that |K|d=1γKK(d)|K|_{d}=\frac{1}{\gamma_{K-K}(d)} and |L|d=1γLL(d)|L|_{d}=\frac{1}{\gamma_{L-L}(d)}. Hence condition (3) holds if and only if

γLL|spanL1cγKK|spanL=γc(KK)|spanL.\gamma_{L-L}|_{\operatorname{span}_{\mathbb{R}}L}\leq\frac{1}{c}\gamma_{K-K}|_{\operatorname{span}_{\mathbb{R}}L}=\gamma_{c(K-K)}|_{\operatorname{span}_{\mathbb{R}}L}.

Using only the definition of the Minkowski gauges γKK\gamma_{K-K} and γLL\gamma_{L-L}, this holds if and only if

c(KK)spanLLL.c(K-K)\cap\operatorname{span}_{\mathbb{R}}L\subseteq L-L.

Hence condition (4) holds with constant c>0c>0 if and only if condition (5) holds with constant C=1/c>0C=1/c>0. ∎

Note that for any general inclusion LKL\subseteq K of compact nc convex sets, we can freely translate to assume 0L0\in L and apply Proposition 4.6. Thus conditions (1), (4), and (5) are equivalent in total generality. Note also that we do not require in 4.6 that (L,0)(L,0) and (K,0)(K,0) are pointed nc convex sets.

Example 4.7.

It is possible that the restriction map A(K,0)A(L,0)A(K,0)\to A(L,0) in Proposition 4.6 is surjective but not an operator space quotient. For instance, let EE be an infinite dimensional Banach space. Let max(E)\max(E) and min(E)\min(E) denote EE equipped with its maximal and minimal operator space norms which restrict to the usual norm on EE [effros2022theory, Section 3.3]. There are standard operator space dualities max(E)min(E)\max(E)^{\ast}\cong\min(E^{\ast}) and min(E)max(E)\min(E)^{\ast}\cong\max(E^{\ast}). As EE is infinite dimensional, the maximal and minimal matrix norms on EE are not completely equivalent [paulsen2002completely, Theorem 14.3]. So, the identity map max(E)min(E)\max(E)\to\min(E) is surjective and not an operator space quotient map. Consider the minimal and maximal nc unit balls

K=n1B1(Mn(min(E)))andL=n1B1(Mn(max(E)))K=\coprod_{n\geq 1}B_{1}(M_{n}(\min(E^{\ast})))\quad\text{and}\quad L=\coprod_{n\geq 1}B_{1}(M_{n}(\max(E^{\ast})))

in (E)\mathcal{M}(E^{\ast}). By the dualities max(E)min(E)\max(E)^{\ast}\cong\min(E^{\ast}) and min(E)max(E)\min(E)^{\ast}\cong\max(E^{\ast}), we have

A(K,0)max(E)andA(L,0)min(E)A(K,0)\cong\max(E)\quad\text{and}\quad A(L,0)\cong\min(E)

completely isometrically. The restriction map A(K,0)A(L,0)A(K,0)\to A(L,0) is just the identity map max(E)min(E)\max(E)\to\min(E), which is surjective, but not an operator space quotient map.

Example 4.8.

Proposition 4.6 provides a guarantee that every matrix-valued affine nc function on LL lifts to an affine nc function on KK with a complete norm bound. However, there is no guarantee that we can lift a positive affine function to one that is positive. For instance, the restriction map of function systems

A([1,1],0)A([0,1],0)A([-1,1],0)\to A([0,1],0)

is an operator space quotient map with constant c=1c=1, but does not map the positives onto the positives because A([1,1],0)+={0}A([-1,1],0)^{+}=\{0\}.

Proposition 4.9.

Let 0LK(E)0\in L\subseteq K\subseteq\mathcal{M}(E) be compact nc convex sets such that (L,0)(L,0) and (K,0)(K,0) are pointed compact nc convex sets. Let ρ:A(K,0)A(L,0)\rho:A(K,0)\to A(L,0) be the restriction map. The following are equivalent

  • (1)

    For all n1n\geq 1, ρn(Mn(A(K,0)+))¯=Mn(A(L,0))+\overline{\rho_{n}(M_{n}(A(K,0)^{+}))}=M_{n}(A(L,0))^{+}.

  • (2)

    The dual map ρ:A(L,0)A(K,0)\rho^{\ast}:A(L,0)^{\ast}\to A(K,0)^{\ast} is a complete order embedding.

  • (3)

    KspanL+LK\cap\operatorname{span}_{\mathbb{R}}L\subseteq\mathbb{R}_{+}L.

  • (4)

    Kncconv¯(L(L))=LK\cap\overline{\mathrm{ncconv}}(L\cup(-L))=L.

Proof.

To prove (1)(2)(1)\iff(2), consider the closed nc convex sets

P=n1Mn(A(L,0))+andQ=n1ρn(Mn(A(K,0))+)¯.P=\coprod_{n\geq 1}M_{n}(A(L,0))^{+}\quad\text{and}\quad Q=\coprod_{n\geq 1}\overline{\rho_{n}(M_{n}(A(K,0))^{+})}.

By the nc Bipolar theorem of Effros and Winkler [effros1997matrix], we have Q=PQ=P if and only if their nc polars QπQ^{\pi} and PπP^{\pi} are equal. But by scaling, we have

Pπ\displaystyle P^{\pi} ={φMk(A(L,0))k,φn(b)1nk for all n1,bPn}\displaystyle=\{\varphi\in M_{k}(A(L,0)^{\ast})\mid k\in\mathbb{N},\real\varphi_{n}(b)\leq 1_{nk}\text{ for all }n\geq 1,b\in P_{n}\}
={φMk(A(L,0))k,φ0}\displaystyle=\{\varphi\in M_{k}(A(L,0)^{\ast})\mid k\in\mathbb{N},\real\varphi\leq 0\}

and similarly

Qπ\displaystyle Q^{\pi} ={φMk(A(L,0))k,φn(ρn(a))1nk for all n1,aMn(A(K,0))+}\displaystyle=\{\varphi\in M_{k}(A(L,0)^{\ast})\mid k\in\mathbb{N},\real\varphi_{n}(\rho_{n}(a))\leq 1_{nk}\text{ for all }n\geq 1,a\in M_{n}(A(K,0))^{+}\}
={φMk(A(L,0))k,ρk(φ)0}.\displaystyle=\{\varphi\in M_{k}(A(L,0)^{\ast})\mid k\in\mathbb{N},\real\rho^{\ast}_{k}(\varphi)\leq 0\}.

Thus P=QP=Q if and only if ρ\rho^{\ast} is a complete order injection.

When we identify A(K,0)=spanK1A(K,0)^{\ast}=\operatorname{span}_{\mathbb{R}}K_{1} and A(L,0)=spanL1A(L,0)^{\ast}=\operatorname{span}_{\mathbb{R}}L_{1} as in Proposition 4.2, the dual map ρ:spanLspanK\rho^{\ast}:\operatorname{span}_{\mathbb{R}}L\to\operatorname{span}_{\mathbb{R}}K is just the inclusion map. Since (K,0)(K,0) and (L,0(L,0) are pointed, the positive cones in Mn(A(K,0))=spanKnM_{n}(A(K,0)^{\ast})=\operatorname{span}_{\mathbb{R}}K_{n} and Mn(A(L,0))=spanLnM_{n}(A(L,0)^{\ast})=\operatorname{span}_{\mathbb{R}}L_{n} are just +Kn\mathbb{R}_{+}K_{n} and +Ln\mathbb{R}_{+}L_{n}, respectively. Hence the inclusion map is a complete order injection if and only if we have

+KspanL=+L.\mathbb{R}_{+}K\cap\operatorname{span}_{\mathbb{R}}L=\mathbb{R}_{+}L.

A rescaling argument shows that this is equivalent to

KspanL+L,K\cap\operatorname{span}_{\mathbb{R}}L\subseteq\mathbb{R}_{+}L,

and so (2) and (3) are equivalent.

If Kncconv¯(L(L))=LK\cap\overline{\mathrm{ncconv}}(L\cup(-L))=L, then scaling gives

+(KspanL)=+KspanL=+L,\mathbb{R}_{+}(K\cap\operatorname{span}_{\mathbb{R}}L)=\mathbb{R}_{+}K\cap\operatorname{span}_{\mathbb{R}}L=\mathbb{R}_{+}L,

which is again equivalent to (3), so (4) implies (3). Now suppose that KspanL+LK\cap\operatorname{span}_{\mathbb{R}}L\subseteq\mathbb{R}_{+}L. Clearly LKncconv¯(L(L))L\subseteq K\cap\overline{\mathrm{ncconv}}(L\cup(-L)). Conversely, if xKncconv¯(L(L))x\in K\cap\overline{\mathrm{ncconv}}(L\cup(-L)), then by Lemma 4.1, we also have xKspanL=+Lx\in K\cap\operatorname{span}_{\mathbb{R}}L=\mathbb{R}_{+}L. Hence

xncconv¯(L(L))+L.x\in\overline{\mathrm{ncconv}}(L\cup(-L))\cap\mathbb{R}_{+}L.

Because (L,0)(L,0) is pointed, this implies xLx\in L, proving that (3) implies (4). ∎

Combining Propositions 4.6 and 4.9 yields

Theorem 4.10.

Let (L,0)(L,0) and (K,0)(K,0) be pointed compact nc convex sets with LK(E)L\subseteq K\subseteq\mathcal{M}(E). The following are equivalent.

  • (1)

    The restriction map A(K,0)A(L,0)A(K,0)\to A(L,0) is a matrix ordered operator space quotient map.

  • (2)

    There is a constant C>0C>0 such that

    • (i)

      (KK)spanLC(LL)(K-K)\cap\operatorname{span}_{\mathbb{R}}L\subseteq C(L-L), and

    • (ii)

      KspanL+LK\cap\operatorname{span}_{\mathbb{R}}L\subseteq\mathbb{R}_{+}L.

5. Dualizability via nc quasistate spaces

Recall that the trace class operators 𝒯(H)=B(H)\mathcal{T}(H)=B(H)_{\ast} inherit a matrix ordered operator space structure via the embedding 𝒯(H)=B(H)B(H)\mathcal{T}(H)=B(H)_{\ast}\subseteq B(H)^{\ast}, where B(H)(B(H))B(H)\cong(B(H)_{\ast})^{\ast} completely isometrically and order isomorphically. By Ng’s [ng2022dual] results, since B(H)B(H) is a C*-algebra, B(H)B(H)^{\ast} is an operator system, and so 𝒯(H)=B(H)B(H)\mathcal{T}(H)=B(H)_{\ast}\subseteq B(H)^{\ast} is also an operator system. The nc quasistate space of 𝒯(H)\mathcal{T}(H) is the compact nc convex set

𝒫(H):=nMn(B(H))1+=n{xMn(B(H))|x0,x1}.\mathcal{P}(H):=\coprod_{n}M_{n}(B(H))_{1}^{+}=\coprod_{n}\{x\in M_{n}(B(H))|x\geq 0,\|x\|\leq 1\}.

Applying Theorem 4.10 and Proposition 3.6 yields the following extrinsic geometric characterization of dualizability for an operator system.

Corollary 5.1.

Let SS be an operator system with pointed nc quasistate space (K,0)(K,0), and let HH be a Hilbert space. The following are equivalent.

  • (1)

    There is a weak-\ast homeorphic complete embedding SB(H)S^{\ast}\to B(H).

  • (2)

    There is a matrix ordered operator space quotient map 𝒯(H)S\mathcal{T}(H)\to S.

  • (3)

    There is a pointed continuous affine nc injection φ:(K,0)𝒫(H)\varphi:(K,0)\to\mathcal{P}(H) such that

    • (i)

      (𝒫(H)𝒫(H))spanφ(K)C(φ(K)φ(K))(\mathcal{P}(H)-\mathcal{P}(H))\cap\operatorname{span}_{\mathbb{R}}\varphi(K)\subseteq C(\varphi(K)-\varphi(K)) for some constant C>0C>0, and

    • (ii)

      𝒫(H)spanφ(K)+φ(K)\mathcal{P}(H)\cap\operatorname{span}_{\mathbb{R}}\varphi(K)\subseteq\mathbb{R}_{+}\varphi(K).

Definition 5.2.

Let EE be an ordered \ast-Banach space with closed positive cone E+E^{+}. We say EE is α\alpha-positively generated or simply α\alpha-generated for a constant α>0\alpha>0 if for each xEsax\in E^{\text{sa}}, we can write

x=yzx=y-z

for y,zE+y,z\in E^{+} satisfying y+zαx\|y\|+\|z\|\leq\alpha\|x\|. Or, equivalently,

B1(E)=αconv(B1(E+)(B1(E+))).B_{1}(E)=\alpha\operatorname{conv}(B_{1}(E^{+})\cup(-B_{1}(E^{+}))).

If XX is a matrix ordered operator space, then we say XX is completely α\alpha-generated if each matrix level Mn(X)M_{n}(X) is α\alpha-generated.

In [ng2022dual, Theorem 3.9], Ng proved that an operator system SS is dualizable if and only if it is completely α\alpha-generated for some α>0\alpha>0. The following definition is the dual property of α\alpha-generation.

Definition 5.3.

An ordered \ast-Banach space EE is α\alpha-normal for some α>0\alpha>0 if for all x,y,zEsax,y,z\in E^{\text{sa}},

(3) xyzyαmax{x,z}.x\leq y\leq z\implies\|y\|\leq\alpha\max\{\|x\|,\|z\|\}.

If XX is a matrix ordered operator space, then XX is completely α\alpha-normal if each matrix level Mn(X)M_{n}(X) is α\alpha-normal.

The condition of α\alpha-normality can be viewed as a strict requirement about how the norm and order structure on EE interact. Normality means that “order bounds" xyzx\leq y\leq z should imply “norm bounds" yαmax{x,z}\|y\|\leq\alpha\max\{\|x\|,\|z\|\}. If one does not care about the exact value of α\alpha, it is enough to check the normality identity (3) on positive elements in the special case x=0x=0.

Proposition 5.4.

If EE is an ordered \ast-Banach space, then EE is α\alpha-normal for some α>0\alpha>0 if and only if there is a constant β>0\beta>0 such that

(4) 0xyxβy0\leq x\leq y\implies\|x\|\leq\beta\|y\|

for x,yE+x,y\in E^{+}.

Proof.

If EE is α\alpha-normal, then (4) holds with β=α\beta=\alpha. Conversely, suppose (4) holds, and let xyzx\leq y\leq z in EsaE^{\text{sa}}. Then 0yxzx0\leq y-x\leq z-x, and so yxβzx\|y-x\|\leq\beta\|z-x\|. Then, we get the bound

y\displaystyle\|y\| yx+x\displaystyle\leq\|y-x\|+\|x\|
βzx+x\displaystyle\leq\beta\|z-x\|+\|x\|
β(z+x)+x\displaystyle\leq\beta(\|z\|+\|x\|)+\|x\|
(2β+1)max{x,z},\displaystyle\leq(2\beta+1)\max\{\|x\|,\|z\|\},

proving EE is (2β+1)(2\beta+1)-normal. ∎

Proposition 5.5.

Let XX be a matrix ordered operator space, with dual matrix ordered operator space XX^{\ast}, and let α>0\alpha>0. If XX is completely α\alpha-generated, then XX^{\ast} is completely 2α2\alpha-normal. Conversely, if XX^{\ast} is completely α\alpha-normal, then XX is completely 2α2\alpha-generated.

Proof.

Suppose that XX is completely α\alpha-generated. Let kk\in\mathbb{N} and suppose x,y,zMk(X)sax,y,z\in M_{k}(X^{\ast})^{\text{sa}} satisfy xyzx\leq y\leq z in the dual matrix ordering on XX^{\ast}. By definition of the dual norm, we have

yMk(X)=sup{\llanglea,x\rranglen1,aMn(X)sa},\|y\|_{M_{k}(X^{\ast})}=\sup\{\|\llangle a,x\rrangle\|\,\mid\,n\geq 1,a\in M_{n}(X)^{\text{sa}}\},

where \llangle,\rrangle\llangle\cdot,\cdot\rrangle denotes the matrix pairing between (X)\mathcal{M}(X) and (X)\mathcal{M}(X^{\ast}) defined by

Mm(X)×Mn(X)Mm×n:(a,x)\llanglea,x\rrangle=[xk,l(ai,j)].M_{m}(X)\times M_{n}(X^{\ast})\to M_{m\times n}:(a,x)\to\llangle a,x\rrangle=[x_{k,l}(a_{i,j})].

Given nn\in\mathbb{N} and aMn(X)saa\in M_{n}(X)^{\text{sa}}, we can write a=bca=b-c where b,cMn(X)+b,c\in M_{n}(X)^{+} satisfy b+cαa\|b\|+\|c\|\leq\alpha\|a\|. Then, we have the operator inequality

\llanglea,y\rrangle\displaystyle\llangle a,y\rrangle =\llangleb,y\rrangle\llanglec,y\rrangle\displaystyle=\llangle b,y\rrangle-\llangle c,y\rrangle
\llangleb,z\rrangle\llanglec,x\rrangle\displaystyle\leq\llangle b,z\rrangle-\llangle c,x\rrangle
(zb+xc)1nk\displaystyle\leq(\|z\|\|b\|+\|x\|\|c\|)1_{nk}
(x+z)αa1nk.\displaystyle\leq(\|x\|+\|z\|)\alpha\|a\|1_{nk}.

Symmetrically,

\llanglea,y\rrangle\displaystyle\llangle a,y\rrangle \llangleb,x\rrangle\llanglec,z\rrangle\displaystyle\geq\llangle b,x\rrangle-\llangle c,z\rrangle
(xb+zc)1nk\displaystyle\geq-(\|x\|\|b\|+\|z\|\|c\|)1_{nk}
(x+z)αa1nk.\displaystyle\geq-(\|x\|+\|z\|)\alpha\|a\|1_{nk}.

It follows that

\llanglea,y\rrangle(x+z)αa.\|\llangle a,y\rrangle\|\leq(\|x\|+\|z\|)\alpha\|a\|.

Since aa was arbitrary, this shows yα(x+z)2αmax{x,z}\|y\|\leq\alpha(\|x\|+\|z\|)\leq 2\alpha\max\{\|x\|,\|z\|\}, proving XX^{\ast} is completely 2α2\alpha-normal.

Now suppose XX^{\ast} is completely 2α2\alpha-normal. Consider the closed matrix convex subsets

K\displaystyle K :=n1B1(Mn(X)sa)=B1((X)sa),\displaystyle:=\coprod_{n\geq 1}B_{1}(M_{n}(X)^{\text{sa}})=B_{1}(\mathcal{M}(X)^{\text{sa}}),
K+\displaystyle K^{+} :=n1B1(Mn(X)+)=K(X)+,\displaystyle:=\coprod_{n\geq 1}B_{1}(M_{n}(X)^{+})=K\cap\mathcal{M}(X)^{+},
L\displaystyle L :=ncconv¯(K+(K+))\displaystyle:=\overline{\mathrm{ncconv}}\left(K^{+}\cup(-K^{+})\right)

of (X)\mathcal{M}(X). We will show that KαLK\subseteq\alpha L.

To prove KαLK\subseteq\alpha L, by the selfadjoint version of the nc separation Theorem of Effros and Winkler [davidson2019noncommutative, Theorem 2.4.1], it suffices to show that the selfadjoint nc polars

Kρ:=n1{xMn(X)sa\llanglea,x\rrangle1nk for all k1,xKk}K^{\rho}:=\coprod_{n\geq 1}\{x\in M_{n}(X)^{\text{sa}}\mid\llangle a,x\rrangle\leq 1_{nk}\text{ for all }k\geq 1,x\in K_{k}\}

and LρL^{\rho} (defined similarly) satisfy LραKρL^{\rho}\subseteq\alpha K^{\rho}. The relevant selfadjoint polars are

Kρ\displaystyle K^{\rho} =k1B1(Mk(X)),\displaystyle=\coprod_{k\geq 1}B_{1}(M_{k}(X^{\ast})),
(K+)ρ\displaystyle(K^{+})^{\rho} =Kρ(X)+\displaystyle=K^{\rho}-\mathcal{M}(X^{\ast})^{+}
=k1{xMk(X)saxy for some yKρ}, and\displaystyle=\coprod_{k\geq 1}\{x\in M_{k}(X^{\ast})^{\text{sa}}\mid x\leq y\text{ for some }y\in K^{\rho}\},\quad\text{ and }
Lρ\displaystyle L^{\rho} =(K+)ρ(K+)ρ\displaystyle=(K^{+})^{\rho}\cap(-K^{+})^{\rho}
=(Kρ(X)+)(Kρ+(X)+)\displaystyle=(K^{\rho}-\mathcal{M}(X^{\ast})^{+})\cap(K^{\rho}+\mathcal{M}(X^{\ast})^{+})
=k1{yMk(X)saxyz for some x,zKρ}.\displaystyle=\coprod_{k\geq 1}\{y\in M_{k}(X^{\ast})^{\text{sa}}\mid x\leq y\leq z\text{ for some }x,z\in K^{\rho}\}.

Hence, if yLρky\in L^{\rho}_{k}, then yy satisfies xyzx\leq y\leq z for some x,zMk(X)+x,z\in M_{k}(X^{\ast})^{+} with x,z1\|x\|,\|z\|\leq 1. By complete α\alpha-normality, this implies yα\|y\|\leq\alpha, so yαKρy\in\alpha K^{\rho}. This proves LραKρL^{\rho}\subseteq\alpha K^{\rho}, so KαLK\subseteq\alpha L.

Hence KαL=αncconv¯(K+(K+))K\subseteq\alpha L=\alpha\overline{\mathrm{ncconv}}(K^{+}\cup(-K^{+})). Using Lemma 4.1, we have

ncconv¯(K+(K+))K+K+.\overline{\mathrm{ncconv}}(K^{+}\cup(-K^{+}))\subseteq K^{+}-K^{+}.

Hence Kα(K+K+)K\subseteq\alpha(K^{+}-K^{+}), and by rescaling every element x(X)sax\in\mathcal{M}(X)^{\text{sa}} can be decomposed as x=yzx=y-z with y,z0y,z\geq 0 and y,zαx\|y\|,\|z\|\leq\alpha\|x\|, and so y+z2αx\|y\|+\|z\|\leq 2\alpha\|x\|. This shows that XX is completely 2α2\alpha-normal. ∎

Remark 5.6.

If HH is a Hilbert space, then B(H)B(H) is completely 11-normal. Consequently, if XX is a matrix ordered operator space which is completely norm and order isomorphic to a subspace of B(H)B(H) (a quasi-operator system), then XX must be α\alpha-normal for some α>0\alpha>0.

Because complete α\alpha-normality is dual to complete α\alpha-generation, [ng2022dual, Theorem 3.9] can be viewed as a partial converse to Remark 5.6. If X=SX=S^{\ast} is the dual of an operator space, then if XX is completely α\alpha-normal, then it is a dual quasi-operator system. Translating the normality condition into a condition on the nc quasistate space gives the following intrinsic characterization of dualizability.

Theorem 5.7.

Let (K,0)(K,0) be a pointed compact nc convex set, with associated operator system S=A(K,0)S=A(K,0). The following are equivalent.

  • (1)

    SS^{\ast} is a dual quasi-operator system.

  • (2)

    SS is completely α\alpha-generated for some α>0\alpha>0.

  • (3)

    SS^{\ast} is completely α\alpha-normal for some α>0\alpha>0.

  • (4)

    There is a constant C>0C>0 such that

    (K+K)+KCK,(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K\subseteq CK,

    where K+KK-\mathbb{R}_{+}K denotes the levelwise Minkowski difference.

  • (5)

    The closed nc convex set (K+K)+K(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K is bounded.

Proof.

The equivalence of (1) and (2) was proved by Ng in [ng2022dual, Theorem 3.9]. Proposition 5.5 shows that (2) and (3) are equivalent. To prove that (3) and (4) are equivalent, we may use Proposition 4.2 to identify n1Mn(S)sa=spanK\coprod_{n\geq 1}M_{n}(S^{\ast})^{\text{sa}}=\operatorname{span}_{\mathbb{R}}K. After doing so, the positive elements in (S)\mathcal{M}(S^{\ast}) correspond to the closed nc convex set +K\mathbb{R}_{+}K, and for d+Knd\in\mathbb{R}_{+}K_{n}, we have dMn(S)=γK(d)\|d\|_{M_{n}(S^{\ast})}=\gamma_{K}(d). Consequently,

(K+K)+K\displaystyle(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K ={dspanK0dx for some xK}\displaystyle=\{d\in\operatorname{span}_{\mathbb{R}}K\mid 0\leq d\leq x\text{ for some }x\in K\}
={dnMn(S)sa0dx for some x>0 in Kn with x1}.\displaystyle=\{d\in\coprod_{n}M_{n}(S^{\ast})^{\text{sa}}\mid 0\leq d\leq x\text{ for some }x>0\text{ in }K_{n}\text{ with }\|x\|\leq 1\}.

Thus (4) holds if and only if

0xy and y1xC,0\leq x\leq y\text{ and }\|y\|\leq 1\implies\|x\|\leq C,

in Mn(S)saM_{n}(S^{\ast})^{\text{sa}} for all nn\in\mathbb{N}. By rescaling, this is equivalent to asserting that

0xyxCy0\leq x\leq y\implies\|x\|\leq C\|y\|

in Mn(S)saM_{n}(S^{\ast})^{\text{sa}}. Then, Proposition 5.4 shows that if (3) holds, then (4) holds with C=αC=\alpha, and if (4) holds, then (3) holds with α=2C+1\alpha=2C+1. Finally, because (K+K)+K(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K is a subset of +K\mathbb{R}_{+}K, on which the matrix norms from SS^{\ast} agree with the Minkowski gauge γK\gamma_{K}, (4) holds if and only if (K+K)+K(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K is bounded by C>0C>0, i.e. if and only if (5) holds. ∎

Note that “bounded” in Theorem 5.7.(5) is in reference to the system of matrix norms on nMn(S)\coprod_{n}M_{n}(S^{\ast}), i.e. uniform boundedness in cb-norm at each level.

Remark 5.8.

The analogous version of Theorem 5.7 holds in the classical case: If (K,0)(K,0) is a pointed compact convex set, then the nonunital function system A(K,0)A(K,0) is α\alpha-generated for some α>0\alpha>0 if and only if (K+K)+K(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K is bounded.

Corollary 5.9.

Let zKLz\in K\subseteq L be compact nc convex sets such that (K,z)(K,z) and (L,z)(L,z) are pointed. If A(L,z)A(L,z) is dualizable, then so is A(K,z)A(K,z).

Proof.

By translating, it suffices to consider this when z=0z=0. This follows by noting that

(K+K)+K(L+L)+L,(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K\subseteq(L-\mathbb{R}_{+}L)\cap\mathbb{R}_{+}L,

and using condition (5) in Theorem 5.7. ∎

In [kennedy2023nonunital, Section 8], quotients of (nonunital) operator systems were defined. There, a quotient of operator systems SS/JS\to S/J corresponds dually to a restriction map A(K,z)A(M,z)A(K,z)\to A(M,z) between pointed compact nc convex sets, where MKM\subseteq K is the annihilator of the kernel JKJ\subseteq K. Applying Corollary 5.9 gives

Corollary 5.10.

If SS is a dualizable operator system, then every quotient of SS is dualizable.

6. Noncommutative simplices

A noncommutative (Choquet) simplex KK is a compact nc convex set such that every point xKx\in K has a unique representing ucp map on the maximal C*-algebra C(K)Cmax(A(K))C(K)\cong C^{\ast}_{\max}(A(K)), which must be the point evaluation δx\delta_{x}. In [kennedy2022noncommutative], the second author and Shamovich characterized nc simplices as corresponding dually to unital C*-systems in the sense of Kirchberg and Wassermann [kirchberg1998c], as follows.

Theorem 6.1.

[kennedy2022noncommutative, Theorems 4.7 and 6.2] Let KK be a compact nc convex set. Then KK is an nc simplex if and only if the bidual A(K)A(K)^{\ast\ast} is unital completely order isomorphic to a C*-algebra. Moreover, if this is the case, the inclusion A(K)A(K)A(K)\hookrightarrow A(K)^{\ast\ast} extends to a \ast-homomorphism C(K)A(K)C(K)\to A(K)^{\ast\ast}, which further extends to a normal conditional expectation of C(K)B(K)C(K)^{\ast\ast}\cong B(K) onto A(K)A(K)^{\ast\ast}.

In fact, if KK is an nc simplex, then we will need to identify the C*-algebra A(K)A(K)^{\ast\ast} as the bidual of the C*-envelope.

Lemma 6.2.

Let KK be a compact nc simplex. Then A(K)A(K)^{\ast\ast} is \ast-isomorphic to the bidual Cmin(A(K))C^{\ast}_{\text{min}}(A(K))^{\ast\ast} via a \ast-isomorphism preserving A(K)A(K).

Proof.

Included in [kennedy2022noncommutative, Theorem 4.7] is the fact that the \ast-homomorphism C(K)A(K)C(K)\to A(K)^{\ast\ast} preserving A(K)A(K) factors through the C*-envelope Cmin(A(K))A(K)C^{\ast}_{\text{min}}(A(K))\to A(K)^{\ast\ast}, still as a \ast-homomorphism. Because A(K)A(K)^{\ast\ast} is a von Neumann algebra, this extends to a normal \ast-homomorphism

π:Cmin(A(K))A(K)\pi:C^{\ast}_{\text{min}}(A(K))^{\ast\ast}\to A(K)^{\ast\ast}

preserving A(K)A(K).

Conversely, if C=C(A(K))A(K)C=C^{\ast}(A(K))\subseteq A(K)^{\ast\ast} is the C*-subalgebra of A(K)A(K)^{\ast\ast} generated by A(K)A(K), then by the universal property of the C*-envelope, there is a \ast-homomorphism CCmin(A(K))C\to C^{\ast}_{\text{min}}(A(K)) preserving A(K)A(K). Upon identifying CA(K)C^{\ast\ast}\hookrightarrow A(K)^{\ast\ast}, we have C=A(K)C^{\ast\ast}=A(K)^{\ast\ast}, because A(K)A(K)^{\ast\ast} is generated as a von Neumann algebra by A(K)A(K). So, double-dualizing the homomorphism CCmin(A(K))C\to C^{\ast}_{\text{min}}(A(K)) gives a normal \ast-homomorphism

σ:A(K)Cmin(A(K))\sigma:A(K)^{\ast\ast}\to C^{\ast}_{\text{min}}(A(K))^{\ast\ast}

that preserves A(K)A(K). Since π\pi and σ\sigma are normal \ast-homomorphisms, and the copies of A(K)A(K) generated A(K)A(K)^{\ast\ast} and Cmin(A(K))C^{\ast}_{\text{min}}(A(K))^{\ast\ast} as von Neumann algebras, it follows that π\pi and σ\sigma are mutual inverses and so A(K)Cmin(A(K))A(K)^{\ast\ast}\cong C^{\ast}_{\text{min}}(A(K))^{\ast\ast} naturally. ∎

An nc Bauer simplex has the additional property that the nc extreme points K\partial K are a closed set in the topology induced from the spectrum of C(K)C(K), and KK is an nc Bauer simplex if and only if A(K)A(K) is itself completely order isomorphic to a C*-algebra [kennedy2022noncommutative, Theorem 10.5]. The second author, Kim, and the third author obtained a noncommutative extension of this result. In [kennedy2023nonunital, Theorem 10.9], they proved that the nonunital system A(K,z)A(K,z) is completely order and norm isomorphic to a C*-algebra if and only if KK is an nc Bauer simplex and zK1z\in K_{1} is an nc extreme point of KK. The corresponding characterization for nc (possibly non-Bauer) simplices is as follows.

Theorem 6.3.

Let (K,z)(K,z) be a pointed compact nc convex set. The operator system A(K,z)A(K,z)^{\ast\ast} is completely isometrically order isomorphic to a C*-algebra if and only if KK is an nc simplex and zKz\in\partial K.

Proof.

The embedding of A(K,z)A(K,z) into its partial unitization A(K,z)=A(K)A(K,z)^{\sharp}=A(K) double-dualizes to a completely isometric order injection of A(K,z)A(K,z)^{\ast\ast} into A(K)A(K)^{\ast\ast} of codimension one. And, A(K)A(K)^{\ast\ast} is naturally viewed as a weak-\ast closed subspace of the C*-algebra B(K)C(K)B(K)\cong C(K)^{\ast\ast} of bounded nc functions on KK. With this identification, we have A(K)=A(K,z)+1A(K)A(K)^{\ast\ast}=A(K,z)^{\ast\ast}+\mathbb{C}1_{A(K)}, from which it follows that A(K)A(K)^{\ast\ast} coincides naturally with the partial unitization of A(K,z)A(K,z)^{\ast\ast}.

Let KK^{\ast\ast} denote the nc state space of A(K)A(K)^{\ast\ast}. Then, KK embeds via an affine nc homeomorphism onto a subset of KK^{\ast\ast}, and we will denote the embedding KKK\hookrightarrow K^{\ast\ast} by xxx\mapsto x^{\ast\ast}. Again, through the identification A(K,z)A(K)B(K)A(K,z)^{\ast\ast}\subseteq A(K)^{\ast\ast}\subseteq B(K), it is straightforward to see that

A(K,z)=A(K,z).A(K,z)^{\ast\ast}=A(K^{\ast\ast},z^{\ast\ast}).

Indeed, the inclusion A(K,z)A(K,z)A(K,z)^{\ast\ast}\subseteq A(K^{\ast\ast},z^{\ast\ast}) is immediate. Conversely, an affine nc function aA(K)a\in A(K)^{\ast\ast} which vanishes on zz^{\ast\ast} can be approximated weak-\ast by functions in A(K)A(K), and–by adding a multiple of 1A(K)=1A(K)1_{A(K)}=1_{A(K)^{\ast\ast}}, by functions which vanish at zz.

Because KK is an nc simplex, A(K)A(K)^{\ast\ast} is isomorphic to a C*-algebra, and so its nc state space KK^{\ast\ast} is an nc Bauer simplex. So, it suffices to prove that zz^{\ast\ast} is an nc extreme point in KK^{\ast\ast}. In the nc Bauer simplex KK^{\ast\ast} the nc extreme points are exactly irreducible representations of the C*-algebra A(K)A(K)^{\ast\ast}.

Since zKz\in\partial K, the point zz is a boundary representation of A(K)A(K), which extends uniquely to an irreducible representation of Cmin(A(K))C^{\ast}_{\text{min}}(A(K)). This extends further to a normal irreducible representation of Cmin(A(K))C^{\ast}_{\text{min}}(A(K))^{\ast\ast}. By Lemma 6.2, we have Cmin(A(K))A(K)C^{\ast}_{\text{min}}(A(K))^{\ast\ast}\cong A(K)^{\ast\ast} naturally, and zz^{\ast\ast} is the unique normal extension of zz. Therefore zz^{\ast\ast} is an irreducible representation and so is nc extreme in A(K)A(K)^{\ast\ast}.

So, KK^{\ast\ast} is an nc Bauer simplex with nc extreme point zz^{\ast\ast}, so [kennedy2023nonunital, Theorem 10.9] implies that

A(K,z)=A(K,z)A(K,z)^{\ast\ast}=A(K^{\ast\ast},z^{\ast\ast})

is isomorphic to a C*-algebra. ∎

Remark 6.4.

Note that if KK is an nc simplex, but zz is not nc extreme, then A(K,z)A(K,z)

Since all C*-algebras are dualizable, we can conclude that all (possibly nonunital) C*-systems are dualizable.

Corollary 6.5.

If KK is an nc simplex, and zK1(K)z\in K_{1}\cap(\partial K) is an nc extreme point, then the operator system A(K,z)A(K,z) is dualizable.

Proof.

By Theorem 6.3, the double dual A(K,z)A(K,z)^{\ast\ast} is a C*-algebra, and therefore a dualizable operator system. Therefore, the triple dual A(K,z)A(K,z)^{\ast\ast\ast} is an operator system. Since the natural map

A(K,z)A(K,z)A(K,z)^{\ast}\to A(K,z)^{\ast\ast\ast}

is a completely isometric order injection, the dual A(K,z)A(K,z)^{\ast} embeds into an operator system and is therefore itself an operator system. ∎

Therefore, upon translating (K,z)(K,z) to (Kz,0)(K-z,0), if KK is an nc simplex containing 0 as an nc extreme point, the nc geometric conditions in Corollary 5.1 and Theorem 5.7 hold for KK.

7. Dualizability for function systems

By a function system, we mean a selfadjoint subspace of a commutative C*-algebra C(X)C(X), for some compact Hausdorff space XX. Classical Kadison duality [kadison1951representation] asserts that function systems are categorically dual to to (ordinary) compact convex sets. What follows is a commutative version of Theorem 5.7.(1)-(3). These results are known already as folklore, but we include proofs for completeness and to contrast the situation in Section 8. C.K. Ng gives a more thorough discussion of what is known for function systems in [ng2022dual, Appendix A.2].

Proposition 7.1.

Let SS be a (possibly nonunital) function system. The following are equivalent.

  • (1)

    SS is positively generated, meaning S=S+S+S=S^{+}-S^{+}.

  • (2)

    SS is α\alpha-generated for some α>0\alpha>0.

  • (3)

    The dual SS^{\ast} is order and norm isomorphic to a function system.

Moreover, the isomorphism in (3) can be chosen to be a homeomorphism from the weak-\ast topology to the topology of pointwise convergence on bounded sets.

Proof.

The implication (2) \implies (1) is immediate. If condition (3) holds, then SS^{\ast} is norm isomorphic to a function system, which is 11-normal, and so SS^{\ast} is α\alpha-normal for some α>0\alpha>0. Therefore SS is α\alpha^{\prime} generated for any α>0\alpha>0. (See [asimow2014convexity, Theorem 2.1.4], which is a classical result corresponding to part of Proposition 5.5.)


If SS is positively generated, then it is a consequence of the Baire Category Theorem that SS is α\alpha-generated, as in [asimow2014convexity, Theorem 2.1.2]. Sketching the proof, let

B=conv(S+1(S+1)),B=\operatorname{conv}(S^{+}_{1}\cup(-S^{+}_{1})),

where S1+=B1(S+)S_{1}^{+}=B_{1}(S^{+}) is the unit ball of S+S^{+}. Then

Ssaα1nB¯,S^{\text{sa}}\subseteq\bigcup_{\alpha\geq 1}n\overline{B},

and so some nB¯n\overline{B} has interior. By shifting and rescaling, we can arrange that

S1sanB¯S_{1}^{\text{sa}}\subseteq n\overline{B}

for some n>1n>1. Then, a series argument shows that B¯(1+ϵ)B\overline{B}\subseteq(1+\epsilon)B for any ϵ>0\epsilon>0, and so S1san(1+ϵ)BS_{1}^{\text{sa}}\subseteq n(1+\epsilon)B. That is, SS is α\alpha-generated for any α>n\alpha>n. Thus, (1) implies (2).


Now, suppose that SS is α\alpha-generated, so that S1αconv(S1+(S1+))S_{1}\subseteq\alpha\operatorname{conv}(S_{1}^{+}\cup(-S_{1}^{+})). Let J:SSJ:S\to S^{\ast\ast} be the natural embedding of SS into its double dual. Let XSX\subseteq S^{\ast\ast} be the closure of (the image of) S1+S_{1}^{+} in the weak-\ast topology of SS^{\ast\ast}. Consider the linear map

ρ:SC(X)\rho:S^{\ast}\to C(X)

which satisfies ρ(f)(J(a))=f(a)\rho(f)(J(a))=f(a) for aS1+a\in S_{1}^{+}. By definition ρ\rho is an order isomorphism onto its range. Since S1+S1S_{1}^{+}\subseteq S_{1}, the map ρ\rho is contractive. Given ff in SS^{\ast}, and a selfadjoint aSsaa\in S^{\text{sa}}, we can find b,cS+b,c\in S^{+} with a=bca=b-c and b+cαa\|b\|+\|c\|\leq\alpha\|a\|. Therefore, b/αab/\alpha\|a\| and c/αac/\alpha\|a\| are in S1+S_{1}^{+}, and so

f(a)\displaystyle\|f(a)\| f(b)+f(c)\displaystyle\leq\|f(b)\|+\|f(c)\|
=αa(f(bαa)+f(cαa))\displaystyle=\alpha\|a\|\left(\left\|f\left(\frac{b}{\alpha\|a\|}\right)\right\|+\left\|f\left(\frac{c}{\alpha\|a\|}\right)\right\|\right)
=αa(ρ(f)+ρ(f))=2αaρ(f).\displaystyle=\alpha\|a\|\left(\|\rho(f)\|+\|\rho(f)\|\right)=2\alpha\|a\|\|\rho(f)\|.

This proves that ρ(f)f/2α\|\rho(f)\|\geq\|f\|/2\alpha, and so ρ\rho is bounded below. Therefore ρ\rho is an order and norm isomorphism onto a function system. Since XX has the weak-\ast topology, it also follows that ρ\rho is a weak-\ast to pointwise homeomorphism on bounded sets. ∎

Remark 7.2.

Item (3) in Proposition 7.1 cannot be extended to say completely order and norm isomorphic, even if we replace (1) or (2) with the stronger hypothesis that SS is completely α\alpha-generated and so dualizable. To see why, if SC(X)S\subseteq C(X), then SS has its minimal operator space structure S=min(S)S=\min(S), in the sense of [effros2022theory, Proposition 3.3.1]. Therefore, as an operator space S=max(S)S^{\ast}=\max(S^{\ast}). Using the map ρ\rho in the proof of Proposition 7.1, ρ(S)\rho(S) is a function system, and so ρ(S)=min(ρ(S))\rho(S)=\min(\rho(S)) as operator spaces. Therefore, if ρ\rho was completely bounded below, it would induce a complete norm isomorphism between max(S)\max(S) and min(S)\min(S). If SS is infinite dimensional, then the maximal and minimal operator space structures are not completely equivalent [paulsen2002completely, Theorem 14.3], and so ρ\rho cannot be completely bounded below.

By [paulsen2002completely, Theorem 3.9], the map ρ\rho is completely positive. But, by definition of ρ\rho, the map ρ\rho is a complete order isomorphism if and only if every positive map SMnS\to M_{n} is completely positive, but this is not true for even finite dimensional operator systems SS.

So, even if SS is a function system that is a dualizable operator system, its dual SS^{\ast} is typically not completely order and norm isomorphic to a function system, and never can be when SS is infinite dimensional. We don’t know whether positive generation of a function system SS is enough to guarantee completely bounded positive generation and so dualizability. We leave this as an open question.

Question 7.3.

If SC(X)S\subseteq C(X) is a positively generated function system, is SS completely α\alpha-generated for some α>0\alpha>0?

If so, then SS is dualizable if and only if it is positively generated. In Proposition 8.4 below, we show that positive generation actually guarantees positive generation at all matrix levels. If Question 7.3 has a negative answer, then by Proposition 8.4 below there is a function system SS for which each Mn(S)M_{n}(S) is αn\alpha_{n}-generated, but the sequence (αn)(\alpha_{n}) cannot be chosen to be bounded.

In Example 8.6 below, we give a matrix ordered operator space which is positively generated , but not completely α\alpha-generated for any α>0\alpha>0. We do not know a function system with this property.

8. Positive generation

In Proposition 7.1 above, we showed that for function systems, positive generation and bounded positive generation coincide. In this section, we discuss the noncommutative situation. First, we show that an operator system SS has complete positive generation, meaning Mn(S)sa=Mn(S)+Mn(S)+M_{n}(S)^{\text{sa}}=M_{n}(S)^{+}-M_{n}(S)^{+} for all n1n\geq 1, if and only if SS is positively generated at the first level. In contrast to the classical situation, complete positive generation need not imply complete α\alpha-generation. In Example 8.6, we give an example of a matrix ordered operator space which is positively generated but not completely α\alpha-generated for any α>0\alpha>0.

One might also consider the following weaker property. Call an ordered Banach space EE approximately positively generated if E+E+E^{+}-E^{+} is dense in EE. Note that even though the postiive cone E+E^{+} is closed, it need not be the case that E+E+E^{+}-E^{+} is closed, even when EE is an operator space, as the following example shows.

Example 8.1.

Let S=C([0,1])S=C([0,1]), and define S+S^{+} to be the closed cone of functions which are both positive and convex. Then S+S+S^{+}-S^{+} is dense in S=C([0,1])S=C([0,1]), because it contains all C2C^{2} functions, but S+S+SS^{+}-S^{+}\neq S, because the convex functions in S+S^{+} are automatically differentiable almost everywhere on the interior (0,1)(0,1). So, SS is an ordered Banach space which is approximately positively generated, but not positively generated. In fact, SS is an operator system. Indeed, if we let

K={φSφ1 and φ(S+)[0,)}K=\{\varphi\in S^{\ast}\mid\|\varphi\|\leq 1\text{ and }\varphi(S^{+})\subseteq[0,\infty)\}

be the classical quasistate space of KK, then since every probability measure on [0,1][0,1] lies in KK, the natural map

SA(K)S\to A(K)

into the continuous affine functions on KK is isometric and order isomorphic. That is, SS is isometrically order isomorphic to a nonunital function system, and so inherits an operator system structure.

There are many examples of the same kind as Example 8.1. It suffices to take any function system SS, and equip it with a new closed positive cone PS+P\subseteq S^{+} for which PPP-P is not closed. In a private correspondence, Ken Davidson suggested another example in which S=c0S=\mathbb{C}\oplus c_{0} is equipped with the new positive cone

P={(t,(xn)n1)c0t0,(xn)n10, and n=1xnt}.P=\Big{\{}(t,(x_{n})_{n\geq 1})\in\mathbb{C}\oplus c_{0}\mid t\geq 0,(x_{n})_{n\geq 1}\geq 0,\text{ and }\sum_{n=1}^{\infty}x_{n}\leq t\Big{\}}.

Here, again PPP-P is dense and not closed in SS.

Proposition 8.2.

Let SS be an operator system with quasistate space KSK\subseteq S^{\ast}. Then SS is approximately positively generated if and only if S+S^{+} separates points in KK.

Proof.

If SS is densely spanned by its positives, then the positives must separate points in KK. Conversely, suppose that SS is not positively generated. Then there exists an element xSsa(S+S+)¯x\in S^{\text{sa}}\setminus\overline{(S^{+}-S^{+})}. By the Hahn-Banach Separation Theorem, there is a self-adjoint linear functional φS\varphi\in S^{\ast} so that for all yS+S+y\in S^{+}-S^{+} we have

φ(x)<φ(y).\displaystyle\varphi(x)<\varphi(y).

But since S+S+S^{+}-S^{+} is a real vector space, this implies that φ\varphi is identically zero on S+S+S^{+}-S^{+}. Moreover, by the Hahn-Jordan decomposition theorem there are positive functionals φ+,φEd\varphi^{+},\varphi^{-}\in E^{d} with φ=φ+φ\varphi=\varphi^{+}-\varphi^{-}. Since φ(x)<0\varphi(x)<0, the functionals φ+\varphi^{+} and φ\varphi^{-} are necessarily distinct, but they are equal on S+S+S^{+}-S^{+} and hence on S+S^{+}. Normalizing φ±\varphi^{\pm} to obtain quasistates shows that S+S^{+} does not separate quasistates. ∎

Remark 8.3.

The Hahn-Jordan decomposition theorem ensures that, as an ordered vector space, the dual space SS^{\ast} is always positively generated.

By the following result, if SS is positively generated then so are each of its matrix levels Mn(S)M_{n}(S). By , Using Kadison Duality, each level Mn(S)M_{n}(S) can itself be viewed as a function system by forgetting the rest of the matrix order, and so Proposition 7.1 implies that each Mn(S)M_{n}(S) is αn\alpha_{n}-generated for some αn>0\alpha_{n}>0. However, in order for SS to be dualizable, we would need the sequence (αn)(\alpha_{n}) to be bounded.

Proposition 8.4.

If SS is positively generated, then so is Mn(S)M_{n}(S) for each nn. That is, a positively generated operator system is automatically completely positively generated.

Before proving this, we will need a technical lemma which proves a much stronger statement in the finite dimensional setting.

Lemma 8.5.

If SS is a finite dimensional and positively generated operator system, then it contains a matrix order unit.

Proof.

Since SS is positively generated, then it admits a basis B={p1,,pm}B=\{p_{1},\ldots,p_{m}\} consisting of positive elements. We claim that e:=i=1mpie:=\sum_{i=1}^{m}p_{i} is an order unit. For any xx in SsaS^{\text{sa}}, we can write xx uniquely as a real linear combination

x=i=1mαipi,x=\sum_{i=1}^{m}\alpha_{i}p_{i},

and we define λx:=max{1,|α1|,,|αm|}\lambda_{x}:=\max\{1,|\alpha_{1}|,\ldots,|\alpha_{m}|\}. It is clear that λxe±x\lambda_{x}e\pm x are positive in SS, so ee is an order unit.

Next we let n0n\geq 0 and show that en:=eIne_{n}:=e\otimes I_{n} is an order unit for Mn(S)M_{n}(S), so fix an X=(xij)i,j=1nMn(S)saX=(x_{ij})_{i,j=1}^{n}\in M_{n}(S)^{\text{sa}}. Since EE is positively generated, for every iji\leq j we can decompose the corresponding entries of XX as

xij=xij+xij+i(Im xij+Im xij).\displaystyle x_{ij}=\real x_{ij}^{+}-\real x_{ij}^{-}+i(\text{Im }x_{ij}^{+}-\text{Im }x_{ij}^{-}).

To find a large enough coefficient of ene_{n} to dominate XX, we let

λX:=λd+λ+λIm .\displaystyle\lambda_{X}:=\lambda_{d}+\lambda+\lambda_{\text{Im }}.

Where λd:=max{λxii}i=1n\lambda_{d}:=\max\{\lambda_{x_{ii}}\}_{i=1}^{n}, λ:=i<jλxij++xij\lambda:=\sum_{i<j}\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}}, and λIm :=i<jλIm xij++Im xij\lambda_{\text{Im }}:=\sum_{i<j}\lambda_{\text{Im }x_{ij}^{+}+\text{Im }x_{ij}^{-}}. Note that it makes sense to write xii±x_{ii}^{\pm} since the xiix_{ii} must all be self-adjoint, as they lie on the diagonal of X=XX=X^{*}.

Fix a concrete representation SB(H)S\hookrightarrow B(H) of SS as a norm closed and \ast-closed subspace of the bounded operators on a Hilbert space. We’ll show that λXen+X0\lambda_{X}e_{n}+X\geq 0 concretely using inner products. Take an arbitrary vector a=(ai)i=1nHn=i=1nHa=(a_{i})_{i=1}^{n}\in H^{n}=\bigoplus_{i=1}^{n}H, and compute

(λXen+X)a,a=\displaystyle\langle(\lambda_{X}e_{n}+X)a,a\rangle= λXena,a+Xa,a\displaystyle\lambda_{X}\langle e_{n}a,a\rangle+\langle Xa,a\rangle
=\displaystyle= λXi=1neai,ai+i=1nxiiai,ai+i<jxijaj,ai+xjiai,aj\displaystyle\lambda_{X}\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+\sum_{i=1}^{n}\langle x_{ii}a_{i},a_{i}\rangle+\sum_{i<j}\langle x_{ij}a_{j},a_{i}\rangle+\langle x_{ji}a_{i},a_{j}\rangle
=\displaystyle= λXi=1neai,ai+i=1nxiiai,ai+i<j2xijaj,ai\displaystyle\lambda_{X}\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+\sum_{i=1}^{n}\langle x_{ii}a_{i},a_{i}\rangle+\sum_{i<j}2\real\langle x_{ij}a_{j},a_{i}\rangle
=\displaystyle= (λdi=1neai,ai+i=1nxiiai,ai)\displaystyle\left(\lambda_{d}\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+\sum_{i=1}^{n}\langle x_{ii}a_{i},a_{i}\rangle\right)
+\displaystyle+ ((λ+λIm )i=1neai,ai+i<j2xijaj,ai)\displaystyle\left((\lambda+\lambda_{\text{Im }})\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+\sum_{i<j}2\real\langle x_{ij}a_{j},a_{i}\rangle\right)
=\displaystyle= (λdi=1neai,ai+i=1nxiiai,ai)\displaystyle\left(\lambda_{d}\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+\sum_{i=1}^{n}\langle x_{ii}a_{i},a_{i}\rangle\right)
+\displaystyle+ (λi=1neai,ai+2i<jxijaj,ai)\displaystyle\left(\lambda\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+2\sum_{i<j}\real\langle\real x_{ij}a_{j},a_{i}\rangle\right)
+\displaystyle+ (λIm i=1neai,ai2i<jIm Im xijaj,ai).\displaystyle\left(\lambda_{\text{Im }}\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle-2\sum_{i<j}\text{Im }\langle\text{Im }x_{ij}a_{j},a_{i}\rangle\right).

For the remainder of the proof, we will show that each of the three terms above is non-negative. Starting with the first term,

λdi=1neai,ai+i=1nxiiai,ai=\displaystyle\lambda_{d}\sum_{i=1}^{n}\langle ea_{i},a_{i}\rangle+\sum_{i=1}^{n}\langle x_{ii}a_{i},a_{i}\rangle= i=1n(λde+xii)ai,ai\displaystyle\sum_{i=1}^{n}\langle(\lambda_{d}e+x_{ii})a_{i},a_{i}\rangle
\displaystyle\geq i=1n(λxiie+xii)ai,ai\displaystyle\sum_{i=1}^{n}\langle(\lambda_{x_{ii}}e+x_{ii})a_{i},a_{i}\rangle
\displaystyle\geq 0,\displaystyle\ 0,

where the last inequality follows from the first paragraph of the proof.

To prove that the second term is non-negative, note

λk=1neak,ak+2i<jxijaj,ai\displaystyle\lambda\sum_{k=1}^{n}\langle ea_{k},a_{k}\rangle+2\sum_{i<j}\real\langle\real x_{ij}a_{j},a_{i}\rangle
=i<j(λxij++xij)k=1neak,ak+2i<jxijaj,ai\displaystyle\quad=\sum_{i<j}(\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}})\sum_{k=1}^{n}\langle ea_{k},a_{k}\rangle+2\sum_{i<j}\real\langle\real x_{ij}a_{j},a_{i}\rangle
=i<j(λxij++xij)k=1neak,ak+2xijaj,ai.\displaystyle\quad=\sum_{i<j}(\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}})\sum_{k=1}^{n}\langle ea_{k},a_{k}\rangle+2\real\langle\real x_{ij}a_{j},a_{i}\rangle.

We now show that for each pair i<ji<j, the corresponding summand is non-negative:

(λxij++xij)k=1neak,ak+2xijaj,ai\displaystyle(\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}})\sum_{k=1}^{n}\langle ea_{k},a_{k}\rangle+2\real\langle\real x_{ij}a_{j},a_{i}\rangle
=(λxij++xij)k=1neak,ak+2(xij+xij)aj,ai\displaystyle\quad=(\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}})\sum_{k=1}^{n}\langle ea_{k},a_{k}\rangle+2\real\langle(\real x_{ij}^{+}-\real x_{ij}^{-})a_{j},a_{i}\rangle
(λxij++xij)eai,ai+(λxij++xij)eaj,aj+2(xij+xij)aj,ai\displaystyle\quad\geq(\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}})\langle ea_{i},a_{i}\rangle+(\lambda_{\real x_{ij}^{+}+\real x_{ij}^{-}})\langle ea_{j},a_{j}\rangle+2\real\langle(\real x_{ij}^{+}-\real x_{ij}^{-})a_{j},a_{i}\rangle
(xij+ai,ai+xij+aj,aj+2xij+aj,ai)\displaystyle\quad\geq\left(\langle\real x_{ij}^{+}a_{i},a_{i}\rangle+\langle\real x_{ij}^{+}a_{j},a_{j}\rangle+2\real\langle\real x_{ij}^{+}a_{j},a_{i}\rangle\right)
+(xijai,ai+xijaj,aj2xijaj,ai)\displaystyle\quad\quad+\left(\langle\real x_{ij}^{-}a_{i},a_{i}\rangle+\langle\real x_{ij}^{-}a_{j},a_{j}\rangle-2\real\langle\real x_{ij}^{-}a_{j},a_{i}\rangle\right)
=xij+(ai+aj),ai+aj+xij(aiaj),aiaj\displaystyle\quad=\langle\real x_{ij}^{+}(a_{i}+a_{j}),a_{i}+a_{j}\rangle+\langle\real x_{ij}^{-}(a_{i}-a_{j}),a_{i}-a_{j}\rangle
0.\displaystyle\quad\geq 0.

The last inequality follows since each xij±\real x_{ij}^{\pm} is a positive operator. The proof that the third term is non-negative is similar. ∎

We now prove Proposition 8.4

Proof of Proposition 8.4.

To show Mn(S)M_{n}(S) is positively generated, fix X=(xij)i,j=1nMn(S)saX=(x_{ij})_{i,j=1}^{n}\in M_{n}(S)^{\text{sa}}. Since SS is positively generated, each xijx_{ij} can be written as a linear combination of four positives xij+\real x_{ij}^{+}, xij\real x_{ij}^{-}, Im xij+\text{Im }x_{ij}^{+}, and Im xij\text{Im }x_{ij}^{-}. Let SXS_{X} denote the linear span of these positives, as ii and jj range from 11 to nn. Since SXS_{X} is a finite dimensional operator system, by the previous lemma there is a matrix order unit eXSXe_{X}\in S_{X} and in particular there is a constant λ>0\lambda>0 so that both λ1neX±X0\lambda 1_{n}\otimes e_{X}\pm X\geq 0. Since X=(λ1neX+X)/2(λ1neXX)/2X=(\lambda 1_{n}\otimes e_{X}+X)/2-(\lambda 1_{n}\otimes e_{X}-X)/2 and all entries are ultimately in SS, this shows Mn(S)M_{n}(S) is positively generated. ∎

So, complete positive generation coincides with positive generation at the first level. However, the following example shows that for matrix ordered operator spaces, positive generation at all matrix levels does not imply complete α\alpha-generation for any α\alpha. We do not know if this example is an operator system.

Example 8.6.

Any Banach space EE has a unique maximal and minimal system of LL^{\infty}-matrix norms which give EE an operator space structure and restrict to the norm on EE at the first matrix level. We denote the resultant operator spaces by max(E)\max(E) and min(E)\min(E), respectively. There are natural operator space dualities max(E)=min(E)\max(E)^{\ast}=\min(E^{\ast}) and min(E)=max(E)\min(E)^{\ast}=\max(E)^{\ast} [effros2022theory, Section 3.3].

We will consider the Banach space 1\ell^{1} and its dual \ell^{\infty}. Because \ell^{\infty} is a commutative CC^{\ast}-algebra, we have =min()\ell^{\infty}=\min(\ell^{\infty}) [effros2022theory, Proposition 3.3.1]. The embedding 1()\ell^{1}\subseteq(\ell^{\infty})^{\ast} gives a matrix ordered operator space structure on 1\ell^{1}, which coincides with the max norm 1=max(1)\ell^{1}=\max(\ell^{1}). Using the natural linear identifications

Mn()=(,Mn)andMn(1)=1(,Mn),M_{n}(\ell^{\infty})=\ell^{\infty}(\mathbb{N},M_{n})\quad\text{and}\quad M_{n}(\ell^{1})=\ell^{1}(\mathbb{N},M_{n}),

the resultant positive cones in \ell^{\infty} and 1\ell^{1} consist of those sequences of matrices which are positive in each entry.

We will consider the minimal operator space min(1)\min(\ell^{1}) equipped with the same matrix ordering as 1=max(1)\ell^{1}=\max(\ell^{1}). Because the matrix cones Mn(1)+=1(,Mn+)M_{n}(\ell^{1})^{+}=\ell^{1}(\mathbb{N},M_{n}^{+}) are closed in the topology of pointwise weak-\ast convergence, which is weaker than the topology induced by either the minimal or maximal norms on Mn(1)M_{n}(\ell^{1}), the matrix cones Mn(1)+M_{n}(\ell^{1})^{+} are closed in the minimal norm topology. Thus min(1)\min(\ell^{1}) has the structure of a matrix ordered operator space. Because MnM_{n} is 11-generated, it follows that each Mn(min(1))=1(,Mn)M_{n}(\min(\ell^{1}))=\ell^{1}(\mathbb{N},M_{n}) is positively generated, so min(1)\min(\ell^{1}) is completely positively generated.

However, we will show that min(1)\min(\ell^{1}) is not completely α\alpha-generated for any α>0\alpha>0. We will do so using Proposition 5.5, by proving the dual matrix ordered operator space min(1)=max()\min(\ell^{1})^{\ast}=\max(\ell^{\infty}) (equipped with the usual matrix ordering on \ell^{\infty}) is not completely α\alpha-normal for any α>0\alpha>0. Since \ell^{\infty} is infinite dimensional, the minimal and maximal matrix norms on \ell^{\infty} are not completely equivalent [paulsen2002completely, Theorem 14.3]. Thus there is a sequence xkMnk()x_{k}\in M_{n_{k}}(\ell^{\infty}) for which

xkmin1andxkmaxk.\|x_{k}\|_{\text{min}}\leq 1\quad\text{and}\quad\|x_{k}\|_{\text{max}}\geq k.

In the C*-algebras Mnk()M_{n_{k}}(\ell^{\infty}), we can write each xkx_{k} as a linear combination

xk=(Rexk)+(Rexk)+i(Imxk)+i(Imxk)x_{k}=(\operatorname{Re}x_{k})^{+}-(\operatorname{Re}x_{k})^{-}+i(\operatorname{Im}x_{k})^{+}-i(\operatorname{Im}x_{k})^{-}

of positive elements (Rexk)±,(Imxk)±(\operatorname{Re}x_{k})^{\pm},(\operatorname{Im}x_{k})^{\pm} of min-norm at most 11. Since xkmax>k\|x_{k}\|_{\text{max}}>k, by suitably choosing yk{(Rexk)±,(Imxk)±}y_{k}\in\{(\operatorname{Re}x_{k})^{\pm},(\operatorname{Im}x_{k})^{\pm}\}, we can obtain a sequence of positive elements ykMnk()+y_{k}\in M_{n_{k}}(\ell^{\infty})^{+} with

ykmin1andykmax>k/4.\|y_{k}\|_{\text{min}}\leq 1\quad\text{and}\quad\|y_{k}\|_{\text{max}}>k/4.

Since the minimal norm on Mnk()M_{n_{k}}(\ell^{\infty}) is just the usual C*-algebra norm, we have 0yk1Mnk()0\leq y_{k}\leq 1_{M_{n_{k}}(\ell^{\infty})}. Because the maximal norms satisfy the LL^{\infty}-matrix norm identity, we have 1Mnk()max=1\|1_{M_{n_{k}}(\ell^{\infty})}\|_{\text{max}}=1. Thus

0yk1Mnk(),1Mnk()max1,andykmax>k/40\leq y_{k}\leq 1_{M_{n_{k}}(\ell^{\infty})},\quad\|1_{M_{n_{k}}(\ell^{\infty})}\|_{\text{max}}\leq 1,\quad\text{and}\quad\|y_{k}\|_{\text{max}}>k/4

for all kk\in\mathbb{N}. So, \ell^{\infty} is not completely k/4k/4-normal, and taking kk\to\infty shows that \ell^{\infty} cannot be completely α\alpha-normal for any α>0\alpha>0.

Example 8.6 is a minimal example of this kind. One cannot restrict to the finite dimensional spaces 1d\ell^{1}_{d} and d=(1d)\ell^{\infty}_{d}=(\ell^{1}_{d})^{\ast} because the maximal and minimal norms on a finite dimensional Banach space are completely equivalent [paulsen2002completely, Theorem 14.3], and so max(1d)min(1d)\max(\ell^{1}_{d})\cong\min(\ell^{1}_{d}) is a dualizable quasi-operator system.

9. Permanence properties

If K=n1KnK=\coprod_{n\geq 1}K_{n} and L=n1LnL=\coprod_{n\geq 1}L_{n} are compact nc convex sets, we denote by

K×L:=n1Kn×LnK\times L:=\coprod_{n\geq 1}K_{n}\times L_{n}

their levelwise cartesian product. In [humeniuk2021jensen], it was shown that A(K×L)A(K\times L) is the categorical coproduct of the unital operator systems A(K)A(K) and A(L)A(L) in the category of unital operator systems with ucp maps as morphisms. The following result will let us assert a similar result in the pointed context, for nonunital operator systems.

Proposition 9.1.

Let (K,z)(K,z) and (L,w)(L,w) be pointed compact nc convex sets. Then (K×L,(z,w))(K\times L,(z,w)) is pointed, and there is a vector space isomorphism

A(K×L,(z,w))A(K,z)A(L,w).A(K\times L,(z,w))\cong A(K,z)\oplus A(L,w).
Proof.

We will prove the result in the special case when z=0z=0 and w=0w=0 in the ambient spaces containing KK and LL. The general case follows by translation. Define a linear map A(K,z)A(L,w)A(K×L,(z,w))A(K,z)\oplus A(L,w)\to A(K\times L,(z,w)) by (a,b)ab(a,b)\mapsto a\oplus b, where (ab)(x,y):=a(x)+b(y)(a\oplus b)(x,y):=a(x)+b(y) for xKx\in K, yLy\in L. Since a(z)=0=b(w)a(z)=0=b(w), it is easy to see that this map is injective. Given cA(K×L,(0,0))c\in A(K\times L,(0,0)), let a(x)=c(x,0)a(x)=c(x,0) and b(y)=c(0,y)b(y)=c(0,y) for xKx\in K, yLy\in L. Then since c(0,0)=0c(0,0)=0,

c(x,y)\displaystyle c(x,y) =2c(x2,y2)\displaystyle=2c\left(\frac{x}{2},\frac{y}{2}\right)
=2(c(x,0)2+c(0,y)2)\displaystyle=2\left(\frac{c(x,0)}{2}+\frac{c(0,y)}{2}\right)
=a(x)+b(y)=(ab)(x,y).\displaystyle=a(x)+b(y)=(a\oplus b)(x,y).

This proves that A(K,0)A(L,0)A(K×L,(0,0))A(K,0)\oplus A(L,0)\to A(K\times L,(0,0)) is a linear isomorphism.

Now, it will follow from this isomorphism that (K×L,(z,w))(K\times L,(z,w)) is pointed. Let ρ:A(K×L,(z,w))Mn\rho:A(K\times L,(z,w))\to M_{n} be any nc quasistate. Then

φ(a)=ρ(a0)andψ(b)=ρ(0b)\varphi(a)=\rho(a\oplus 0)\quad\text{and}\quad\psi(b)=\rho(0\oplus b)

define nc quasistates on A(K,0)A(K,0) and A(L,0)A(L,0), respectively. Because (K,0)(K,0) and (L,0)(L,0) are pointed, all nc quasistates are point evaluations, so we have φ(a)=a(x)\varphi(a)=a(x) and φ(b)=b(y)\varphi(b)=b(y) for some (x,y)(K×L)n(x,y)\in(K\times L)_{n} and all aA(K,0)a\in A(K,0), bA(L,0)b\in A(L,0). From linearity, it follows that ρ\rho is just point evaluation at (x,y)(x,y), so (K×L,(0,0))(K\times L,(0,0)) is pointed. ∎

Definition 9.2.

Let SS and TT be operator systems with respective nc quasistate spaces (K,0)(K,0) and (L,0)(L,0). We define the operator system coproduct to be the vector space STS\oplus T equipped with the operator system structure such that

STA(K,0)A(L,0)A(K×L,(0,0))S\oplus T\cong A(K,0)\oplus A(L,0)\cong A(K\times L,(0,0))

is a completely isometric complete order isomorphism.

Explicitly, the matrix norms on STS\oplus T satisfy

(x,y)Mn(ST)=sup{φn(x)+ψn(y)φK,ψL}\|(x,y)\|_{M_{n}(S\oplus T)}=\sup\{\|\varphi_{n}(x)+\psi_{n}(y)\|\mid\varphi\in K,\psi\in L\}

for (x,y)Mn(ST)=Mn(S)Mn(T)(x,y)\in M_{n}(S\oplus T)=M_{n}(S)\oplus M_{n}(T). The matrix cones just identify Mn(ST)+=Mn(S)+Mn(T)+M_{n}(S\oplus T)^{+}=M_{n}(S)^{+}\oplus M_{n}(T)^{+}.

Proposition 9.3.

The bifunctor (S,T)ST(S,T)\mapsto S\oplus T is the categorical coproduct in the category of operator systems with ccp maps as morphisms. That is, given any operator system RR and ccp maps φ:SR\varphi:S\to R and ψ:TR\psi:T\to R, the linear map φψ:STR\varphi\oplus\psi:S\oplus T\to R is ccp.

Proof.

This follows either by the explicit description of the matrix norms and order on STS\oplus T, or by showing that (K×L,(0,0))(K\times L,(0,0)) is the categorical product of (K,0)(K,0) and (L,0)(L,0) in the category of pointed compact nc convex sets, and using Theorem 2.11. ∎

Remark 9.4.

The operator space norm on STS\oplus T is neither the usual \ell^{\infty}-product nor the 1\ell^{1}-product of the operator spaces SS and TT. For example, if

K=L=n1{xMn+0x1n}K=L=\coprod_{n\geq 1}\{x\in M_{n}^{+}\mid 0\leq x\leq 1_{n}\}

is the nc simplex generated by [0,1][0,1], and aA(K,0)a\in A(K,0) is the coordinate function a(x)=xa(x)=x, then

aaA(K2,(0,0))\displaystyle\|a\oplus a\|_{A(K^{2},(0,0))} =2>aaand\displaystyle=2>\|a\oplus a\|_{\infty}\quad\text{and}
a(a)A(K2,(0,0))\displaystyle\|a\oplus(-a)\|_{A(K^{2},(0,0))} =1<aa1.\displaystyle=1<\|a\oplus a\|_{1}.
Proposition 9.5.

Let SS and TT be operator systems. If SS and TT are dualizable, then STS\oplus T is dualizable.

Proof 1.

We will use Theorem 5.7. Let the nc quasistate spaces of SS and TT be (K,0)(K,0) and (L,0)(L,0), respectively. Then (K+K)+K(K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K and (L+L)+L(L-\mathbb{R}_{+}L)\cap\mathbb{R}_{+}L are norm bounded. Checking that

(K×L+(K×L))+(K×L)((K+K)+K)×((L+L)+L)(K\times L-\mathbb{R}_{+}(K\times L))\cap\mathbb{R}_{+}(K\times L)\subseteq((K-\mathbb{R}_{+}K)\cap\mathbb{R}_{+}K)\times((L-\mathbb{R}_{+}L)\cap\mathbb{R}_{+}L)

shows that (K×L+(K×L))+(K×L)(K\times L-\mathbb{R}_{+}(K\times L))\cap\mathbb{R}_{+}(K\times L) is bounded, so STA(K×L,(0,0))S\oplus T\cong A(K\times L,(0,0)) is dualizable. ∎

It is also possible to give a proof of Proposition 9.5 using only Ng’s bounded decomposition property, which appears in 5.7.(2).

More generally, we can form finite pushouts in the operator system category by taking pullbacks in the category of pointed compact nc convex sets.

Definition 9.6.

Let

R{R}S{S}T{T}φ\scriptstyle{\varphi}ψ\scriptstyle{\psi}

be a diagram of operator systems with ccp maps as morphisms. Let SS, TT, and RR, have respective quasistate spaces (K,0)(K,0), (L,0)(L,0), and (M,0)(M,0). We define the pushout SR,φ,ψTS\oplus_{R,\varphi,\psi}T as the operator system

A(K×M,φ,ψL,(0,0)),A(K\times_{M,\varphi^{\ast},\psi^{\ast}}L,(0,0)),

where

K×M,φ,ψL={(x,y)K×Lφ(x)=ψ(y)}K×L,K\times_{M,\varphi^{\ast},\psi^{\ast}}L=\{(x,y)\in K\times L\mid\varphi^{\ast}(x)=\psi^{\ast}(y)\}\subseteq K\times L,

equipped with the natural maps

ιS:SST\displaystyle\iota_{S}:S\to S\oplus T SR,φ,ψTand\displaystyle\to S\oplus_{R,\varphi,\psi}T\quad\text{and}
ιT:TST\displaystyle\iota_{T}:T\to S\oplus T SR,φ,ψT\displaystyle\to S\oplus_{R,\varphi,\psi}T

which make the diagram

(5) R{R}S{S}T{T}SR,φ,ψT{S\oplus_{R,\varphi,\psi}T}φ\scriptstyle{\varphi}ψ\scriptstyle{\psi}ιS\scriptstyle{\iota_{S}}ιT\scriptstyle{\iota_{T}}

commute.

When the morphisms φ\varphi and ψ\psi are understood, we will usually just write SRTS\oplus_{R}T and K×MLK\times_{M}L. Note that the coproduct STS\oplus T coincides with the pushout S0TS\oplus_{0}T of the diagram

0{0}S{S}T{T}0\scriptstyle{0}0\scriptstyle{0}

as expected, where 0 denotes the 0 operator system.

To verify that A(K×ML,(0,0))A(K\times_{M}L,(0,0)) is an operator system, we need to show that:

Proposition 9.7.

(K×ML,(0,0))(K\times_{M}L,(0,0)) is pointed.

Proof.

Let ρ:A(K×ML,(0,0))Mn\rho:A(K\times_{M}L,(0,0))\to M_{n} be an nc quasistate. Pulling ρ\rho back to A(K×L,(0,0))A(K\times L,(0,0)) gives a point evaluation at some point (x,y)K×L(x,y)\in K\times L. It will suffice to show that (x,y)K×ML(x,y)\in K\times_{M}L, in which case ρ\rho must be point evaluation at (x,y)(x,y).

We must show that φ(x)=ψ(y)\varphi^{\ast}(x)=\psi^{\ast}(y) in MM. Given aRA(M,0)a\in R\cong A(M,0). Since the diagram (5) commutes, upon pulling back to STS\oplus T, we have

ρ(ιSφ(a))=(φ(a)0)(x,y)=(0ψ(a))(x,y)=ρ(ιTψ(a)),\rho(\iota_{S}\varphi(a))=(\varphi(a)\oplus 0)(x,y)=(0\oplus\psi(a))(x,y)=\rho(\iota_{T}\psi(a)),

that is, φ(a)(x)=a(φ(x))=ψ(a)(y)=a(ψ(y))\varphi(a)(x)=a(\varphi^{\ast}(x))=\psi(a)(y)=a(\psi^{\ast}(y)). Since aR=A(M,0)a\in R=A(M,0) was arbitrary, this proves φ(x)=ψ(y)\varphi^{\ast}(x)=\psi^{\ast}(y), so (x,y)K×ML(x,y)\in K\times_{M}L. ∎

Proposition 9.8.

The diagram 5 is a pushout in the category of operator systems with ccp maps as morphisms.

Proof.

It is easiest to verify that the diagram

(K×ML,(0,0)){(K\times_{M}L,(0,0))}(K,0){(K,0)}(L,0){(L,0)}(M,0){(M,0)}φ\scriptstyle{\varphi^{\ast}}ψ\scriptstyle{\psi^{\ast}}

is a pullback in the category of pointed compact nc convex sets with pointed continuous affine nc functions as morphisms, where the unlabeled maps are just the coordinate projections. Checking this is fairly immediate, using the fact that the point-weak-\ast topology on K×MLK×LK\times_{M}L\subseteq K\times L coincides with the restriction of the product topology. By the contravariant equivalence of categories Theorem 2.11, it follows that (5) is a pushout. ∎

Proposition 9.9.

If SS and TT are dualizable operator systems, then any pushout SR,φ,ψTS\oplus_{R,\varphi,\psi}T is also dualizable.

Proof.

This follows from Proposition 9.5 combined with Corollary 5.9 used with the inclusion (0,0)K×MLK×L(0,0)\subseteq K\times_{M}L\subseteq K\times L. ∎

It follows by induction that any pushout of a finite family of dualizable operator systems is again dualizable.

References